Buscar en
Brazilian Journal of Microbiology
Toda la web
Inicio Brazilian Journal of Microbiology Ethanol production in Brazil: a bridge between science and industry
Journal Information
Vol. 47. Issue S1.
Pages 64-76 (December 2016)
Share
Share
Download PDF
More article options
Visits
5676
Vol. 47. Issue S1.
Pages 64-76 (December 2016)
Biotechnology and Industry Microbiology
Open Access
Ethanol production in Brazil: a bridge between science and industry
Visits
5676
Mario Lucio Lopes
Corresponding author
mario@fermentec.com.br

Corresponding author at: Av. Antonia Pazinatto Sturion, 1155 Piracicaba, Brazil.
, Silene Cristina de Lima Paulillo, Alexandre Godoy, Rudimar Antonio Cherubin, Marcel Salmeron Lorenzi, Fernando Henrique Carvalho Giometti, Claudemir Domingos Bernardino, Henrique Berbert de Amorim Neto, Henrique Vianna de Amorim
Fermentec, Piracicaba, São Paulo, SP, Brazil
This item has received

Under a Creative Commons license
Article information
Abstract
Full Text
Bibliography
Download PDF
Statistics
Figures (5)
Show moreShow less
Tables (3)
Table 1. Main characteristics of ethanol production in the USA and Brazil.
Table 2. Comparison of continuous and fed-batch fermentations with cell recycling for ethanol production in Brazil.37
Table 3. Increase of ethanol concentration in wine and reduction of vinasse volumes for a distillery with an ethanol production of 5200m3/week.
Show moreShow less
Abstract

In the last 40 years, several scientific and technological advances in microbiology of the fermentation have greatly contributed to evolution of the ethanol industry in Brazil. These contributions have increased our view and comprehension about fermentations in the first and, more recently, second-generation ethanol. Nowadays, new technologies are available to produce ethanol from sugarcane, corn and other feedstocks, reducing the off-season period. Better control of fermentation conditions can reduce the stress conditions for yeast cells and contamination by bacteria and wild yeasts. There are great research opportunities in production processes of the first-generation ethanol regarding high-value added products, cost reduction and selection of new industrial yeast strains that are more robust and customized for each distillery. New technologies have also focused on the reduction of vinasse volumes by increasing the ethanol concentrations in wine during fermentation. Moreover, conversion of sugarcane biomass into fermentable sugars for second-generation ethanol production is a promising alternative to meet future demands of biofuel production in the country. However, building a bridge between science and industry requires investments in research, development and transfer of new technologies to the industry as well as specialized personnel to deal with new technological challenges.

Keywords:
Ethanol
Yeast
Fermentation
Distillery
Full Text
Introduction

The automobile was one of the most impressive inventions that changed the life style of humanity. The first cars were designed to run on ethanol. However, due to the discovery of new oil fields, its abundance and low prices at the beginning of century 20, engines were being modified, giving preference to petroleum based-fuels. Then, gasoline became the primary fuel option for cars worldwide. Nevertheless, because of 1970's oil crisis, the Brazilian government launched the “Proalcool” program in 1975 with the aim to reduce the country's dependence on oil imports.1 In the first phase of the program, ethanol was added to gasoline. After the second oil crisis in 1979, the automobile industry started to produce the first car to run on ethanol only for the Brazilian market.2

In 1980, production of light vehicles that ran on ethanol reached 95% of all fleet produced in Brazil. It increased ethanol consumption and reduced significantly oil dependence. However, a combination of factors involving oil price drops, reduction of subsidies to producers and rise of sugar prices contributed to fuel shortage that led to a major downturn in the demand for ethanol-run cars.2

At the beginning of 21st century, the use of ethanol as fuel was resumed chiefly motivated by high oil prices in the international market and the development of flex-fuel technology.1 Flex-fuel cars can run either on 100% hydrous ethanol or on different blends of ethanol and gasoline. Electronic sensors detect the fuel blending and automatically adjust the engine combustion. Furthermore, the use of ethanol blends replaced toxic additives in gasoline, such as tetraethyl lead and oil-derived benzene. Leaded gasoline was banned in several countries due to serious contamination problems to the environment and humans.3 Moreover, the use of ethanol blends has improved air quality in large urban centers, reducing emissions of carbon monoxide from 50g/km driven to less than 5.8g/km driven.4 Nowadays, gasoline contains a blending of 27% of anhydrous ethanol.

The flex-fuel car technology allowed the consumer to choose for the most convenient and cheaper fuel. The production and selling of flex fuel cars and light commercial vehicles represented 85.45% of the total in 2015.5 Currently, the use of ethanol as biofuel in Brazil has been the most successful program to replace fossil fuels worldwide. Several countries have started their own programs for production and use of ethanol as fuel to reduce oil dependence and GHG (greenhouse gases) emissions.6–8

Ethanol can be produced either by chemical or microbiological processes. The chemical route is based on ethylene hydration while the microbiological process is chiefly carried out by yeast Saccharomyces cerevisiae, although other microorganisms may also produce ethanol.9–11

Currently, the main industrial route used for ethanol production worldwide is the microbiological process, also referred as alcoholic or ethanolic fermentation.12–15 During this process, sugars are converted into ethanol, energy, cellular biomass, CO2 and other byproducts by yeast cells. These sugars may come from different feedstocks and crop wastes.16,17 In Brazil, the main feedstock is sugarcane while the United States of America (USA) produces ethanol from corn.16,18 Sugarcane, corn and sorghum are C4 plants with high efficiency to convert atmospheric CO2 and water into sugars and polymers such as starch, cellulose and hemicellulose through photosynthesis. This process uses sun light energy to fix carbon and release oxygen into the air.19 Then, all CO2 resulting from ethanol burning is recycled through photosynthesis (Fig. 1).

Fig. 1.

Through photosynthesis, sugarcane converts carbon dioxide (CO2), water (H2O) and energy from light into sugars and polysaccharides releasing oxygen (O2) into atmosphere. After harvesting and crushing, sugars are fermented by yeasts and converted into ethanol. It is separated by the distillation and used as biofuel by cars. The burning of ethanol generates CO2 that returns to the atmosphere closing the cycle.

(0.17MB).

Ethanol produced from sugarcane has reduction rates between 40 and 62% in GHG emissions compared to gasoline.20 Readily fermentable sugars such as sucrose, glucose and fructose are directly converted into ethanol during alcoholic fermentation by the yeast S. cerevisiae while starch, cellulose and hemicellulose need to be hydrolyzed to simple sugars to be fermented.21,22 In both cases, yeast cells convert a solid substrate into a liquid fuel. Moreover, to convert sugars into ethanol, yeasts charge only 7% of the energy contained in sugar molecules. The end balance shows that 93% of the energy present in sugars is conserved in ethanol.23 In addition, ethanol is a molecule that can be easily separated from water through distillation process because its boiling point (78°C) is very different from water (100°C). Moreover, ethanol production from sugarcane in Brazilian distilleries presents a highly positive energy balance. In terms of equivalence, for each unit of fossil energy consumed during its production, 9.3 units of renewable energy were produced in 2005, but with potential to reach 11.6 in 2020.6 Moreover, for second-generation ethanol processes, energy balance for ethanol production from cellulosic materials is expected to be better than that of sugarcane or corn.24 All these characteristics turned the Brazilian distilleries a renewable and environmental friendly biofuel industry.25

Several scientific studies in the microbiological field have been crucial for evolution of the ethanol industry in Brazil, namely the improvement and development of new fermentation processes, selection of robust industrial yeast strains, better control of bacterial contaminants, improvement of chemical and microbiological control at distilleries.26–29 Moreover, investments in research on second-generation ethanol have stimulated the participation of some companies to produce ethanol from sugarcane bagasse and vegetal trash. Nowadays, new scientific and technological frontiers are faced by the industry to improve the fermentation process for ethanol production in Brazil.

The main alcoholic fermentation processes in the world

Nowadays, the USA and Brazil are the world's largest ethanol producers. Together, both countries account for more than 94 billion liters of ethanol produced per year, accounting for around 85% of worldwide production.30 However, there are huge differences in the fermentation processes (Table 1). Besides feedstock, a primary difference is the recycling of yeast cells. Brazilian distilleries use an improved fermentation process patented in 1937 by Firmino Boinot from the Melle region, France.31 After the end of each fermentation, the raw wine is centrifuged to separate the yeast cells in a concentrated cream while the wine goes to distillation. After a treatment with water-diluted sulfuric acid (pH 2.0–2.5 for 1–2h), these yeast cells return to large-volume tanks (250–3000L) for a new fermentation cycle.21,26,27 Distilleries in the USA do not recycle yeast cells due to high concentration of solids and all fermented media (including yeast cells) is distilled.32 Then, yeast cells are not recycled and more sugar is deviated for multiplication of cells rather than to ethanol production. Fermentation yield is lower in processes without recycling of cells.33

Table 1.

Main characteristics of ethanol production in the USA and Brazil.

Characteristics  USA  Brazil 
Main feedstock  Corn  Sugarcane 
Fermentation process  Without yeast recycling  With yeast recycling 
Solids in suspension  >30%  <1% 
Fermentation yield  85–90%  90–92% 
Fermentation time  45–606–12
Yeast concentration  3–4%  8–12% 
Ethanol concentration  12–18% (v/v)  7–12% (v/v) 
Days of plant operation  345  200–240 
Main by-product/residue  DDGS for animal feed  Vinasse for ferti-irrigation 

Once the recycling processes start with higher cell concentrations (8–12%, v/v), fermentation times of Brazilian distilleries are faster (6–12h) when compared to fermentations in USA (54–72h) without yeast cell recycling. However, corn fermentations reach higher ethanol concentrations (12–18%, v/v) in comparison to Brazilian processes (7–12%, v/v). Once corn grains can be stored for several weeks, fermentations can be carried out during 345 days a year while Brazilian distilleries run their processes for 200–240 days. Differently from corn grains, sugarcane needs to be crushed soon after harvesting to avoid loss caused by microbial contamination. In addition, Brazilian distilleries are subject to interruptions of sugarcane harvesting, crushing and fermentation due to rains.1,21

The main co-product from corn fermentations is DDGS (distillers dried grains with solubles) used for animal feed and has a high market value due to nutritional characteristics.34 On the other hand, Brazilian distilleries generate huge volumes of potassium-rich vinasse, which is used for sugarcane ferti-irrigation, reducing costs on imports of chemical fertilizers.35

Brazilian distilleries present several differences from one to another in terms of the fermentation process and sugarcane must composition. Distilleries attached to sugar factories usually ferment sugarcane musts of molasses diluted with water or a mix of juice and molasses while autonomous distilleries ferment only juice.36

Concerning fermentation, fed-batch processes account for around 83% of the distilleries while continuous fermentations are 17%.37 Despite higher investments for installation, fed-batch processes have presented better results than continuous fermentation (Table 2). A study carried out during eight consecutive years with 62 distilleries (51 fed-batch and 11 continuous fermentation) showed that fed-batch processes reached the highest fermentation yields. In addition, it was observed a lower bacterial contamination in wine and lower consumption of chemicals such as sulfuric acid and antibiotics in comparison to continuous processes.37 However, a misconducted fed-batch process will be worse than a well-managed continuous fermentation. Sometimes, the benefits of a continuous fermentation are masked by improper engineering conception and low cost adaptations.38

Table 2.

Comparison of continuous and fed-batch fermentations with cell recycling for ethanol production in Brazil.37

Characteristics  Unit  Continuous  Fed-batch 
Fermentation yield  87.0–89.5  88.9–90.5 
Bacterial contamination of winea  bacterial rods/mL  3.8–9.9×107  2.0–3.9×107 
Consumption of sulfuric acid  g/L ethanol  8–14  6–7 
Consumption of antibiotics  mg/L ethanol  5.7–13.0  3.8–8.0 
Consumption of antifoaming  mg/L ethanol  0.47–0.75  0.45–0.70 
a

Living bacterial rods are counted by optical microscopy.

In fed-batch fermentations, each tank is filled, managed and cleaned separately from the other, while in continuous processes all tanks ferment simultaneously in line and cannot be cleaned with the same frequency. After centrifugation, the yeast cream receives an acidic treatment and returns to fermentation tanks (Figs. 2 and 3).

Fig. 2.

Simplified drawing of a continuous fermentation process currently adopted by Brazilian distilleries.

(0.2MB).
Fig. 3.

Simplified drawing of a fed-batch fermentation process with the recycling of yeast cells currently adopted by Brazilian distilleries.

(0.22MB).
Fermentation: a living process

Although Eduard Buchner showed that alcoholic fermentation could be carried out with extract of dead yeast,39 the Melle-Boinot fermentation depends on living cells.31 These cells must be alive at the end of fermentation and after the acidic treatment for reuse in the next cycle. Typical Brazilian distilleries run 400 fermentation cycles during the sugarcane harvesting season. Because they are recycled several times (two cycles a day), yeast cells are subject to stressful conditions of industrial fermentations (Fig. 4). These conditions may change from one distillery to another, as well as, in the same distillery during the sugarcane harvesting season. Moreover, yeast cells are subject to effects of two or more stressing conditions.40,41

Fig. 4.

Stressing factors that affect yeast cells in industrial fermentations for ethanol production and common mechanisms of cell defense based on trehalose, glycogen, succinic acid and glycerol.

(0.15MB).

However, yeast cells have mechanisms of stress response to industrial conditions of fermentation and acid treatment.42,43 Trehalose and glycogen are reserve carbohydrates for additional energy when cells are starved for sugars.44,45 Moreover, trehalose is a protecting compound to membrane that helps cells to tolerate dehydration and high ethanol concentrations as well as other industrial stressing factors.46,47 Succinic acid in combination with ethanol has an antibacterial effect,48 while glycerol is a regulator for osmotic shocks.49,50

According to the Gay Lussac equation, theoretically, fermentation produces 511g of ethanol and 489g of CO2 from 1000g of glucose. However, in industrial processes this fermentation yield is not achieved because yeast cells drive sugars to the production of cellular biomass and secondary compounds such as glycerol, succinic, malic and acetic acids, fusel oil and other minor by-products. For this reason, the maximum fermentation yield achieved is 92–93%.28 Moreover, fermentation yield is highly dependent on the yeast strain. A fermentation carried out with cell recycles in bench scale showed that baker's yeast had a fermentation yield 3.9% lower than that of industrial strains.26 It seems little, but it means 8.5 million liters of ethanol during 200 working days for an autonomous distillery that crushes 14,000 sugarcane tons daily.

Monitoring and selection of industrial and tailored-yeast strains

Before 1990, identification and monitoring of industrial strains belonging to S. cerevisiae were often ambiguous and uncertain when performed by cell and colony morphology, physiological and biochemical methodologies.51 Morphology of cells and colonies is quite simple to distinguish strains and different strains can frequently present the same morphology. On the other hand, the same strain can change its cell and colony morphology in response to stress conditions or nutrient starvation and overexpression of genes related to cellular dimorphism like PHD1.50,52 Several strains of Saccharomyces can alternate their multiplication pattern from single cell budding to pseudohyphae, when the buds remain attached to mother cells.53 Physiological and biochemical traits are based on the ability to assimilate and/or ferment different kinds of sugars, the use of nitrogen sources, tolerance to inhibitors, hydrolysis of complex molecules, growth at different temperatures and other conditions that may allow to distinguish different strains from others. However, all these methodologies have shown limitations to identify and monitor yeast populations in practice.

During several years, the dynamic of yeast populations in industrial processes of alcoholic fermentation in Brazil remained unknown and unexplored due to difficulties for strain identification by traditional methodologies based on morphological and physiological traits. While some distilleries claimed that the baker's yeasts, M300A and IZ1904, used as starter strains were very good fermenters, others had an opposite opinion due to low fermentation yields.54 Due to lack of proper methodologies to identify and monitor starter and wild strains, it was not possible to follow changes in population during successive recycles. For this reason, when the initial yeast population was replaced by wild Saccharomyces with good characteristics, the people believed that the strain used as starter was superior. On the other hand, when wild strains with poor fermentation abilities contaminated the process, starter yeasts were appointed as responsible for problems at the distillery. The advent of molecular techniques based on DNA analyses allowed to use differences in genome as molecular markers for strain identification, monitoring and selection.

After 1990, several laboratories have introduced molecular techniques to identify new yeast strains from industrial fermentations worldwide. One methodology was the electrophoretic karyotyping based on chromosome-length polymorphism of yeast cells.55–57 After a pulsed-field gel electrophoresis to separate large DNA molecules, yeast strains show differences in chromosome profile regarding number and size that make each karyotype unique and strain-specific.58–60 This technique had been used to identify and monitor yeast strains in grape fermentations with excellent results.56 In 1990, after a visit to INRA, Montpelier, France, Professor Luiz Carlos Basso (ESALQ-USP) introduced the electrophoretic karyotyping to identify, monitor and select industrial yeast strains based on chromosome polymorphism. Karyotyping showed that the baker's yeast, IZ1904, and laboratory strains do not survive successive fermentation cycles.56 After three or four weeks, these strains are replaced by wild Saccharomyces that contaminate the industrial process of alcoholic fermentation.26 However, several years were necessary to convince people from distilleries that baker's yeast, IZ1904 and laboratory strains do not survive in industrial fermentations with cell recycling and are quickly replaced by wild Saccharomyces.61 In addition, karyotyping has shown that PE2 and CAT1, used as starter strains, were able to replace baker's yeast even at proportions as low as 0.5kg for 12tons of pressed baker's yeast.26 Nowadays, there are only six industrial strains of major importance for ethanol production in Brazil: PE2, CAT1, FT858L and Fermel® (selected by Fermentec), BG1 and SA1 (selected by CTC). These strains have been used by Brazilian distilleries that account for 70% of all ethanol production in the country.

Industrial yeast strains have more complexes and heterogeneous chromosomes than laboratory strains do.62 The high polymorphism is unique for each strain. However, these chromosomes are subject to mitotic and meiotic recombination as well as segregation errors during successive mitoses under stress conditions or after sporulation and conjugation of spores.58,59,63 The introduction of a simplified methodology for the mitochondrial DNA analysis allowed to distinguish variants from starters and contaminating strains.64 The combination of karyotyping and fragment length polymorphism of mitochondrial DNA has proven to be a powerful tool to monitor yeast population diversity in industrial fermentations as well as to select new strains and variants from starter yeasts that become more adapted to conditions of each distillery in particular.61 Variability of mitochondrial DNA does not follow the Mendelian laws of segregation and independent assortment of chromosomes.65 The main differences are related to occasional mutations that are incorporated into DNA molecules of mitochondria. Thus, the combination of karyotyping and mitochondrial DNA analyses has allowed to distinguish strains derived from starter yeasts and contaminants as well as to select robust yeast strains for ethanol production.54

The genome complexity of industrial yeast strains represents a new opportunity to select variants with infinite combinations of genes in different fermentation processes. Two very important contributions on genome sequencing of two industrial yeast strains traditionally used by Brazilian distilleries are PE266 and CAT1.67,68 The yeast strain PE2 showed a high genome plasticity mainly in telomeric and subtelomeric regions of chromosomes. This plasticity could explain why this strain adapts to different fermentation processes.66 CAT1, selected from a continuous fermentation process presented large genomic regions of heterozygosity loss while 58% of the 6652 predicted protein-coding genes were different alleles when compared to reference genome of strain S288c.68 Both strains, PE2 and CAT1, have an increased copy number of genes for biosynthesis of thiamin (B1) and pyridoxine (B6) in comparison to the baker's yeast and laboratory strains.67 An increased copy number can confer better fitness to these strains in vitamin-free fermentation medium with high concentration of sugars. CAT1 presented a short lag phase in comparison to S288c, a laboratory strain, when both yeasts were subjected to high sugar concentration in a medium without vitamins B1 and B6.67 These results are directly related to factors that affect the yeast dominance and persistence in industrial fermentation processes of ethanol production and explain why industrial strains are more robust and tolerant than the baker's yeast and laboratory strains. This means that monitoring yeast populations in industrial fermentations regarding their dominance and persistence characteristics is the basis to select tailored and high-performance strains according to the peculiar characteristics of each distillery. Currently, 18 distilleries use Tailored Yeast Strains® in their process. Some of them have two, three or even four customized strains (Fig. 5). This approach was designed as Process-Driven Selection®.54

Fig. 5.

Number of distilleries that use Tailored Yeast Strains® as starter for ethanol production in Brazil (gray bars) and total number of strains (dark bars) selected since 2008 by karyotyping and the mitochondrial DNA analysis.

(0.06MB).

However, yeast selection and biodiversity of wild strains still remain a scientific challenge to be explored.22,69 Moreover, it has been proposed that biodiversity of yeasts isolated from different environments can reveal new strains with desired characteristics to industrial applications such as ethanol production.70

Contaminating yeast and bacteria

Because of large volumes, industrial fermentations are subject to bacterial contamination, such as Saccharomyces and non-Saccharomyces wild species.26,71,72 Several wild yeasts compete with starter strains causing serious operational difficulties to industrial processes of ethanol production.26,54,73

A research carried out during five years demonstrated that all 73 wild yeasts isolated from two distilleries were classified as S. cerevisiae.74 However, contaminant strains can present different characteristics in relation to starter strains. After to evaluate 340 contaminating yeasts isolated from fermentation tanks of 50 Brazilian distilleries it was shown that 80% had undesirable characteristics for an optimum fermentation such as foaming, flocculation and high concentrations of residual sugars in the wine.26 Moreover, some strains presented a combination of these traits and most of them belonged to genus Saccharomyces. The main contaminant yeasts, classified as non-Saccharomyces, were Dekkera, Schizosaccharomyces and Candida. Dekkera is a common genus of contaminant yeasts causing significant reduction of fermentation yield in distilleries that does not warm the sugarcane juice and works at very low ethanol concentrations in the wine.75,76

Contaminating wild Saccharomyces show a high frequency of undesirable traits such as flocculation, pseudohyphae development and foam excess. Although flocculation and chain formation of cells are characteristics well documented, foam production by different strains is not well known. However, it has been associated to hydrophobicity of the yeast cell wall in ethanol and sake production.77,78

Although contaminants can enter the process through different ways, sugarcane is the major source and new strategies based on non-cultural techniques have been applied to access microbial communities that can affect industrial processes of bioethanol production.79,80 Besides wild yeasts, bacteria compete for sugars and nutrients with starter yeast strains causing many problems, such as inhibition of fermentation, production of organic acids, reduction of industrial yields and flocculation.81–83

A common procedure adopted by Brazilian distilleries that recycle yeast cells is the use of diluted sulfuric acid (pH 2.0–2.5 for 1–2h) to kill bacteria.26 Without acidic treatment, industrial fermentations are subject to sugar losses due to bacterial contamination. Microbiological analyses carried out in a distillery demonstrated that sulfuric acid treatment of the yeast cream reduced bacterial population by 44.55%.84 Control of bacterial contamination in industrial fermentations for ethanol production has been one of the most important factors that contributed to increase the fermentation yield in Brazilian distilleries.37

The development of a methodology to count living bacteria by microscopy in 15min, evolution of fast tests to identify the best antibiotics in only 6h and the use of antimicrobial products reduced significantly bacterial contamination in the industry.29,71 From 2007 onward, new antimicrobials were introduced to control bacterial contamination.85 This change was necessary due to pressure to reduce the use of antibiotics, mainly by distilleries that produce inactive dry yeast for animal feed. Chlorine dioxide and hop acids derivatives (alpha and beta fraction) are among the new antimicrobials used. These products are alternatives to control bacterial contaminants without use of antibiotics. Microscopic countings as low as 105 bacterial rods per mL of wine have been obtained in several distilleries. This means a thousand and ten thousand times less bacteria in fermentation than 40 years ago.37 Moreover, several compounds can affect not only bacteria but also the yeast cells and should be evaluated before to be used by the industry.86

Regarding bacterial biodiversity in industrial fermentations, 334 strains were isolated from Brazilian distilleries and classified using a matrix of similarity.84 The main bacterial communities were represented by Gram-positive (98.52%), rods (87.76%) and not-sporulated (73.95%). Among the main bacterial groups, the largest were Lactobacillus (59.75%) and Bacillus (26.58%) while other representatives were at minor proportion (14.67%). For the Lactobacillus group, the main species observed was L. fermentum (15.04%).84 Despite bacterial contamination is caused chiefly by Gram-positive bacteria, it was reported a case with drastic reduction of fermentation yield and ethanol production due to Gram-negative bacteria identified as Acetobacter pasteurianus by the sequencing of 16S rDNA gene.83

Lactic acid bacteria are the main contaminants of alcoholic fermentations. These bacteria are classified into two biochemical groups, according to metabolism of glucose: homo and heterofermentative. The first group is characterized by the production of lactate from glucose, while the second drives the metabolism for the production of lactate, acetate, ethanol and CO2.87

Since lactic acid bacteria are the main contaminants of alcoholic fermentations, lactic acid is an important indicator to monitor bacterial contamination. However, some bacterial species can produce different isomers of lactic acid or a mix of both. A research carried out with 27 strains of Lactobacillus isolated from industrial fermentations of ethanol production demonstrated that 70% of them produced dl-lactate while the other 30% produced only one isomer (d or l)-lactate.88 Furthermore, heterofermentative bacteria also produce mannitol which has been used as an indicator of sugarcane deterioration by microorganisms as well as of contamination of alcoholic fermentation processes by heterofermentative bacteria.89,90 Homo and heterofermentative lactic acid bacteria can affect the fermentation of different ways.91 Because of their importance, some distilleries have monitored mannitol, lactate, acetate and residual sugars in wine through high performance liquid chromatography improving their process control.29

Another very important issue is the effect of bacterial contamination on the yeast. A very common effect is flocculation of yeast cells. This kind of flocculation depends on physical contact between bacteria and yeast cells. Among lactic acid bacteria, L. fermentum, L. fructosus, L. buchneri and L. plantarum can cause yeast flocculation.92 When the ratio bacteria:yeast reaches 4.8:1, flocculation occurs.93 Furthermore, a fermentation assay carried out with three strains (baker's yeast, PE-2 and M-26), contaminated by L. fermentum, demonstrated that baker's yeast was the most sensible strain to bacterial contamination. Fermentations with baker's yeast presented the lowest rates of cell viability and the highest rates of bacterial multiplication. The results showed that dead cells release nutrients such as amino acids, vitamins and minerals that stimulate bacterial metabolism. In addition, PE-2 and M-26 showed the highest cell viability rates and lowest bacterial contamination by Lactobacillus.94 Later, the high tolerance of industrial and tailored yeast strains to bacterial contamination by L. fermentum while the baker's yeast was more sensible.54

One of the most important bacterial contamination sources of sugarcane juice is the soil. There is one billion of bacterial cells for each 1g of soil.95 Another important source of bacterial contamination is sugarcane infestation by borer. In addition, sugarcane stalks perforated by borer accumulate organic acids and phenolic compounds that can inhibit the fermentation.96

Bacterial contamination can affect the ethanol production differently as well as the indirect determination of fermentation yield based on by-products such as glycerol, yeast biomass, residual sugars and acidity. Some years ago, a distillery presented a fermentation process with contamination above 107bacterial rods/mL and a wine acidity that was lower than acidity of sugarcane must. It was not expected for an industrial fermentation. After isolation and biochemical tests were identified bacterial species belonging to malolactic group.97 These bacterial species have been isolated from grape wine fermentations but not from ethanol production. Malolactic bacteria are able to convert malic acid into lactic acid for ATP production.98 Malic acid is a dicarboxylic acid present in sugarcane juice, but it can also be produced by yeasts during the fermentation.49,99 Once lactic acid is monocarboxylic, these bacteria reduce the overall acidity of fermented medium. Because the acidity balance at the end of fermentation was negative, it overestimated the fermentation yield by indirect methodology based on by-products. In 2013, another case observed was the contamination by Acetobacter indonesiensis. It is a Gram-negative bacteria associated to soil and roots of sugarcane. This contamination was associated to a drastic glycerol reduction in wine, overestimating the real fermentation yield based on by-products.

All of these cases show the diversity and complexity of yeast and bacterial communities that contaminate industrial process of ethanol production and the different ways that these microorganisms can affect alcoholic fermentation and compromise the current ethanol production. The biodiversity of contaminants may affect not only ethanol production of first generation but also new technologies.

New technologies for production of first and second-generation ethanol

Despite reaching alcoholic fermentation yields as high as 92%, there are several scientific and technological challenges to be overcome for first and second-generation ethanol.22,100 In comparison to corn fermentations in the USA, Brazilian distilleries run their processes for eight months (or less) in a year because sugarcane cannot be harvested during the rainy season and cannot be stored as corn can. However, new Brazilian distilleries were built to ferment cornstarch, reducing the off-season without ethanol production.101 It has opened a new opportunity to increase the production of first-generation ethanol, reducing industrial fixed costs, employing specialized labor force and improving competitiveness of distilleries. Moreover, these processes have an enormous advantage because they do not generate vinasses. The main by-product is distillers dried grains with solubles (DDGS). It has a commercial value as high as that of ethanol, it is richer in protein than corn grain and can be used directly for animal feeding.102

More recently, a new Brazilian process of corn fermentation has been developed based on the reuse of surplus yeast cells from sugarcane fermentation that would be discarded by the plant. This Brazilian process integrates the fermentation of sugarcane juice (or molasses) and starchy feedstocks, providing a number of advantages over conventional distilleries.103 This process has a faster fermentation (34–36h) in comparison with the traditional process (45–60h) adopted by distilleries in the USA and less sugar is deviated for yeast multiplication and production of cellular biomass. In addition, this technology allows to use the same distillation system used for sugarcane and extend the period of ethanol production to 345 days a year, reducing initial investments and fixed costs. Each corn ton allows to produce 415L of ethanol and 250kg of DDGS.103 Moreover, integration of ethanol production from sugarcane and corn has the potential to offer significant economic advantages in comparison to stand-alone units, since important operations, yeast, feedstock, laboratories and technical personnel may be shared between both processes.

Flex fuel plants of corn and sugarcane have an energy balance and reduction of GHG emissions that do not compromise the fermentation performance. They are economically viable in regions with corn supply at low prices and high demand of DDGS for animal feed. It is an opportunity for Brazilian distilleries in corn producing regions that are far from ports.104

Besides ethanol production from starchy feedstocks, the last frontier of biofuels has been the use of lignocellulosic biomass-derived sugars.105 Cellulose and hemicellulose account for more than 60% of dry weight of sugarcane bagasse and can be converted into fermentable sugars either by acid or enzymatic hydrolysis.106 To achieve this goal, several research groups and companies worldwide have expended efforts and resources to produce not only ethanol, but also other biofuel molecules such as butanol and isobutanol.107–109

Research on synthetic biology and metabolic engineering has improved productivity and expectations of high yield for different biofuels. Moreover, selection of microorganisms to be used as biotechnological platform for metabolic engineering approaches has been considered a success key factor.110 These strains need to be improved regarding velocity of sugar assimilation, co-fermentation of pentoses and hexoses, tolerance to inhibitors produced from hydrolysates, tolerance to ethanol concentrations and recycling processes.22,110 Selection of industrial yeast strains has been a viable approach for a new generation of strains that can be used as platform to produce second-generation ethanol while other initiatives have focused on bacteria.111,112

Nowadays, in Brazil, there are two industrial plants in operation for the production of second-generation ethanol and one that use a semi-industrial process. These processes were originated from investment programs of Brazilian government in second-generation ethanol launched in 2011. Nevertheless, production costs of second-generation ethanol are still high, mainly concerning equipment handling at bagasse pre-treatment and the use of enzymes.113 Several initiatives have focused on selection of microbial strains and optimization of conditions for production of enzymes required for hydrolysis and fermentation in industrial processes.114–118

Lignocellulosic pre-treatments, such as steam explosion, alkaline hydrogen peroxide, diluted acid, ammonia, use of solvents, among others, are carried out prior to enzymatic hydrolysis in order to make cellulose more accessible to attack by enzymes.119 However, the pre-treatment of sugarcane bagasse still has technological challenges to be overcame in industrial scale. New technologies to disassembly the sugarcane fiber as well as the use of enzymes will improve the efficiency of bagasse hydrolysis as well as the production of ethanol 2G.120,121

Reduction of vinasse volumes

Vinasse is the resulting residue after distillation of wine. For each liter of ethanol produced in Brazil, on average, are generated other 12L of vinasse.122 Considering the ethanol production in 2015 of 30 billion liters, it means a production of 360 billion liters of vinasse per year. Vinasse is potassium-rich and it has been used in ferti-irrigation of sugarcane fields, reducing costs with chemical fertilizers.35 In addition, vinasse increases the soil pH some days after its application due to microbial activity stimulated by organic compounds and other nutrients.123

However, vinasse transport and application in the field are costly and require careful procedures to avoid contamination of table waters and soil saturation with cations.124 A reduction of vinasse volume increases the economic distance of transport and application in sugarcane crops.125–127 More concentrated, vinasse can be transported to distant areas from distillery reducing costs of application in the field.

The vinasse volume is directly related to ethanol concentration in wine (Table 3). Despite the overall yield of fermentation process increased from 75–80% to 90–92% in the last 30 years, ethanol concentrations in wine (8.0–8.5) has remained almost the same with exception of some distilleries that have run their processes with alcoholic contents above 10% (v/v).22,128,129

Table 3.

Increase of ethanol concentration in wine and reduction of vinasse volumes for a distillery with an ethanol production of 5200m3/week.

Ethanol concentration in wine (v/v)  L vinasse/L ethanol  Vinasse volume (m3/week) 
11.9  61,880 
10.6  55,120 
10  9.4  48,880 
11  8.5  44,200 
12  7.7  40,040 
13  7.0  36,400 
14  6.4  33,280 
15  5.9  30,680 
16  5.4  28,080 

Source: Fermentec.

A key factor to increase the ethanol concentrations in wine and reduce vinasse volumes has been the selection and use of robust yeast strains. Tests carried out at a pilot scale revealed that some industrial strains are more robust and viable for fermentations with high ethanol concentrations in wine than other traditional yeasts.130 These yeast strains open a new perspective to Brazilian distilleries increase the concentration of ethanol in wine for reduction of vinasse volumes.

Final remarks

Nowadays, several Brazilian distilleries are looking for technological innovations to improve their performance and ensure competitiveness of ethanol production. Evolution of first and second-generation ethanol depends of new feedstocks and a continuous improvement of the microbiological processes.22,131,132

Although production of first-generation ethanol has been considered a mature technology, there are huge opportunities of research, development and innovation for Brazilian distilleries.133 It includes reduction of sugars losses, new strategies to control bacterial contaminants, selection of new yeast strains, technologies for reduction of vinasse volumes and alternative uses by the industry, energy saving, better use of water, development of new fermentation processes for alternative feedstocks and biorrefineries for high-value added products.22

Moreover, it is necessary to develop strategies for transfer of new technologies to the industry.134 This movement of technologies is fundamental to increase efficiency and reduce costs. A study conducted in 2016 showed that for every R$ 1.00 invested in research and development, there is potential to return R$ 17.11 only in terms of reduction of production costs in Brazilian distilleries.135 Finally, investments in scientific and technological development, formation of researchers and specialized professionals will build solid bridges between science and industry for sustainable future of ethanol production in Brazil.136

Acknowledgment

To Marta Maria Moraes Leitão for drawings and illustrations.

References
[1]
H.V. Amorim, M.L. Lopes.
Ethanol production in a petroleum dependent world.
Sugar J, 67 (2005), pp. 11-14
[2]
N.I. Potter.
How Brazil achieved energy independence and the lessons the United States should learn from Brazil's experience.
Wash U Glob Stud Law Rev, 7 (2008), pp. 331-351
[3]
K.P.K. Olympio, C. Gonçalves, W.M.R. Günther, E.J.H. Bechara.
Neurotoxicity and aggressiveness triggered by low-level lead in children: a review.
Pan Am J Public Health, 26 (2009), pp. 266-275
[4]
J. Goldemberg.
The Brazilian biofuels industry.
Biotechnol Biofuels, 1 (2008), pp. 6
[5]
Anfavea.
Brazilian automotive industry yearbook.
Brazilian Automotive Indus-try Association, (2016),
[6]
I.C. Macedo, J.E.A. Seabra, J.E. Silva.
Greenhouse gases emissions in the production and use of ethanol from sugarcane in Brazil: the 2005/2006 averages and a prediction for 2020.
Biomass Bioenergy, 32 (2008), pp. 4
[7]
C. Pasha, L.V. Rao.
Thermotolerant yeasts for bioethanol production using lignocellulosic substrates.
Yeast biotechnology: diversity, pp. 551-558
[8]
J. Goldemberg.
Sugarcane ethanol: strategies to a successful program in Brazil.
Advanced biofuels and bioproducts, pp. 13-20
[9]
N.S. Hidzir, A.S. Som, Z. Abdullah.
Ethanol production via direct hydration of ethylene: a review.
International conference on global sustainability and chemical engineering (ICGSE),
[10]
D. Radecka, V. Mukherjee, R.Q. Mateo, M. Stojiljkovic, M.R. Foulquie-Moreno, J.M. Thevelein.
Looking beyond Saccharomyces: the potential of non-conventional yeast species for desirable traits in bioethanol fermentation.
FEMS Yeast Res, 15 (2015),
[11]
J.S. Rokem, C.L. Greenblatt.
Making biofuels competitive: the limitations of biology for fuel production.
JSM Microbiol, 3 (2015), pp. 1023
[12]
B. Hahn-Hägerdal, T. Lindén, T. Senac, K. Skoog.
Ethanolic fermentation of pentoses in lignocellulose hydrolysates.
Appl Biochem Biotechnol, 28–29 (1991), pp. 131-144
[13]
A.M. Jones, K.C. Thomas, W.M. Ingledew.
Ethanolic fermentation of blackstrap molasses and sugarcane juice using very high gravity technology.
J Agric Food Chem, 42 (1994), pp. 1242-1246
[14]
G.P. Silva, E.F. Araújo, D.O. Silva, W.V. Guimarães.
Ethanolic fermentation of sucrose, sugarcane juice and molasses by Escherichia coli strain ko11 and Klebsiella oxytoca strain P2.
Braz J Microbiol, 36 (2005), pp. 395-404
[15]
L.L. Madaleno, G.D. Minari, F.R. Annunzio, et al.
Use of antimicrobials for contamination control during ethanolic fermentation.
Científica, 44 (2016), pp. 226-234
[16]
D.A. Monceaux.
Alternative feedstocks for fuel ethanol production.
The alcohol textbook. A reference for the beverage, fuel and industrial alcohol industries, 5th ed., pp. 47-57
[17]
F. Wang, Y. Jiang, W. Guo, et al.
An environmentally friendly and productive process for bioethanol production from potato waste.
Biotechnol Biofuels, 9 (2016), pp. 50
[18]
H.V. Amorim, M.L. Lopes.
Tecnologia sobre processamento de leveduras vivas, inativas e seus derivados: conceitos básicos.
Congresso Internacional Sobre uso da Levedura na Alimentação Animal, pp. 5-20
[in Portuguese]
[19]
R. Sage, R.K. Monson.
C4 plant biology.
Elsevier, (1999),
[20]
M. Wang, J. Han, J.B. Dunn, H. Cai, A. Elgowainy.
Well-to wheels energy use and greenhouse gas emissions of ethanol from corn, sugarcane and cellulosic biomass for US use.
Environ Res Lett, 7 (2012), pp. 045905
[21]
A.E. Wheals, L.C. Basso, D.M.G. Alves, H.V. Amorim.
Fuel ethanol after 25 years.
Trends Biotechnol, 17 (1999), pp. 482-487
[22]
H.V. Amorim, M.L. Lopes, J.V.C. Oliveira, M. Buckeridge, G.H. Goldman.
Scientific challenges of bioethanol production in Brazil.
Appl Microbiol Biotechnol, 91 (2011), pp. 1267-1275
[23]
S.C. Bhatia.
Ethanol.
Advanced renewable energy systems. Part I, pp. 523-572
[24]
E.D. Larson.
A review of life-cycle analysis studies on liquid biofuels systems for the transport sector.
Energy Sustain Dev, 10 (2006), pp. 109-126
[25]
S.T. Coelho, J. Goldemberg, O. Lucon, P. Guardabassi.
Brazilian sugarcane ethanol: lessons learned.
Energy Sustain Dev, 10 (2006), pp. 26-39
[26]
L.C. Basso, H.V. Amorim, A.J. Oliveira, M.L. Lopes.
Yeast selection for fuel ethanol production in Brazil.
FEMS Yeast Res, 8 (2008), pp. 1155-1163
[27]
H.V. Amorim, M. Gryschek, M.L. Lopes.
The success and sustainability of the Brazilian sugarcane–fuel ethanol industry.
Sustainability of the sugar and sugar ethanol industries, pp. 73-82
[28]
B.E. Della-Bianca, T.O. Basso, B.U. Stambuk, L.C. Basso, A.K. Gombert.
What do we know about the yeast strains from the Brazilian fuel ethanol industry?.
Appl Microbiol Biotechnol, 97 (2013), pp. 979-991
[29]
E.P. Borges, M.L. Lopes, C. Bernardino, et al.
The benefits of applied research: 37 years of discoveries, adaptations and solutions.
Sugar Ind, 140 (2015), pp. 209-216
[30]
E. Bertrand, L.P.S. Vandenberghe, C.R. Soccol, J.C. Sigoillot, C. Faulds.
First generation bioethanol.
Green fuels technology, pp. 175-212
[31]
Boinot F. Process for carrying out industrial alcoholic fermentations. United States Patent Office PI 2,230,318; 1941. Available from: https://docs.google.com/viewer?url=patentimages.storage.googleapis.com/pdfs/US2230318.pdf.
[32]
W.M. Ingledew, G.D. Austin, D.R. Kelsall, C. Kluhspies.
The alcohol industry: how has it changed and matured?.
The alcohol textbook. A reference for the beverage, fuel and industrial alcohol industries, pp. 1-6
[33]
U.A. Lima, L.C. Basso, H.V. Amorim.
Produção de etanol.
Biotecnologia Industrial: processos fermentativos e enzimáticos, pp. 1-43
[chapter 3]
[34]
C. Martinez-Amezcua, C.M. Parsons, V. Singh, R. Srinivasan, G.S. Murthy.
Nutritional characteristics of corn distillers dried grains with solubles as affected by the amounts of grains versus solubles and different processing techniques.
Poult Sci, 86 (2007), pp. 2624-2630
[35]
R.M. Prado, G. Caione, C.N.S. Campos.
Filter cake and vinasse as fertilizers contributing to conservation agriculture.
Appl Environ Soil Sci, (2013),
[36]
H.V. Amorim, L.C. Basso, M.L. Lopes.
Sugar cane juice and molasses, beet molasses and sweet sorghum: composition and usage.
The alcohol textbook: a reference for the beverage, fuel, and industrial alcohol industries, pp. 39-46
[37]
A. Godoy, H.V. Amorim, M.L. Lopes, A.J. Oliveira.
Continuous and batch fermentation processes: advantages and disadvantages of these processes in the Brazilian ethanol production.
Int Sugar J, 110 (2008), pp. 175-181
[38]
M.G.S. Andrietta, S.R. Andrietta, C. Steckelberg, E.N.A. Stupiello.
Bioethanol – 30 years of Proálcool.
Int Sugar J, 109 (2007), pp. 195-200
[39]
J.A. Barnett, F.W. Lichtenthaler.
A history of research on yeasts 3: Emil Fischer, Eduard Buchner and their contemporaries, 1880–1900.
[40]
C. Laluce, M.C. Palmieri, C.L. Cruz.
Growth and fermentation characteristics of new selected strains of Saccharomyces at high temperatures and high cell densities.
Biotechnol Bioeng, 37 (1990), pp. 528-536
[41]
C. Dorta, P. Oliva-Neto, M.S. de-Abreu-Neto, N. Nicolau-Junior, A.I. Nagashima.
Synergism among lactic acid, sulfite, pH and ethanol in alcoholic fermentation of Saccharomyces cerevisiae (PE-2 and M-26).
World J Microbiol Biotechnol, 22 (2006), pp. 177-182
[42]
N.A. Brown, P.A. Castro, B.C.P. Figueiredo, et al.
Transcriptional profiling of Brazilian Saccharomyces cerevisiae strains selected for semi-continuous fermentation of sugarcane must.
FEMS Yeast Res, 13 (2013), pp. 277-290
[43]
V.R. Reis, A.P.G. Bassi, J.C.G. da Silva, S.R. Ceccato-Antonini.
Characteristics of Saccharomyces cerevisiae yeasts exhibiting rough colonies and pseudohyphal morphology with respect to alcoholic fermentation.
Braz J Microbiol, 44 (2013), pp. 1121-1131
[44]
L.V. Ferreira, H.V. Amorim, L.C. Basso.
Fermentação de trealose e glicogênio endógenos em Saccharomyces cerevisiae.
Ciênc Tecnol Alim, 19 (1999), pp. 29-32
[in Portuguese]
[45]
S.C.L. Paulillo, F. Yokoya, L.C. Basso.
Mobilization of endogenous glycogen and trehalose of industrial yeasts.
Braz J Microbiol, 34 (2003), pp. 249-254
[46]
A.C. Panek, J.J.A. Mansure, M.F. Paschoalin, D. Panek.
Regulation of trehalose metabolism in Saccharomyces cerevisiae mutants during temperature shifts.
Biochemie, 72 (1990), pp. 77-79
[47]
T. D’Amore, R. Crumplen, G.G. Stewart.
The involvement of trehalose in yeast stress tolerance.
J Ind Microbiol, 7 (1991), pp. 191-195
[48]
L.C. Basso, D.M.G. Alves, H.V. Amorim.
The antibacterial action of succinic acid produced by yeast during fermentation.
Rev Microbiol, 28 (1997), pp. 77-82
[49]
D.M.G. Alves.
Respostas fisiológicas de duas linhagens de Saccharomyces cerevisiae frente ao potássio durante a fermentação alcoólica.
Universidade Estadual Paulista “Julio de Mesquita”, (2000),
118 pp. [in Portuguese]
[50]
G.M. Walker.
Yeast physiology and biotechnology.
John Wiley & Sons, (1998),
[51]
L.S. Watrud, K. Martin, K.K. Donegan, J.K. Stone, C.G. Coleman.
Comparison of taxonomic, colony morphotype and PCR-RFLP methods to characterize microfungal diversity.
Mycologia, 98 (2006), pp. 384-392
[52]
C.J. Gimeno, G.R. Fink.
Induction of pseudohyphal growth by overexpression of PHD1, a Saccharomyces cerevisiae gene related to transcriptional regulators of fungal development.
Mol Cell Biol, 14 (1994), pp. 2100-2112
[53]
S.R. Ceccato-Antonini, P.E. Sudbery.
Filamentous growth in Saccharomyces cerevisiae.
Braz J Microbiol, 35 (2004), pp. 173-181
[54]
M.L. Lopes, S.C.L. Paulillo, R.A. Cherubin, A. Godoy, H.B. Amorim Neto, H.V. Amorim.
Tailored-yeast strains for ethanol production: process-driven selection.
[55]
F. Vezinhet, B. Blondin, J.N. Hallet.
Chromosomal DNA patterns and mitochondrial DNA polymorphism as tools for identification of enological strains of Saccharomyces cerevisiae.
Appl Microbiol Biotechnol, 32 (1990), pp. 568-571
[56]
F. Vezinhet, J.N. Hallet, M. Valade, A. Poulard.
Ecological survey of wine yeast strains by molecular methods of identification.
Am J Enol Viticult, 43 (1992), pp. 83-86
[57]
L.C. Basso, A.J. Oliveira, V.F.M. Orelli, A.A. Campos, C.R. Gallo, H.V. Amorim.
Dominância das leveduras contaminantes sobre as cepas industriais avaliada pela técnica da cariotipagem.
V Congr Nac STAB, 5 (1993), pp. 246-250
[in Portuguese]
[58]
M.E. Zolan.
Chromosome-length polymorphism in fungi.
Microbiol Rev, 59 (1995), pp. 686-698
[59]
A.C. Codón, T. Benítez, M. Korhola.
Chromosomal reorganization during meiosis of Saccharomyces cerevisiae Baker's yeasts.
Curr Genet, 32 (1997), pp. 247-259
[60]
A.C. Codón, T. Benítez, M. Korhola.
Chromosomal polymorphism and adaptation to specific industrial environments of Saccharomyces strains.
Appl Microbiol Biotechnol, 49 (1998), pp. 154-163
[61]
H.V. Amorim, M.L. Lopes.
Ciência e tecnologia na seleção de leveduras para produção de etanol.
Microrganismos em Agroenergia: da Prospecção aos Bioprocessos, pp. 42-59
[62]
T. Benítez, P. Martínez, A.C. Codón.
Genetic constitution of industrial yeast.
Microbiologia, 12 (1996), pp. 371-384
[63]
M.L. Lopes.
Estudo do polimorfismo cromossômico em Saccharomyces cerevisiae (linhagem PE-2) utilizada no processo industrial de produção de etanol.
Universidade Estadual Paulista “Julio de Mesquita Filho” UNESP, (2000), pp. 118
[in Portuguese]
[64]
V. López, A. Querol, D. Ramo, M.T. Fernandez-Espinar.
A simplified procedure to analyse mitochondrial DNA from industrial yeasts.
Int J Food Microbiol, 68 (2001), pp. 75-81
[65]
W.C. Birky Jr..
The inheritance of genes in mitochondria and chloroplasts: laws, mechanisms, and models.
[66]
J.L. Argueso, M.F. Carazzolle, P.A. Mieczkowski, F.M. Duarte, O.V.C. Netto, S.K. Missawa.
Genome structure of a Saccharomyces cerevisiae strain widely used in bioethanol production.
Genome Res, 19 (2009), pp. 2258-2270
[67]
B.U. Stambuk, B. Dunn, S.L. Alves Junior, E.H. Duval, G. Sherlock.
Industrial fuel ethanol yeasts contain adaptive copy number changes in genes involved in vitamin B1 and B6 biosynthesis.
Genome Res, 19 (2009), pp. 2271-2278
[68]
F. Babrzadeh, R. Jalili, C. Wang, et al.
Whole-genome sequencing of the efficient industrial fuel-ethanol fermentative Saccharomyces cerevisiae strain CAT-1.
Mol Genet Genomics, 6 (2012), pp. 485-494
[69]
M.A. Lachance.
Yeast biodiversity: how many and how much?.
Biodiversity and ecophysiology of yeasts,
[chapter 1]
[70]
C.L. Ramos, W.F. Duarte, A.L. Freire, D.R. Dias, E.C.A. Eleutherio, R.F. Schwan.
Evaluation of stress tolerance and fermentative behavior of indigenous Saccharomyces cerevisiae.
Braz J Microbiol, 44 (2013), pp. 935-944
[71]
H.V. Amorim, A.J. Oliveira.
Infecção na fermentação: como evitá-la.
Açúcar Álcool, 5 (1982), pp. 12-18
[in Portuguese]
[72]
B.T. Lucena, B.M. dos Santos, J.L. Moreira, et al.
Diversity of lactic acid bacteria of the bioethanol process.
BMC Microbiol, 10 (2010), pp. 298
[73]
M.C.F.L. Oliveira, F.C. Pagnocca.
Aplicabilidade de meios seletivos empregados na indústria cervejeira para detecção de leveduras selvagens em unidades sucroalcooleiras.
Simpósio Nacional de Fermentação, VIII, Sinaferm, (1988), pp. 78-81
[in Portuguese]
[74]
C.T. Stroppa.
Dinâmica Populacional de leveduras caracterizadas por eletro–cariótipo e desempenho fermentativo em processos de fermentação alcoólica.
Universidade Estadual de Campinas, (2002), pp. 133
[in Portuguese]
[75]
A.T.D. Liberal, A.C.M. Basilio, A.D. Resende, et al.
Identification of Dekkera bruxellensis as a major contaminant yeast in continuous fuel ethanol fermentation.
J Appl Microbiol, 102 (2007), pp. 538-547
[76]
A.C.M. Basilio, P.R.L. de Araujo, J.O.F. de Morais, E.A. de Silva, M.A. de Morais, D.A. Simoes.
Detection and identification of wild yeast contaminants of the industrial fuel ethanol fermentation process.
Curr Microbiol, 56 (2008), pp. 322-326
[77]
C.M. Figueiredo.
Análise molecular da floculação e da formação de espuma por leveduras utilizadas na produção industrial de álcool combustível no Brasil.
Universidade Federal de Santa Catarina, (2008), pp. 82
[78]
H. Shimoi, K. Sakamoto, M. Okuda, R. Atthi, K. Iwashita, K. Ito.
The AWA1 gene is required for the foam-forming phenotype and cell surface hydrophobicity of sake yeast.
Appl Environ Microbiol, 68 (2002), pp. 2018-2025
[79]
O.Y.A. Costa, B.M. Souto, D.D. Tupinambá, et al.
Microbial diversity in sugarcane ethanol production in a Brazilian distillery a culture-independent method.
J Ind Microbiol Biotechnol, 42 (2015), pp. 73-84
[80]
R.S.C. Souza, V.K. Okura, J.S.L. Armanhi, et al.
Unlocking the bacterial and fungal communities assemblages of sugarcane microbiome.
[81]
P. Oliva-Neto, F. Yokoya.
Effects of nutritional factors on growth of Lactobacillus fermentum mixed with Saccharomyces cerevisiae in alcoholic fermentation.
Rev Microbiol, 28 (1997), pp. 25-31
[82]
K.M. Ludwig, P. Oliva-Neto, D.F. Angelis.
Quantificação da floculação de Saccharomyces cerevisiae por bactérias contaminantes da fermentação alcoólica.
Ciênc Tecnol Aliment, 21 (2001), pp. 63-68
[in Portuguese]
[83]
M.L. Lopes, H.V. Amorim, A. Godoy, A.J. Oliveira, R.A. Cherubin, L.C. Basso.
Interaction between yeast and acetic acid bacteria in industrial fermentation for ethanol production: a case study.
11th International Congress on Yeasts; vol. 1, pp. 183
[84]
C.R. Gallo.
Determinação da microbiota bacteriana de mosto e de dornas de fermentação alcoólica.
Universidade Estadual de Campinas, (1989), pp. 387
[85]
S.P. Meneghin, F.C. Reis, P.G. Almeida, S.R. Ceccato-Antonini.
Chlorine dioxide against bacteria and yeasts from the alcoholic fermentation.
Braz J Microbiol, 39 (2008), pp. 337-343
[86]
P. Oliva-Neto, F. Yokoya.
Susceptibility of Saccharomyces cerevisiae and lactic acid bacteria from the alcohol industry to several antimicrobial compounds.
Braz J Microbiol, 32 (2001), pp. 10-14
[87]
M.T. Madigan, J.H. Martinko, K.S. Bender, D.H. Buckley, D.A. Stahl.
Brock biology of microorganisms.
Pearson, (2015), pp. 1041
[88]
V.M. Costa, T.O. Basso, L.H.P. Angeloni, M. Oetterer, L.C. Basso.
Production of acetic acid, ethanol and optical isomers of lactic acid by Lactobacillus strains isolated from industrial ethanol fermentations.
Cienc Agrotecnol, 32 (2008), pp. 503-509
[89]
G. Eggleston, B. Legendre, T. Tew.
Indicators of freeze-damage sugarcane varieties which can predict processing problems.
Food Chem, 87 (2004), pp. 119-133
[90]
G. Eggleston, L.C. Basso, H.V. de Amorim, S.C.D. Paulillo, T.O. Basso.
Mannitol as a sensitive indicator of sugarcane deterioration and bacterial contamination in fuel alcohol production.
Zuckerindustrie, 132 (2007), pp. 33-39
[91]
T.O. Basso, F.S. Gomes, M.L. Lopes, H.V. Amorim, G. Eggleston, L.C. Basso.
Homo- and heterofermentative lactobacilli differently affect sugarcane-based fuel ethanol fermentation.
Antonie van Leeuwenhoek, 105 (2014), pp. 169-177
[92]
V.E. Alcarde.
Avaliação de parâmetros que afetam a floculação de leveduras e bactérias isoladas de processos industriais de fermentação alcoólica.
Universidade Estadual de Campinas, (2001), pp. 91
[in Portuguese]
[93]
F. Yokoya, P. Oliva-Neto.
Características da floculação de leveduras por Lactobacillus fermentum.
Rev Microbiol, 22 (1991), pp. 12-16
[in Portuguese]
[94]
R.A. Cherubin.
Efeitos da viabilidade da levedura e da contaminação bacteriana na fermentação alcoólica.
Escola Superior de Agricultura “Luiz de Queiroz”, (2003), pp. 137
[in Portuguese]
[95]
J. Gans, M. Wolinsky, J. Dunbar.
Computational improvements reveal great bacterial diversity and high metal toxicity in soil.
Science, 309 (2005), pp. 1387-1390
[96]
J.P. Stupiello.
Complexo broca/podridões: avaliação-interpretação da intensidade de dano e efeitos no processamento de álcool.
Rev STAB – Açúcar Álcool Subprod, 17 (1999),
[in Portuguese]
[97]
S.C.L. Paulillo, E.P. Borges, V.M. Costa, et al.
Bactérias malolácticas na fermentação alcoólica.
Anais dos Resumos do 25° Congresso Brasileiro de Microbiologia,
ID 63.1 [in Portuguese]
[98]
S. Kolb, H. Otte, B. Nagel, B. Schink.
Energy conservation in malolactic fermentation by Lactobacillus plantarum and Lactobacillus sake.
Arch Microbiol, 157 (1992), pp. 457-463
[99]
L.E. Gutierrez, S.E. Ferrari, A.A. Orelli Jr..
Ácidos orgânicos de caldo de três variedades de cana-de-açúcar (Saccharum spp) em quatro estádios de maturidade.
An Esc Super Agric Luiz Queiroz, 46 (1989), pp. 41-52
[in Portuguese]
[100]
A. Abreu-Cavalheiro, G. Monteiro.
Solving ethanol production problems with genetically modified yeast strains.
Braz J Microbiol, 44 (2013), pp. 665-671
[101]
J.L. Borsatto Jr., R.F. Souza, Weiss Lass, D.G. Dal Vesco.
Ensaio teórico sobre viabilidade da produção de etanol a partir do milho: uma possibilidade para as usinas sucroalcooleiras e produtores de grãos do estado de Mato Grosso.
Annals of IV international symposium on project management, innovation and sustainability,
[102]
M.E. Corrigan, R.A. Mass.
Utilisation of distiller's dried grains with solubles (DGGS) by cattle.
The alcohol textbook. A reference for the beverage, fuel and industrial alcohol industries, 5th ed., pp. 323-335
[103]
Fermentec Tecnologias em Açúcar e Álcool Ltda.
Processo de reaproveitamento de biomassa de levedo na integração de fermentações alcoólicas de cana e substratos amiláceo.
(2015),
BR 10 2015 021056 6 [in Portuguese]
[104]
A.Y. Milanez, D. Nyko, M.S. Valente, et al.
A produção de etanol pela integração do milho-safrinha às usinas de cana-de-açúcar: avaliação ambiental, econômica e sugestões de política.
[105]
A. Kang, T.S. Lee.
Converting sugars to biofuels: ethanol and beyond.
Bioengineering, 2 (2015), pp. 184-203
[106]
N. Pereira Jr., M.A.P.G. Couto, L.M.M. Santa Anna.
Biomass of lignocellulosic composition for fuel ethanol production within the context of biorefinery,
[107]
A.P. Mariano, M.O.S. Dias, T.L. Junqueira, et al.
Utilization of pentoses from sugarcane biomass: techno-economics of biogas vs. butanol production.
Bioresour Technol, 142 (2013), pp. 390-399
[108]
J. Natalense, Z. Désirée.
Technology roadmapping for renewable fuels: case of biobutanol in Brazil.
J Technol Manag Innov, 8 (2013), pp. 143-152
[109]
Z.W. Pang, W. Lu, H. Zhang, et al.
Butanol production employing fed-batch fermentation by Clostridium acetobutylicum GX01 using alkali-pretreated sugarcane bagasse hydrolysed by enzymes from Thermoascus aurantiacus QS 7-2-4.
Bioresour Technol, 212 (2016), pp. 82-91
[110]
E.A. Silva, W.F. Melo, D.F. Antunes, et al.
Isolation by genetic and physiological characteristics of a fuel-ethanol fermentative Saccharomyces cerevisiae strain with potential for genetic manipulation.
J Ind Microbiol Biotechnol, 32 (2005), pp. 481-486
[111]
T.P. Basso.
Improvement of Saccharomyces cerevisiae by hybridization for increased tolerance towards inhibitors from second-generation ethanol substrate.
Escola Superior de Agricultura “Luiz de Queiroz”, (2015),
104 pp.
[112]
S.M. Scully, J. Orlygsson.
Recent advances in second generation ethanol production by thermophilic bacteria.
Energies, 8 (2015), pp. 1-30
[113]
J.L. Ramos, M. Valdivia, F. Garcia-Lorente, A. Segura.
Benefits and perspectives on the use of biofuels.
Microb Biotechnol, 9 (2016), pp. 436-440
[114]
R. Silva, E.S. Lago, C.W. Merheb, M.M. Macchione, Y.K. Park, E. Gomes.
Production of xylanase and cmcase on solid state fermentation in different residues by Thermoascus aurantiacus Miehe.
Braz J Microbiol, 36 (2005), pp. 235-241
[115]
M.M.S. Moretti, D.A. Bocchini-Martins, R. Silva, A. Rodrigues, L.D. Sette, E. Gomes.
Selection of thermophilic and thermotolerant fungi for the production of cellulases and xylanases under solid-state fermentation.
Braz J Microbiol, 43 (2012), pp. 1062-1071
[116]
M.P. Vásquez, J.N. da Silva, M.B. de Souza Jr., N. Pereira Jr..
Enzymatic hydrolysis optimization to ethanol production by simultaneous saccharification and fermentation.
Appl Biochem Biotechnol, 137–140 (2007), pp. 141-153
[117]
V. Ferreira, M.O. Faber, S.S. Mesquita, N. Pereira Jr..
Simultaneous saccharification and fermentation process of different cellulosic substrates using a recombinant Saccharomyces cerevisiae harbouring the β-glucosidase gene.
Electron J Biotechnol, 13 (2010),
[118]
V.M. Benassi, R.C. de Lucas, J.A. Jorge, M.L.T.M. Polizeli.
Screening of thermotolerant and thermophilic fungi aiming β-xylosidase and arabinanase production.
Braz J Microbiol, 45 (2014), pp. 1459-1467
[119]
Y. You, S. Yang, L. Bu, J. Jiang, D. Sun.
Comparative study of simultaneous saccharification and fermentation byproducts from sugarcane bagasse using steam explosion, alkaline hydrogen peroxide and organosolv pretreatments.
RSC Adv, 6 (2016), pp. 13723-13729
[120]
M.C.A. Wanderley, C. Martín, G.J.M. Rochad, E.R. Gouveia.
Increase in ethanol production from sugarcane bagasse based on combined pretreatments and fed-batch enzymatic hydrolysis.
Bioresour Technol, 128 (2013), pp. 448-453
[121]
M.S. Buckeridge, W.D. Santos, A.P. Souza.
Routes for cellulosic ethanol in Brazil.
Sugarcane bioethanol — R&D for productivity and sustainability, pp. 365-380
[122]
F.P. Romano.
Benefícios econômicos do aumento do teor alcoólico no vinho de fermentação.
Escola Superior de Agricultura “Luiz de Queiroz”, (2015),
48 pp. [in Portuguese]
[123]
J.R. Almeida, G. Ranzani, O. Valsechi.
Contribuition à 1’étude de la vinasse dans l’Agriculture.
VIII Congr Int Ind Agric, 1 (1950), pp. 13-21
[124]
C.A. Christofoletti, J.P. Escher, J.E. Correia, J.F.U. Marinho, C.S. Fontanetti.
Sugarcane vinasse: environmental implications of its use.
Waste Manage, 33 (2013), pp. 2752-2761
[125]
H.A. Germek, G.E. Feigl.
Processo de redução de vinhaça.
Rev STAB Açúcar Álcool Subprod, 2 (1987), pp. 42-50
[in Portuguese]
[126]
G.A. Silva.
Possibilidades tecnológicas para a destinação da vinhaça de cana-de-açúcar quanto ao aspecto de desenvolvimento ambiental e econômico.
Escola de Engenharia de São Carlos, (2012),
208 pp. [in Portuguese]
[127]
H.V. Amorim, M.S. Lorenzi, F.C. Boscariol, G.M. Ferreira.
Alto teor alcoólico e concentração de vinhaça: Processos se complementam.
Idea News, 134 (2012), pp. 26-32
[in Portuguese]
[128]
J. Doerfler, H.V. Amorim.
Applied bioethanol technology in Brazil.
Sugar Ind, 132 (2007), pp. 694-697
[129]
F.A.F.C. Vicente.
Seleção, avaliação e utilização de uma levedura personalizada para a produção de etanol.
Universidade Estadual Paulista “Julio de Mesquita Filho”, (2015),
113 pp. [in Portuguese]
[130]
H.V. Amorim, M.S. Lorenzi, F.C. Boscariol, G.M. Ferreira.
Information on Fermentec/Dedini process.
Sugar J, 74 (2011), pp. 34-35
[131]
E. Winter, A.A. Lima, C.D.S. Mendes.
Mapeamento tecnológico da cadeia produtiva do etanol proveniente da cana-de-açúcar sob enfoque dos pedidos de patente: cenário brasileiro.
Bioetanol de Cana-de-Açúcar: P&D para produtividade e sustentabilidade,
[chapter 10, in Portuguese]
[132]
S. Salles-Filho, A.B. Sicsú, A. Bina, et al.
Futuros do Bioetanol: Brasil na liderança.
1st ed., Elsevier, (2015),
[in Portuguese]
[133]
A.T. Furtado, M.I.G. Scandiffio, L.A.B. Cortez.
The Brazilian sugarcane innovation system.
Energy Policy, 39 (2011), pp. 56-66
[134]
E.M. Rogers, S. Takegami, J. Yin.
Lessons learned about technology transfer.
Technovation, 21 (2001), pp. 253-261
[135]
F.H.C.A Giometti.
Impacto das tecnologias industriais nos custos de produção.
37ª Reunião Anual Fermentec,
[in Portuguese]
[136]
H.V. Amorim, R.M. Leão.
Fermentação Alcoólica: Ciência e Tecnologia.
1st ed., Fermentec, (2005),
[chapter 11, in Portuguese]
Copyright © 2016. Sociedade Brasileira de Microbiologia
Article options
Tools