You are on page 1of 7

Biomass and Bioenergy 119 (2018) 411–417

Contents lists available at ScienceDirect

Biomass and Bioenergy


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

Research paper

Hydraulic and organic rates applied to pilot scale UASB reactor for sugar T
cane vinasse degradation and biogas generation
Valéria Del Nery∗, Inaê Alves, Márcia Helena Rissato Zamariolli Damianovic, Eduardo Cleto Pires
Department of Hydraulics and Sanitation, University of São Paulo (EESC/USP), Av. Trabalhador São-carlense, 400, 13566-590, São Carlos, SP, Brazil

A R T I C LE I N FO A B S T R A C T

Keywords: A pilot-scale upflow anaerobic sludge blanket (UASB) reactor (120 L) was operated for biogas production by
Biogas vinasse degradation. The reactor was operated for 700 days using a recirculation rate of 1:3 with organic loading
High rate reactor rates (OLR) ranging from 0.5 to 32.4 kgCOD.m−3.d−1, upflow velocities ranged from 0.008 to 0.292 m h−1(feed
Pilot-UASB reactor added with recirculation) and the hydraulic retention time ranged from 33.33 to 0.86 days. The reactor de-
Renewable energy
monstrated a stable performance at OLR at all applied loads. The COD removal efficiencies throughout the
Vinasse
experiment were 87.5 ± 5.3 and 90.5 ± 3.6% for raw and soluble COD, respectively. Total volatile acids
ranged from 59 mg L−1 to 585 mg L−1. The concentration of acetic and propionic acids at the reactor effluent
reached 28 mg L−1 on OLR 20 kgCOD.m−3.d−1. The methane content in the biogas was 68.8 ± 7.14%.The
methane yield of 0.299 ± 0.066 LCH4.g−1COD corresponds to about 76.4% of the theoretical methane yield.
The methane productivity reached the highest value of 8.059 LCH4.L−1.d−1 at 32.27 kgCOD.m−3.d−1., cor-
responding to 83.9% of the theoretical value (9.611 LCH4.L−1.d−1). This result is very promising for power
generation in the sugar and ethanol industry, complementing the energy balance of sugarcane, as shown by an
energy balance evaluation.

1. Introduction energetic potential of vinasse. Anaerobic digestion minimizes vinasse


pollution load without losing the fertilization potential and generates
The Brazilian sugar and ethanol industry has received much atten- biogas. Generating and using biogas is an important technological ad-
tion in recent years due to pioneering, significant growth and techno- vancement in the area of bioenergy production [3,4]. Renewable en-
logical development. This progress has resulted in improved energy ergy generation is a strong attraction to use anaerobic digestion of vi-
efficiency and the recovery of by-products from the sugarcane industry, nasse [5,6].
highlighting this industry as a food and energy production unit. Anaerobic reactors of different configurations and scales have been
Brazil has a worldwide presence as a major producer of sugarcane, studied extensively for anaerobic degradation of vinasse from sugar
sugar and ethanol. The 2017–18 crop produced 641.07 million tons of cane aiming to acquire knowledge and/or design parameters to im-
sugarcane, generating 38.60 million tons of sugar and 27.86 billion prove reactor technology for high generation of biogas. Fixed-film
liters of alcohol [1]. [7,8]; fluidized-bed [9,10], UASB [11,12], sequential batch [13], hy-
Taking into account the potential energy of sugar cane, 40% is brid reactors [14,15], two-stage anaerobic membrane bioreactor [16]
transformed into alcohol and 31% remain in the by-products as bagasse are some examples of anaerobic reactors applied to vinasse anaerobic
(26%) and vinasse (5%). Bagasse is used extensively to meet the energy degradation.
demand in the production process; however the energy content re- Concomitantly to the progress of studies using anaerobic digestion
maining in vinasse, which has not been exploited, is very significant in to treat vinasse, restrictive conditions are highlighted in the treatment
the energy balance. Vinasse is the by-product of the alcohol distillation capacity and in the application of this technology. Some of the limita-
process generated in a proportion of 10–18 L per liter of ethanol pro- tions concerning the technological application of the anaerobic process
duced, characterized by a high concentration of organic matter and to recover the energy content of the vinasse are mainly related to the
nutrient salts [2]. start-up of the reactor (adaptation of the biomass to the effluent to be
The anaerobic biological process of organic matter conversion is a treated) and the maintaining the active biomass in the reactor. These
suitable option for reducing both the organic load and the waste of factors are made worse by the high concentration of organic matter and


Corresponding author.
E-mail address: vdelnery@terra.com.br (V. Del Nery).

https://doi.org/10.1016/j.biombioe.2018.10.002
Received 30 March 2018; Received in revised form 19 September 2018; Accepted 1 October 2018
0961-9534/ © 2018 Elsevier Ltd. All rights reserved.
V. Del Nery et al. Biomass and Bioenergy 119 (2018) 411–417

organic load applied to the reactor, which has toxic effect and the need Table 1
for adapting biomass to by-products of anaerobic digestion and biogas Characterization of the vinasse and the anaerobic reactor effluent.
production markedly. Parameter Vinasse Effluent
Despite the efforts, the use of efficient compact anaerobic reactors to
−1
be used in large scales is far from being solved. Anaerobic degradation COD raw (mg.L ) 19220 ± 2410 2270 ± 710
COD filtered (mg.L−1) 15300 ± 1850 1440 ± 550
of vinasse on an industrial scale is a challenge and is linked to the use of
Total volatile acids (mg.L−1) 2120 ± 723 174 ± 44
scientific and technological basis for optimizing the biochemical pro- Total solids (mg.L−1) 14570 ± 3420 8830 ± 1940
cess and reactor design. Then, due to the difficulties of maintaining the Total volatile solids (mg.L−1) 9340 ± 4125 2140 ± 520
stability of high rate anaerobic process for degradation of vinasse, Total suspended solids (mg.L−1) 2210 ± 890 990 ± 302
studies addressing organic loads and hydraulic rates associated with Volatile suspended solids (mg.L−1) 1880 ± 850 710 ± 135
Acetic acid (mg.L−1) 334 ± 104 32.20 ± 0.87
methane generation should be carried out to better understand the
Propionic acid (mg.L−1) 46 ± 35 4.85 ± 7.87
phenomena of high rate anaerobic degradation of vinasse. The lack of Butyric acid (mg.L−1) 38 ± 40 12.77 ± 3.04
information makes it difficult to enable projects of high rate compact Valeric acid (mg.L−1) 15 ± 12 13.45 ± 7.73
and efficient anaerobic reactors. Ethanol (mg.L−1) 148 ± 153 0.30 ± 0.87
Sulfate (mg SO42−.L−1) 648 ± 545 46.17 ± 29.86
The current research aimed to present results from a long term and
Sulfide (mg S2−.L−1) 3.04 ± 7.34 38.36 ± 48.25
stable operation of a pilot-scale UASB reactor for which stepped organic Orthophosphate (mg PO43− L−1) 102 ± 166 48 ± 25
loads increases were associated with hydraulic rate adjustments, using Ammoniacal nitrogen (mg NH3-N.L−1) 45 ± 24 43 ± 17
recirculation, to reach high biogas production by anaerobic degradation Nitrate (mg NO3-N.L−1) 441 ± 397 86 ± 49
of vinasse. Using biogas production the potential of electric energy that Total Kjeldahl nitrogen (TKN), (mgN.L−1) 345 ± 140 219 ± 33
pH 4.20 ± 0.19 7.75 ± 0.23
could be supplied using internal gas engine coupled to generators was
Total alkalinity (mgCaCO3.L−1) – 3996 ± 1038
predicted for a model sugar processing facility. Partial alkalinity (mgCaCO3.L−1) – 3239 ± 905
Calcium (mgCa.L−1) 540 ± 213 424 ± 185
Iron (mgFe.L−1) 39 ± 24 10 ± 3
2. Material and methods
Magnesium (mgMg.L−1) 237 ± 75 192 ± 64
Sodium (mg Na.L−1) 277 ± 529 1170 ± 466
2.1. Experimental set up Potassium (mgK.L−1) 1470 ± 565 1220 ± 500

The experimental set up used was shown in Fig. 1. A pilot scale


UASB reactor with a capacity of 60 L in the reaction zone (2 m height) TG 0.5.
and total volume of 120 L (4 m total height) were constructed using The reactor was seed with 60 L of granular sludge (37 gVSS.L−1)
PVC pipes with 19.5 cm diameter. The reactor was equipped with gas- from a UASB reactor installed at a poultry slaughterhouse industry
solid separators [17] and provided with recirculation devices. Along the [18]. The volume of the granular sludge filled the named reaction zone
vertical eight sampling ports (from P1 to P8) 40 cm apart from each of the reactor. The reactor was operated at the laboratory temperature
other allowed the collection of samples for vertical profiles studies. A of 22 ± 3 C.
200 L tank for reactor effluent and vinasse addition was used as storage The organic loading rates (OLR) associated with upflow velocities
to feed the reactor. The reactor was operated using a recirculation rate were applied to the reactor in steps in order to adapt the anaerobic
of 1:3 aiming to adjust the reactor influent concentration by dilution sludge to the vinasse, aiming to reach as high as 30 kgCOD.m−3.d−1. at
with the reactor effluent and to decrease the amount of alkali needed to the reaction zone. The OLR increments were 0.3 kgCOD.m−3.d−1. They
keep the pH within the usual operational range. Sodium bicarbonate were performed with reactor efficiency above 80% for at least 7 days.
was used as alkali to adjust the pH and the alkalinity in the reactor (0.3 The reactor was operated for 700 days.
gNaHCO3.g−1COD). The biogas flux was measured using a Ritter Type
2.2. Reactor feed

The characterization of the vinasse used in this study during and the
effluent used to diluted the vinasse during the operating period is
shown in Table 1.

2.3. Monitoring analyses

The monitoring analyses were carried out according to the Standard


Methods for the Examination of Water and Wastewater [19]. Alkalinity
was carried out as described by Ripley et al. [20], Total Volatile Acids
(TVA) by titration was carried out as described by Dilallo and Albertson
[21], Volatile Fatty Acids (VFA) were determined by gas chromato-
graphy using HP 6890 equipment, with a HP-INNOVAX
(30 m × 0.25 mm x 0.25 mm) column. The measurement of biogas
composition was performed in a Shimadzu gas chromatography (GC
2014 model), with thermal conductivity detector and an HP-PLOT
column (30 m × 0.53 mm).

3. Results

The OLR applied to the reactor reaction zone ranged from 0.5 to
32.4 kgCOD.m−3.d−1. and from 0.5 to 39.0 kgCOD.m−3.d−1. (adding
the OLR related to the effluent recirculation). The upflow velocities
Fig. 1. Schematic design of the pilot-UASB reactor. applied to the reactor ranged from 0.008 to 0.292 m h−1(feed added

412
V. Del Nery et al. Biomass and Bioenergy 119 (2018) 411–417

with recirculation) and the hydraulic retention time ranged from 33.33
to 0.86 days. The results of COD removal efficiency, generation of vo-
latile acids and flow of methane will be assessed for the OLR applied.

3.1. Characterization of the vinasse and reactor effluent

The vinasse characterization over the reactor operational period


(Table 1) indicates the presence of ethanol, probably related to the
residual of the distillation column and VFA probably due to fermenta-
tion of the vinasse during the storage step. The characterization showed
relatively low nitrogen and phosphorus concentrations and high po-
tassium concentration. In terms of micronutrients, lower concentrations
were found of calcium, magnesium, iron and sulphur. The macro and
micronutrients determined in the operational period did not negatively
affect the performance and stability of the anaerobic process. Despite
the relatively low nutrient the minimum ratio 350COD: 100 N:1 P was
contemplated.
Fig. 3. Total volatile acids in the reactor effluent during the organic loading
These characteristics are close to the vinasse generated in autono-
rate applied to the reactor.
mous distilleries that use only the sugar cane juice to produce ethanol.
More concentrated vinasse is generated in the production of alcohol
from molasses or a mixture of molasses and juice. The vinasse generated considered) (Fig. 2).
in distilleries has lower concentrations of organic matter, potassium, The results indicate the success of the anaerobic process in COD
calcium and magnesium [11,22,23]. Moreover, physicochemical char- reduction of vinasse and methane generation. The recirculation con-
acteristics of sugarcane vinasse change along the season and are in- tributed to increase the upflow velocity in the reactor, which improves
trinsically related to the type of industry – sugar and alcohol (molasses), mass transport from the bulk of the flow to the granules, as is well
distillery (juice) or mixed (molasses and juice) –; the fermentation stablished in the literature on mass transfer, thus improving the effi-
method; processing efficiency; inputs added as chemicals, acids, anti- ciency of the anaerobic degradation. Higher upflow velocities may also
biotics and nitrogen and phosphorus, as well as the quality of water contribute to methane release. Besides that, the recirculation con-
used, work system and influence of the mill operators [9,24–26]. tributes to diluting the vinasse, reducing the negative effect of high
The reactor effluent characterization (Table 1) showed a high or- strength effluent and the presence of some inhibitory substances in
ganic matter reduction (88% of COD), a reduction of total solids and anaerobic process [25,27].
total suspended solids of 39.4 and 55.4%, respectively.
The anaerobic process reduced ethanol concentration and very low 3.3. Total and organic volatile acids
concentrations of organic acids were found in the effluent. Although the
anaerobic process has reduced organic and solid matter presented in The OLR applied to the reactor between 0.5 and 32 kgCOD.m−3.d−1
vinasse, there was no significant reduction of nutrients and micro nu- resulted in low concentrations of TVA (Fig. 3) that ranged from
trients concentration. This result is very important because vinasse is 59 mg L−1 to 585 mg L−1. The concentration of acetic, propionic, bu-
widely used for irrigation and fertilization for total or partial replace- tyric and valeric acids in the reactor effluent for OLR up to 20
ment of chemical fertilizers in sugarcane crops [22]. As the anaerobic kgCOD.m−3.d−1 were very low (Table 2). Although a high concentra-
process did not change the fertilizing properties, the effluent from the tion of organic matter or the complexity of the vinasse constituents
anaerobic reactor can be used to fertilize soil after energy generation. could lead to microbial inhibition by excess substrate, these results
confirm the success and stability of an anaerobic process of vinasse at
3.2. COD removal high organic loads. The strategy of effluent recirculation may have
stimulated an environment suitable for the degradation of organic
The COD removal efficiencies throughout the experiment were matter which promotes dilution of vinasse in the influent of the reactor.
87.5 ± 5.3 and 90.5 ± 3.6% for raw COD and soluble COD, respec- The production of organic acids depends primarily on the characteristic
tively. There is an almost perfect correlation between the applied and of the wastewater apart from OLR applied [28]. Another important
removed OLR (r2 = 0.992 without consideration of the load resulting factor may be a suitable stepwise increase in the OLR applied to the
from recirculation and r2 = 0.9576 when the load from recirculation is reactor over the operational period to achieve high OLR such as 20 to

Fig. 2. Organic loading rate removed as a function of organic loading rate (without recirculation) (□) and organic loading rate (with recirculation) (♦) applied to the
reactors.

413
V. Del Nery et al. Biomass and Bioenergy 119 (2018) 411–417

Table 2 greatly increase the peak methane content are in the range of 0.5–5
Volatile fatty acids in the reactor effluent at the organic loading rate applied to kgCOD.m−3.d−1 (Fig. 6), meaning that the volume of the reactor would
the reactor. be at least twice the volume of the reactor designed to operate under 10
OLR (kgCOD.m−3.d−1) Acid concentration (mg.L−1) kgCOD.m−3.d−1. At the same time, although peak production increases
to up to 80%, the median methane content would be the same, as seem
Acetic Propionic Butyric Valeric in Fig. 6. Thus the capital expenditure (CAPEX) would increase because
of doubling the reactor volume without a significant gain in the volume
5 18.4 < 0.2 18.0 15.10
7.5 25.7 < 0.2 8.1 < 0.2 of methane produced. Another reason to choose 10 kgCOD.m−3.d−1 is
10 56.5 7.2 12.0 < 0.2 the consistency of methane production at this load as the narrower
12.5 29.7 < 0.2 11.4 < 0.2 range indicates (Fig. 6). The methane content used in this estimate
15 22.7 0.3 10.5 < 0.2
ranged from 60 to 68%, with a median of 65% (Fig. 6). These values are
17.5 45.4 21.5 14.1 < 0.2
20 27.8 27.8 15.3 < 0.2 consistent with other studies that used vinasse [7]. Concerning the
methane yield, according to the current experiment, it was obtained
0.299 ± 0.066 LCH4.g−1CODremoved, at 25 °C and 1 atm, which
32 kgCOD.m−3.d−1, avoiding an accumulation of long-chain volatile corresponds, on average, to about 76% of the theoretical methane yield,
acids and keeping the favourable thermodynamic conditions for the that is 0.391 ± 0.004 LCH4.g−1CODremoved (Fig. 7a).
balanced development of the anaerobic process. Low concentration of The methane productivity increased linearly with the stepped in-
propionic acid also confirms the suitability of an anaerobic environ- crease of OLR removed (Fig. 7b). A regression analysis resulted in a
ment developed in the reactor reaction zone [29]. correlation coefficient (Pearson's r) equal to 0.84, while the linear re-
The success of anaerobic digestion of vinasse under high OLR de- gression indicates that for the highest OLR applied – 32.3
pends on the balance of the microbial consortium interactions present kgCOD.m−3.d−1 –a methane yield between 6.4 and 9.6
in the reactor sludge. LCH4.L−1reactor.d−1 is expected. It should be mentioned that the
highest value corresponds to the theoretical methane yield, that is, a
3.4. Characterization of reactor vertical profile yield that cannot be surpassed. The most probable yield for this load is
8.63 LCH4.L−1reactor.d−1, or 5.66 gCH4.L−1reactor.d−1. This result
The evaluation of organic matter consumption in the reactor profile compares favourably with the value obtained by Uellendahl and Ahring
indicated that the most significant organic matter reduction occurred in [30] who achieved 1.2 LCH4.L−1reactor.d−1 using a pilot-scale UASB
P1 and P2, regardless of the OLR applied to the reactor, corresponding reactor under mesophilic conditions treating high strength wastewater
to 24 L (40% of the reaction zone and 20% of the total volume of the from second generation bioethanol production. The mentioned authors
reactor). operated the reactor up to an OLR of 13.2 kgCOD.m−3.d−1.
While the overall COD soluble removal efficiency throughout the Thus, assuming the same conversion efficiency in terms of methane
experiment was 90.5 ± 3.6%, the removal at the first 40% of the re- production per gram of removed COD obtained in this work, a simple
action zone was 88.5 ± 3.7% (Fig. 4). Concerning the OLR applied to mass and energy balance provides an estimate of the amount of elec-
the reactor, the highest concentration of acetic and propionic acids trical energy that can be supplied by an ethanol plant. As an example,
occurred in P1 (438 ± 156 and 307 ± 124 mg L−1, respectively) the vinasse generation at the plant that supplied the vinasse for this
while in P2 the concentration of these acids was 35.8 ± 29.9 and research was 1,820,000 m3 during the 2014–2015 crop. For this
15.7 ± 24.8 mg L−1. In the reactor effluent the concentration of acetic amount of vinasse a volume of 7000 m3 for the anaerobic reaction zone
and propionic acids dropped to 32.20 ± 13.7 and operating at 32 kgCODm−3.d−1 is needed, as in this experiment. This
4.85 ± 8.62 mg L−1, respectively (Fig. 5). These results confirm the reactor would produce 39,600 kgCH4.d−1, or at least 29,400
efficiency of the transformation of complex organic compounds by high kgCH4.d−1 if it operates providing the lowest expected methane yield.
rate anaerobic reactors. Moreover, these results can contribute to car- Thus, using Equation (1), the amount of electrical energy can be esti-
rying out projects of high rate compact and efficient anaerobic reactors mated.
to treat vinasse. Eel = 0.2778 × Hc × MCH4 × ηconv (1)

3.5. Methane yield, productivity and electrical energy recovery potential In Equation (1), Eel represents the amount of electrical energy, Hc is
the calorific power of methane, MCH4 is the mass of methane burned
For the purpose of estimating the energy recovery potential, the and ηconv is the efficiency of converting the chemical energy of methane
authors used only results for organic loading rates higher than 10 into electrical energy and 0.2778 is the conversion factor from MJ to
kgCOD.m−3.d−1. Although, as seen in Fig. 6, lower OLRs provided kWh. In this estimation, ηconv was set to 43.7% per data from Caterpillar
higher methane content, the authors consider that it is not economical (2016) and Hc is 50.6 MJkg−1 (The Engineering Toolbox, 2016). Sub-
to operate anaerobic reactors with such low loads. Loads that would stituting these numbers in Equation (1), Eel results from 180,600 to

Fig. 4. COD concentration along the reactor vertical. For better visualization, the results for 7.5; 12.5 and 17.5 kgCOD.m3.d−1 are not shown.

414
V. Del Nery et al. Biomass and Bioenergy 119 (2018) 411–417

Fig. 5. Volatile fatty acids in the vertical profile of the reactor during the organic loading rate increase applied to the reactor.

243,250 to kWh per day, or 5,418,000 to 7,297,500 kWh per month,


during the cropping season. Data from 2004 to 2015, for home energy
consumption in the southeast of Brazil, where the sugar mill is located,
indicate that the average yearly consumption is 172 kWh per month per
house. Thus, this plant could provide energy for at least 31,500 houses,
if the lower methane yield is considered, or 42,400 houses if the most
probable yield is considered. Considering power generation, the result
from the current work is promising and would help to complement the
energy balance of a sugarcane mill.
It should be mentioned that biogas needs to be cleaned before it can
be supplied to internal combustion engines as even small amounts of
H2S, which is normally found in biogas from anaerobic digestion, can
cause severe corrosion in pipelines and in the engine itself. However,
there are clean-up technologies that require low amounts of energy and
capital costs and can be used in such cases [31,32]. A complete as-
sessment of these costs is beyond the scope of this paper and the reader
is advised to access the cited work for further information.

Fig. 6. Content of methane in biogas over organic loading rates applied to the
reactor. 4. Conclusion

Anaerobic reactors to process vinasse for methane production are


viable and can operate under high organic loading rates, above 32.4

415
V. Del Nery et al. Biomass and Bioenergy 119 (2018) 411–417

Fig. 7. Methane yield during the organic loading rates ap-


plied to the reactor (a) and methane productivity during the
organic loading rates applied to the reactor (b) (◊) theoretical
(■) reactor.

kgCOD.m−3.d−1, during long term operation as shown in this research, ethanol production in Brazil: challenges and perspectives, Renew. Sustain. Energy
when the reactors were continuously operated during 700 days. The Rev. 44 (2015) 888–903, https://doi.org/10.1016/j.rser.2015.01.023.
[6] L.T. Fuess, B.C. Klein, M.F. Chagas, M.C.A.F. Rezende, M.L. Garcia, A. Bonomi,
methane yield and production can reach values above 75% of the M. Zaiat, Diversifying the technological strategies for recovering bioenergy from the
theoretical values. The long term reactor stability of organic matter two-phase anaerobic digestion of sugarcane vinasse: an integrated techno-economic
reduction and methane production pointed out the benefits of the op- and environmental approach, Renew. Energy 122 (2018) 674–687, https://doi.org/
10.1016/j.renene.2018.02.003.
erational strategies in terms of vinasse dilution by effluent recirculation [7] A. Bories, J. Raynal, F. Bazile, Anaerobic digestion of high-strength distillery was-
and very small increments of OLR during start-up. The use of methane tewater (cane molasses stillage) in a fixed-film reactor, Biol. Wastes 23 (1988)
to produce electricity is justifiable as it was shown that, considering the 251–267, https://doi.org/10.1016/0269-7483(88)90014-6.
[8] T. Nandy, S. Shastry, S.N. Kaul, Wastewater management in a cane molasses dis-
energy consumption pattern of the southeast of Brazil, a large ethanol tillery involving bioresource recovery, J. Environ. Manag. 65 (2002) 25–38,
plant can produce sufficient energy to supply power to more than https://doi.org/10.1006/jema.2001.0505.
31,000 houses during the sugar cane harvesting season. [9] N. Fernández, S. Montalvo, R. Borja, L. Guerrero, E. Sánchez, I. Cortés,
M.F. Colmenarejo, L. Travieso, F. Raposo, Performance evaluation of an anaerobic
fluidized bed reactor with natural zeolite as support material when treating high-
Acknowledgements strength distillery wastewater, Renew. Energy 33 (2008) 2458–2466, https://doi.
org/10.1016/j.renene.2008.02.002.
The authors acknowledges National Council for Scientific and [10] L.M. Siqueira, E.S.G. Damiano, E.L. Silva, Influence of organic loading rate on the
anaerobic treatment of sugarcane vinasse and biogás production in fluidized bed
Technological Development (CNPq - Conselho Nacional de reactor, J. Environ. Sci. Heal. - Part A Toxic/Hazardous Subst. Environ. Eng. 48
Desenvolvimento Científico e Tecnológico, Brazil) 04394/2009-2, (2013) 1707–1716, https://doi.org/10.1080/10934529.2013.815535.
158721/2012-8 and 142211/2015-0, São Paulo Research Foundation [11] F.J.C.B. Costa, B.B.M. Rocha, C.E. Viana, A.C. Toledo, Utilization of vinasse ef-
fluents from an anaerobic reactor, Water Sci. Technol. 18 (1986) 135–141.
(Fapesp-Fundação de Amparo à Pesquisa doEstado de São Paulo, Brazil) [12] A.D.N. Ferraz Júnior, M.H. Koyama, M.M. de Araújo Júnior, M. Zaiat, Thermophilic
2013/17591-1, for supporting the development of this study, and Rio anaerobic digestion of raw sugarcane vinasse, Renew. Energy 89 (2016) 245–252,
Pardo S/A for providing vinasse for this research. https://doi.org/10.1016/j.renene.2015.11.064.
[13] a. P. Miqueleto, C.C. Dolosic, E. Pozzi, E. Foresti, M. Zaiat, Influence of carbon
sources and C/N ratio on EPS production in anaerobic sequencing batch biofilm
References reactors for wastewater treatment, Bioresour. Technol. 101 (2010) 1324–1330,
https://doi.org/10.1016/j.biortech.2009.09.026.
[14] G.S. Kumar, S.K. Gupta, G. Singh, Biodegradation of distillery spent wash in
[1] UNICA - União da Indústria de Cana-de-Açúcar - por safra, (n.d). http://www.
anaerobic hybrid reactor, Water Res. 41 (2007) 721–730, https://doi.org/10.1016/
unicadata.com.br/historico-de-producao-e-moagem.php?idMn=32&
j.watres.2006.11.039.
tipoHistorico=4&acao=visualizar&idTabela=1984&safra=2017%252F2018&
[15] C.B. Shivayogimath, T.K. Ramanujam, Treatment of distillery spentwash by hybrid
estado=RS%252CSC%252CPR%252CSP%252CRJ%252CMG%252CES%252CMS
UASB reactor, Bioprocess Eng. 21 (1999) 255–259, https://doi.org/10.1007/
%252CMT%252CGO%252CDF%252CBA%252CSE%252CAL%252CPE%252CPB
s004490050673.
%252CRN%252CCE%252CPI%252CMA%252CTO%252CPA%252CAP%25 (ac-
[16] F.S. Santos, B.C. Ricci, L.S. França Neta, M.C.S. Amaral, Sugarcane vinasse treat-
cessed September 17, 2018).
ment by two-stage anaerobic membrane bioreactor: effect of hydraulic retention
[2] A.C. van Haandel, Integrated energy production and reduction of the environmental
time on changes in efficiency, biogas production and membrane fouling, Bioresour.
impact at alcohol distillery plants, Water Sci. Technol. 52 (2005) 49–57.
Technol. 245 (2017) 342–350, https://doi.org/10.1016/j.biortech.2017.08.126.
[3] B.S. Moraes, T.L. Junqueira, L.G. Pavanello, O. Cavalett, P.E. Mantelatto,
[17] P.F.F. Cavalcanti, E.J.S. Medeiros, J.K.M. Silva, A. van Haandel, Excess sludge
A. Bonomi, M. Zaiat, Anaerobic digestion of vinasse from sugarcane biorefineries in
discharge frequency for UASB reactors, Water Sci. Technol. 40 (1999) 211–219,
Brazil from energy, environmental, and economic perspectives: profit or expense?
https://doi.org/10.1016/S0273-1223(99)00628-9.
Appl. Energy 113 (2014) 825–835, https://doi.org/10.1016/j.apenergy.2013.07.
[18] V. Del Nery, M.H.Z. Damianovic, R.B. Moura, E. Pozzi, E.C. Pires, E. Foresti, Poultry
018.
slaughterhouse wastewater treatment plant for high quality effluent, Water Sci.
[4] K.C. Surendra, D. Takara, A.G. Hashimoto, S.K. Khanal, Biogas as a sustainable
Technol. 73 (2016) 309–316, https://doi.org/10.2166/wst.2015.494.
energy source for developing countries: opportunities and challenges, Renew.
[19] APHA/AWWA/WEF, Standard methods for the examination of water and waste-
Sustain. Energy Rev. 31 (2014) 846–859, https://doi.org/10.1016/j.rser.2013.12.
water, Stand. Methods. (2012) 541 ISBN: 9780875532356.
015.
[20] A.L.E. Ripley, W.C. Boyle, J.C. Converse, Alkalimetric Improved for anaerobic di-
[5] B.S. Moraes, M. Zaiat, A. Bonomi, Anaerobic digestion of vinasse from sugarcane
gestion wastes monitoring of, Water Pollut, Contron Fed 58 (1986) 406–411,

416
V. Del Nery et al. Biomass and Bioenergy 119 (2018) 411–417

https://doi.org/10.2307/25042933. 1016/j.biortech.2012.12.126.
[21] R. DiLallo, O. Albertson, Volatile acids by diect titration, Water Pollut Control Fed [27] Y. Chen, J.J. Cheng, K.S. Creamer, Inhibition of anaerobic digestion process: a re-
33 (1961) 356–365, https://doi.org/10.2307/25034391. view, Bioresour. Technol. 99 (2008) 4044–4064, https://doi.org/10.1016/j.
[22] L.T. Fuess, M.L. Garcia, Implications of stillage land disposal: a critical review on biortech.2007.01.057.
the impacts of fertigation, J. Environ. Manag. 145 (2014) 210–229, https://doi.org/ [28] P. Elefsiniotis, D.G. Wareham, M.O. Smith, Effect of a starch-rich industrial was-
10.1016/j.jenvman.2014.07.003. tewater on the acid-phase anaerobic digestion process, Water Environ. Res. 77
[23] C.A. Christofoletti, J.P. Escher, J.E. Correia, J.F.U. Marinho, C.S. Fontanetti, (2005) 366–371.
Sugarcane vinasse: environmental implications of its use, Waste Manag. 33 (2013) [29] Q. Wang, M. Kuninobu, H.I. Ogawa, Y. Kato, Degradation of volatile fatty acids in
2752–2761, https://doi.org/10.1016/j.wasman.2013.09.005. highly efficient anaerobic digestion, Biomass Bioenergy 16 (1999) 407–416,
[24] A.C. Wilkie, K.J. Riedesel, J.M. Owens, Stillage characterization and anaerobic https://doi.org/10.1016/S0961-9534(99)00016-1.
treatment of ethanol stillage from conventional and cellulosic feedstocks, Biomass [30] H. Uellendahl, B.K. Ahring, Anaerobic digestion as final step of a cellulosic ethanol
Bioenergy 19 (2000) 63–102, https://doi.org/10.1016/S0961-9534(00)00017-9. biorefinery: biogas production from fermentation effluent in a UASB reactor - pilot-
[25] Y. Satyawali, M. Balakrishnan, Wastewater treatment in molasses-based alcohol scale results, Biotechnol. Bioeng. 107 (2010) 59–64, https://doi.org/10.1002/bit.
distilleries for COD and color removal: a review, J. Environ. Manag. 86 (2008) 22777.
481–497, https://doi.org/10.1016/j.jenvman.2006.12.024. [31] M.D. Ong, R.B. Williams, S.R. Kaffka, Comparative Assessment of Technology
[26] T. Onodera, S. Sase, P. Choeisai, W. Yoochatchaval, H. Sumino, T. Yamaguchi, Options for Biogas Clean ‐ up, (2014).
Y. Ebie, K. Xu, N. Tomioka, M. Mizuochi, K. Syutsubo, Development of a treatment [32] Capital Costs for Biogas, (2016) http://cleanleap.com/6-biogas/61-capital- costs-
system for molasses wastewater: the effects of cation inhibition on the anaerobic biogas , Accessed date: 19 November 2016.
degradation process, Bioresour. Technol. 131 (2013) 295–302, https://doi.org/10.

417

You might also like