Next Article in Journal
Wing Deformation of an Airborne Wind Energy System in Crosswind Flight Using High-Fidelity Fluid–Structure Interaction
Previous Article in Journal
A Facile Approach for the Encapsulation of Perovskite Solar Cells
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Promising Bioalcohols for Low-Emission Vehicles

by
Manju Dhakad Tanwar
1,*,
Felipe Andrade Torres
2,
Ali Mubarak Alqahtani
3,
Pankaj Kumar Tanwar
1,
Yashas Bhand
1 and
Omid Doustdar
4
1
Organic Recycling Systems Limited, Navi Mumbai 400703, India
2
Department of Mechanical Systems, Center of Exact and Technological Sciences, Federal University of Recôncavo of Bahia, Cruz das Almas 44380-000, Brazil
3
Mechanical Engineering Department, Jubail Industrial College, Jubail 31961, Saudi Arabia
4
Department of Mechanical Engineering, School of Engineering, University of Birmingham, Birmingham B15 2TT, UK
*
Author to whom correspondence should be addressed.
Energies 2023, 16(2), 597; https://doi.org/10.3390/en16020597
Submission received: 25 November 2022 / Revised: 23 December 2022 / Accepted: 30 December 2022 / Published: 4 January 2023

Abstract

:
In recent decades, many kinds of research have been conducted on alternative fuels for compression ignition (CI) engines. Low/zero-carbon fuels, such as bioalcohols and hydrogen, are the most promising alternative fuels and are extensively studied because of their availability, ease of manufacturing, and environmental benefits. Using these promising fuels in CI engines is environmentally and economically beneficial. The most common alcohols are methanol, ethanol, isopropanol, propanol, butanol, n-butanol, tert-butanol, iso-butanol, and pentanol. The primary objective of this review paper is to examine the impact of bioalcohols and their blends with conventional diesel fuel in CI engines since these fuels possess characteristic properties that impact overall engine performance and exhaust emissions. This research also indicated that alcohols and blended fuels could be used as fuels in compression ignition engines. Chemical and physical properties of alcohols were examined, such as lubricity, viscosity, calorific value, and cetane number, and their combustion characteristics in compression ignition engines provide a comprehensive review of their potential biofuels as alternative fuels.

1. Introduction

According to World Energy Balances, fossil fuels provided around 81% of the total primary energy supply in 2019, with renewables providing nearly 14% [1]. Although the concern about oil reserve depletion has driven research into petroleum-based fuel substitutes, hydro, wind, and solar energies are aimed at large-scale power production. The transportation industry has the biggest need for petroleum-based derivatives, accounting for approximately 60% of total demand, with road transportation accounting for the majority [2,3]. Furthermore, scientists are investigating alternative fuel resources in response to rising fossil fuel prices, detrimental greenhouse gas emissions, and energy production diversity [4,5].
Conventional transportation infrastructure continues to be a major source of pollution, influencing overall greenhouse gas emissions and their impact on human health [6]. Although fully electric vehicles have been found to reduce pollutants, their extremely high cost and unreliability of energy output imply that they are not a short-term solution. This has resulted in a search for alternative fuels and lubricants, including low-carbon and zero-carbon fuels [7]. Solar energy, as an alternative source of energy, can still be used in a variety of ways. Solar rays are used by plants for photosynthesis, which converts them into stored energy in chemical bonds within plant tissue, which can be converted to fuel through different thermos-chemical processes [8]. Despite that electrification has been increasing for passenger vehicles fleet, both EVs (electric vehicles) and HEVs (hybrid electric vehicles), it is essential to focus on the development of advanced combustion and aftertreatment techniques as a short- and medium-term plan [9]. In this scenario, biofuel (low/zero-carbon fuels) produced from renewable sources is considered the most viable alternative, in a short-term scenario, for replacing conventional diesel fuel in compression ignition engines and is still and will continue to be necessary for cleaner mobility.
The industrial revolution resulted in the production of steam and waterpower, and since then, energy sources have changed to meet every increasing human demand. In 1900, Rudolph Diesel, the first diesel cycle engine designer, used vegetable oil (groundnut oil) as a fuel for his prototype [6]. His objective was to enhance efficiency and reduce the cost and size of engines in comparison to large steam engines so that smaller businesses could afford the machinery [8]. It was also suggested that using vegetable oil will significantly aid countries’ agricultural efforts [10]. Many countries, such as Brazil and the United States, have diversified their biofuel production. For example, after the 1970s oil crisis, Brazil deployed ethanol fuel from sugar cane; however, the benefits began to fade due to an increase in ethanol fuel costs in 1989 [6]. Over the last few decades, there has been an increase in global warming and climate change caused by fossil fuels, specifically petroleum-based liquid fuels, natural gas, and coal [11]. This made biofuels derived from vegetable oils a promising alternative fuel [4].
Because of the oil resource, human health, and environmental impacts of fossil fuels, many researchers investigate the usage of vegetables to produce alternative fuels, usually edible vegetables. Most investigations have been done based on the use of fuels in diesel engines without any modification. Based on the literature, many researchers have attempted to review important investigations on biodiesel production processes, physical and chemical properties, and their performance and emissions characteristics as compression ignition engine fuels. Among the alternatives, alcohol fuels as low-carbon fuels possess certain advantageous properties that give them the potential to lower pollutants and CO2 emissions from IC engines. However, alcohols also have some limitations in their application, such as miscibility, lower cetane number, and lower heating value than conventional diesel fuel. In this review paper, the potential future directions for bioalcohol fuels for low-emission vehicles have been comprehensively investigated in a way toward net-zero emission transportation.

2. Why Bioalcohols?

Alcohol fuels are once again becoming a popular keyword for clean fuel utilization as a means of mitigating global climate change and meeting the needs of less-used local energy sources. As raw feedstock, lignocellulosic biomass and algae could be derived from grain-based raw materials. It is critical to use locally available, underutilized feedstocks for local energy security and distributed energy infrastructure. Alcohols as low-carbon fuels such as methanol, ethanol, propanol, and butanol are the most common alcohols for automotive applications [12,13,14]. Investigations on the effects of propanol isomers (n-propanol and isopropanol) on combustion engines are uncommon and mostly limited to spark ignition (SI) engines. Although the above short-chain alcohols’ high volatilities, low viscosities, and high oxygen contents could improve spray combustion and emissions, some disadvantages, such as high heat of transformation of vaporization, low heating values, and low cetane numbers (CNs), cause a lower proportion of applications in compression ignition (CI) engines.
Furthermore, when combined with diesel, a low alcohol concentration exhibits phase separation, cold start problems (i.e., ethanol-C2H5OH), stability issues, storage problems, and corrosion (i.e., methanol-CH3OH) [15]. Even in the case of diesel/biodiesel/ethanol ternary blends, the proportion of ethanol in the mix should be kept as low as feasible because NOx emissions from ternary blend combustion result in higher values than diesel fuel [16]. As a result, higher alcohol combined with diesel oil has recently received significant attention as an alternative fuel for diesel engines. Butanol (C4H9OH), pentanol (C5H11OH), hexanol (C6H13OH), octanol (C8H17OH), decanol (C10H23OH), dodecanol (C12H25OH), and phytol are all straight-chain alcohols with four or more carbon atoms (C20H41OH). Butanol and pentanol have lately gained significant attention as alternative fuels for diesel engines since they release fewer greenhouse gases and hazardous pollutants.

2.1. Methanol (CH3OH)

CH3OH is the simplest form of alcohol and is colourless, odourless, and poisonous. It is corrosive to some materials and is frequently created from a combination of sources, including synthetic gas (syngas), formic acid, formaldehyde, and methane. CH3OH is hygroscopic, which causes phase separation in CH3OH-gasoline blends [17]. Methanol is beneficial in the reduction of NOx and particles due to its high oxygen content and also its high latent heat of evaporation.
Methanol has some disadvantages as a fuel additive: low lubricity, low viscosity, low cetane number, and ignitability. An enhancer of the cetane number and lubricity is needed to solve this problem. According to Cheung et al. [17], biodiesel and methanol-blended fuels have a higher fuel consumption because they have a lower heat value than oxygenated fuels. Compared to methanol, ethanol has a higher energy density, is nontoxic, and is available in large quantities, reducing racers’ vulnerability to price increases and supply disruptions.
Additionally, when the proportion of methanol in mixed fuels increased, NOx emissions were reduced. Researchers concluded that methanol’s cooling impact dominated NOx generation in blended fuels. Because of the reduction in soot generation, methanol-mixed fuels produce fewer particulates [18].
Methanol as an alternative fuel can be produced in a variety of sources, including natural gas, biomass [19], and coking plants. During the continuous manufacturing of biodiesel with supercritical methanol [20], the gas [21] can be recovered by flash evaporation. Many synthesis techniques have emerged, including low-pressure synthesis, novel designs, and sophisticated reforming processes [18]. Biomass processing is the most profitable method for producing methanol from renewable resources [22,23]. Reno et al. [24] conducted a life cycle assessment on the production of methanol from bagasse, and the results showed that the production of methanol from bagasse is a viable alternative in terms of output–input ratio. Holmgren et al. [25] developed the synthesis chain due to the interdependence of the various stages of the biomass and methanol gasification process; the results can be used to calculate energy efficiency and greenhouse gas emissions economic indicators for independent and integrated methanol synthesis plants with biomass gasification. Clausen et al. [26] investigated the viability of a methanol production system based on biomass gasification and water electrolysis. Methanol is a substance with numerous advantageous features that make it valuable as an energy source, chemical raw material, component, or intermediate in a variety of consumer products. Replacing existing generation fossil fuels with cleaner fuels has the potential to reduce demand as well as the economic expenses associated with traditional increased pollution [27,28].

2.2. Ethanol (C2H5OH)

C2H5OH can be a clear, colourless, poisonous liquid with a distinct odour. It has a higher octane rating than gasoline, making it ideal for combining as a liquid fuel. Ethanol has less energy per volume than gasoline while denatured ethanol (98% C2H5OH) has around 30% less energy per volume than gasoline [29]. Because C2H5OH includes oxygen, using it as a gasoline addition reduces carbon monoxide gas emissions by up to 25% [30]. It is also soluble in both polar and nonpolar solvents, has a higher vapour pressure than gasoline, and has a 35% oxygen content. C2H5OH has a limited half-life in surface water and subterranean aquifers, reducing the risk to aquatic organisms significantly. Furthermore, even when it is spilt, the environmental impact is minor. Because ethanol dissolves swiftly in the natural environment, biodegradation occurs quickly in soil, groundwater, and surface water, with projected half-lives ranging from several hours to ten days [31].

2.3. Butanol (C4H9OH)

Butanol has mostly been employed in the cosmetics and pharmaceutical sectors as a solvent, chemical intermediary, and extractant as well as in the assembly of butyl acrylate and methacrylate [32,33,34]. When created biologically through the fermentation of starchy and sugar feedstock, this alcohol class is commonly referred to as biobutanol.
The use of ethanol blended with gasoline increases the octane rating, including drawbacks such as metal component corrosion and vapour lock. Such troubleshooting is frequently resolved by modifying the engine and equipment, but the addition of alcohols with large carbon numbers, such as biobutanol, allows use in the existing system without requiring any changes. Alcohols with high carbon contents are frequently produced from syngas via a catalytic process involving modified Fischer–Tropsch or methanol catalysts. In recent years, researchers have taken an interest in n-butanol as an alternative biofuel to bioethanol. Even though most studies and enterprises have focused on ethanol as a fuel rather than butanol [35], butanol may be a significantly superior direct option.
Because n-butanol has a lower Reid vapour pressure than ethanol, it is less evaporative/explosive [7]. Because the air–fuel ratio and energy content of n-butanol are similar to those of gasoline, it is commonly mixed with gasoline in larger ratios than ethanol for use in existing cars without modification [8]. Furthermore, the high octane number makes the alcohol more appropriate for use in combustion engines. A fuel with a lower octane level is more prone to knocking (very quick and compression ignition) and can reduce efficiency. Engine damage might also result from knocking. Unlike other alcohols, the Environmental Energy Company (US) proved that n-butanol is frequently utilized as a full substitute for gasoline [10].
Although ethanol is promising as a biofuel, recent studies have shown that ethanol is a serious problem that must be solved before it can be used in engines. It is corrosive and can cause defects in the existing fuel line layout and injection system [36]. Ethanol also has a very low flash point and evaporation capacity, so extra precautions need to be taken when using it [37].
In addition to these shortcomings, ethanol can also reduce local air quality, thereby endangering human health. Compared with ethanol, butanol is a more promising choice [36,38,39]. For carbon atoms, the atoms can form a straight or branched chain [40]. The more carbon atoms, the higher the energy content of the alcohol fuel. In addition, compared with ethanol, butanol has less toxicity and higher energy consumption [36,41]. It can also be made from cellulose waste from agricultural products or even fibre residue from sugar factories [42,43].
Compared with ethanol, butanol is much less corrosive, which allows it to be transported through existing traditional fuel transportation systems [36,44]. Since butanol is less corrosive, it can also be used in automobiles without modification [45,46]. If it is contaminated by water, it is difficult to separate from the base fuel (gasoline/diesel). It also contains energy units similar to gasoline, but the specific energy per litre is 25% higher than that of ethanol [36]. The performance results in higher engine output, so the vehicle can easily increase butanol consumption. As we all know, due to the complexity of manufacturing, long-chain alcohol fuels have not been widely used; today, the production cost of butanol is gradually decreasing; rapid progress in production technology has made it possible to produce more butanol [47]. Butanol in gasoline engines is not as extensively studied as methanol or ethanol, despite the increasing number of publications. The following article examines the power, combustion, and performance effects of mixing butanol with gasoline. There are several loopholes in adding butanol to gasoline engines that deserve further study based on the comments in [47].

2.4. Pentanol (C5H11OH)

Pentanol blended with conventional fuels has fewer greenhouse gas emissions, leading it to gain a lot of attention as an alternative fuel for diesel engines. Because of pentanol’s high potential as a blending component and its physicochemical features, binary blends, such as diesel/pentanol and biodiesel/pentanol, as well as ternary blends, such as diesel/biodiesel/pentanol, have been researched primarily within conventional diesel. Diesel, biodiesel, pure vegetable oil, and pentanol quaternary mixes were also studied. There is limited information available in the literature about the use of pentanol in advanced compression ignition (CI) engines. Diesel/pentanol mixes with exhaust gas recirculation (EGR) technology could lower NOx and soot emissions from the CI engine at the same time. Furthermore, diesel/pentanol blends emitted more CO and hydrocarbon (HC) emissions than diesel oil. However, by combining the cetane improver with the blends, CO and HC emissions were dramatically reduced. Up to 45–50% n-pentanol/diesel blends can frequently be used safely in diesel engines with no engine modifications or additives. However, in terms of kinematic viscosity, lubricity, and oxidation stability of diesel/pentanol blends, pentanol concentrations should be kept below 10%. In general, investigations using biodiesel/pentanol blends in CI engines revealed a reduction in CO, HC, NOx, and smoke emissions when compared to straight biodiesel. Researchers focused on diesel/biodiesel/pentanol blends for ternary blends. The emission characteristics of CO, HC, NOx, and smoke opacity showed distinct trends that were consistent with the pentanol fraction in the ternary blends, particularly in [31].

2.5. Production of Bioalcohols

Alcohol fuels are frequently generated from all available organic elements, such as gas, coal, biomass, and organic waste. Alcohol fuels are primarily made from corn and sugar cane as raw materials, although the synthesis and manufacturing of alcohol fuels from nonfood crops and agricultural leftovers is a current focus. Energy crops, cellulosic residues, and wastes are examples of nonfood lignocellulosic biomass. Grain-based ethanol has been tried as the main generation to vary to second-generation cellulosic ethanol and other advanced cellulosic biofuels. After the 2000s, cellulosic ethanol was identified as a crucial biochemical method for converting biomass to fuels [48]. Third-generation algae-based feedstock for alcohol fuels emerged as a possibility as a big staple for the future alcohol fuel sector.
Figure 1 depicts the generations of raw feedstock used in the production of alcohol fuel as well as the most visible material used in different countries.
There is a clear difference between developing and developed countries in terms of priority selection, but the underlying notion should be the same: use locally accessible, underutilised feedstock, and choose feedstock that has a tipping fee to treat the feedstock as municipal/industrial wastes. However, when wastes are used as feedstock, it should be highlighted that the not-in-my-backyard (NIMBY) problem is now a common occurrence in nearly every country. The European Commission recently voted against the use of biofuels synthesised from biomass derived from food crop sources by 2030. For the timely commercialization of cellulosic bioethanol, a second-generation bioalcohol, intensive interdisciplinary efforts are expected.
Ethanol is made by either using coal and biomass syngas or synthesizing it from petroleum-based ethylene. It can also be obtained by fermenting sweet substances.
On the other hand, bioethanol is produced by fermenting regenerative biomass or through the procedures of distillation and purification. The United States and Brazil are the world leaders in the production of bioethanol, corresponding to roughly 85% of global production in 2021 [50]. The main feedstocks used for the production of bioethanol are corn in the United States and sugarcane in Brazil. Bioethanol Figure 2 and Figure 3 show the overall manufacturing process of bioethanol. Pretreatment involves converting biomass into alcohol fuels. The first step to preparation is reducing the size and uniformizing the density/size. Often, steam, hot water, or slight carbonization is used as a pretreatment. Basically, saccharification (hydrolysis) is the process of breaking down cellulose/hemicellulose to produce sugars, such as glucose and xylose. Based on the concentration ratio and the adsorption ratio of the component enzymes, three distinct cellulase enzymes perform the overall hydrolysis. In fermentation, sugar and starch are converted into ethanol by microorganisms. Concentrating and cleaning the ethanol produced by distillation is called distillation and rectification and last is drying.
Agricultural waste often comprises a high concentration of alkali metals (potassium and sodium) and other inorganic elements, such as calcium, magnesium, and, in rare cases, chlorine and sulphur. When converting these wastes to alcohol fuels using thermal techniques, alkaline metal components act to supply low-melting salts, causing clogging and other ash-related difficulties. In contrast, the fermenting approach can lessen the likelihood of ash issues, which may be a desirable component of the manufacturing process. Rice husk, in particular, contains over 90% ash, and rice straw contains almost 30% silica, though this varies depending on rice stock, climate, and geographical context.
These inorganic materials act as a barrier to the thermal conversion process, and fermenting is frequently a better technique to transform this biomass fuel. To supply ethanol, starch and carbohydrates are employed as first-generation raw materials. In 2013, about 90% of bioethanol was created from starch and carbohydrate. Major crops include corn, grain, and cassava (Figure 4). The degradation of the environment during crop cultivation and ethanol production as well as the exploitation of vital food resources as fuel production are disadvantages. As a result, significant agricultural countries such as the United States, Brazil, and China are currently major producers of biofuels, including alcohol fuels. In the United States, 95% of ethanol is derived from starch [52].
As the production of bioethanol from grain-based raw materials has recently been constrained, second-generation bioethanol production from non-grain-based biomass is now gaining a constantly growing importance. Bioethanol is quickly becoming a viable alternative to gasoline, and it is often made from the starch of corn and cassava or the sugary contents of sugar cane and sugar turnip. Bioethanol is also made from crop wastes that contain lignin and cellulose. Sugar and starch are easily converted to bioethanol, but their supply and cost are limited. As a result, efforts are being made to investigate various procedures for producing bioethanol from lignocellulose-based raw materials to capitalize on their vast supply in nature while also ensuring market viability [29].
Kandasamy et al. [53] investigated the ability to supply alcohol from lignocellulosic material; nevertheless, while not a protein chemical reaction degradation, the ultrasound technique produces more alcohol than that normally described in the literature [54].
According to the authors, the bulk of plastic alcohol production units in the EU are only on a pilot or demonstration scale. According to Cotana et al. [55], in the context of a greater focus on the assembly of second-generation bioethanol, the use of lignocellulosic feedstock, monocot genus australis (common reed, which could be a perennial grass growing in wetlands or close to upcountry waterways), is an especially good option. In the mentioned research, a steam explosion was used to optimize the bioethanol production process from the monocot genus australis. The entire approach had a potency of sixteen and a 56 g ethanol/100 g staple. Eggert et al. [56] conducted a study to promote second-generation biofuels, specifically cellulose-based organic chemistry ethanol; additionally, the authors explore relevant regulations that could aid in the fight for those fuels. They believe that first-generation biofuels have been and continue to be heavily supported, which has contributed to the increased production and usage of such fuels.
The question is whether primary-generation biofuels pave the way for second-generation biofuels. Thus, they have the advantage that even if favourable conditions for innovation and scale efficiencies are created, necessary value reductions may not be realized. The GHG emissions from land-use change, which are linked to the large-scale increase of plastic feedstock, may counterbalance the gains from dynamic fuel.
Switchgrass (a warm-season grass) looks to have a high potential yield in marginal areas, making it an excellent nonfood bioenergy feedstock for bioethanol production in China [57]. Hansdah et al. [58] discussed the potential for producing bioethanol from the Madhuca indica flower, which is a tropical forest tree that grows abundantly in Asia and Australia. Madhuca indica seeds, on the other hand, are a potential feedstock for biodiesel production in India. Chen et al. [59] examined Landoltia punctata (a worldwide hydrophyte strain with the ability to collect starch) and its application as a novel feedstock for bioethanol production. To increase ethanol production, pectinase pretreatment was used, which resulted in the highest alcohol concentration reported to this date and a mistreatment hydrophyte because of the feedstock. Moreover, the invention of recent raw materials (filamentous fungi) and also the improvement of the method of obtaining bioethanol by integrating first- and second-generation alcohol production processes are described in [60], which that reports a discount of energy consumption by a pair of 5% and a rise of alcohol production by four-times. Connolly et al. [61] have elaborated different ways of manufacturing transport fuels, including alcohol and alcohol in terms of the resources and conversion processes used and the transport demands met. In keeping with the results of this study, the best answer at the present consists of liquid fuels, for instance, alcohol or dimethyl ether. Biomass chemical action and electrolysis area unit technologies require improvements. As introduced before, the third generation of bioethanol feedstocks area unit protoctist contains lipids, proteins, carbohydrates/polysaccharides, and skinny plastic walls. In keeping with [32], while “algal lipids area unit principally extracted and reworked into biodiesel, the left-over cake of starch (the storage component) and polysaccharide (the skinny wall component) may be reborn into bio-ethanol”. Moreover, the ensuing combined biodiesel/bioethanol method additionally offers the advantage of manufacturing CO2 fermentation, which might be captured and used.

2.6. Properties of Bioalcohols

Alcohols can be produced from fossil fuels and lignocellulosic biomass via synthesis gas. These alcohols can be used as neat or blended in current engines. From the early years of research in the area of alternative fuels, alcohol was evaluated as a fuel alternative for diesel vehicles [17]. Methanol (CH3OH), ethanol (C2H5OH), propanol (C3H7OH), butanol (C4H9OH), and pentanol (C5H11OH) are five alcohols that will be reviewed in this section. Short-chained alcohols such as ethanol, propanol, and butanol have not been considered neat fuels for CI. However, they can be used as blends with fossil diesel fuels or biodiesel [62,63]. Generally, alcohols in CI engines reduce CO2 and PM emissions with some changes in their cetane index and combustion process. Alcohol-based fuels can easily be used as blends in a diesel without any major changes and modifications to the engine [64].
Before using any new alternative fuels in an engine, it is fundamental to investigate their chemical and physical properties. This analysis shows how it is possible to use these new fuels without any major modification or property enhancer. The advantages of alcohol application in an engine regarding its reduction of soot formation and particle emissions make it more attractive than other alternative fuels.
In the 1970s in South Africa and in the 1980s in Japan, Germany, and the United States, ethanol was used as a fuel in diesel engines [62]. Ethanol combustion in diesel engines led to reductions in particulate matter emissions. Ethanol has low ignitibility, low lubricity, limited miscibility, and a low heating value, which makes its blending proportions limited. The best solution to use ethanol as an additive is to blend it with biodiesel to improve the stability, cetane number, and lubricity of blends [65]. Lapuerta et al. [63] reported that biodiesel fuel acts as a stabilizing component in ethanol-diesel bends. A higher ethanol content in blends increases the total hydrocarbon emissions, decreases smoke emissions, and may present a nuclei mode because the high THC emissions lead to the nucleation of hydrocarbons. Researchers concluded that the effect of the oxygen content in these blends on the particle size depends upon its functional group. Ethanol has low energy density and high hygroscopicity in comparison to gasoline.
Methanol is an alternative, renewable, and environmentally and economically attractive fuel; it is considered to be one of the most suitable fuels for traditional fossil fuels. Recently, methanol has been used as a substitute for traditional internal combustion engine (IC) fuel to solve some environmental and economic problems. Significant progress has been made in recent major research projects that is worth reporting. Methanol is considered to be one of the most suitable fuels for engines, for example: (a) it can be used in gasoline engines with high compression ratios and can replace diesel in some professional applications; (b) it can be used in intake manifold injection engines; (c) it can be used in high compression SI direct injection engines; (d) it can be used in direct injection engines; and (e) it can be used in medium power and high compression turbine engines with port injection [66,67,68,69,70].
Propanol as a possible alternative fuel can be produced from biomass through various kinds of processes. It has a higher energy density and cetane number compared to ethanol and methanol [71]. Low carbon chain length alcohols, such as methanol and ethanol, have been used extensively in SI engines because of their good antiknock properties and lower gaseous emissions, such as CO and THC [72,73,74]. Short-chain alcohols with a low cetane number in diesel engines result in a significant increase in the duration of ignition delay [75]. The addition of ethanol or butanol has been observed to have a longer ignition delay and a significant reduction in PM formation [76] because an increase in fuel ignition delay allows more time for fuel and air mixing before the start of combustion and leads to fewer fuel rich-zones and a decrease in PM emissions [75]. These fuels are not of good compatibility with diesel fuels due to their poor blending stability, poor cetane number, and lower heating value.
Butanol is a promising fuel with several advantages over ethanol, such as higher energy content, less corrosive, better miscibility than current fuels, and lower water absorption [77,78]. The same production process for ethanol can be adapted to produce butanol from different feedstocks. There are other technologies, such as fermentation and transesterification [79], with lower required energy, which lead to the production of desirable alcohols, such as butanol [71]. The THC emissions from ethanol blended with diesel are higher than butanol blended with diesel due to the higher heat of vaporization of ethanol [77] Longer chain alcohols are more suitable for blending with diesel fuel. They have proper lubricity, high calorific value, lower volatility, a high flash point, high viscosity, and cetane number compared to lighter alcohols [69,71].
It has been discussed that pentanol, butanol, and other higher alcohols have a low affinity for water and a higher energy density and fit well into current vehicles without any needed modification to the vehicle’s engine [80]. The literature has shown that pentanol has good physical properties as an alternative fuel [80]. Pentanol is also an excellent alternative fuel because it can be produced from biological processes such as microbial fermentation using micro-organisms and biosynthesis from glucose [71]. Pentanol, a long-chain alcohol, requires even less production energy than butanol compared to other alcohols [81].
Heavier alcohols are more advantageous than light-chain alcohols due to their better blending stability, hydrophobic properties, higher cetane numbers, heating value, and low carbon counterparts [82,83,84]. Butanol as a blend with diesel fuel has been extensively studied in diesel engines [85,86,87,88,89]. It was concluded that the alcohol content in a diesel engine increased ignition delay due to its low cetane number. In-cylinder pressure and heat release rate increased due to an enhanced premixed combustion phase. The presence of oxygen improved the combustion process and reduced soot formation. Pentanol’s similar properties to diesel and also the fewer emissions produced together with good engine performance make it advantageous as a short-term replacement in diesel engines. It has shown more advantages than butanol and other light alcohols [81,90].
The use of pentanol as an alternative fuel in a diesel engine is not as well-known as butanol, so it needs more investigation due to its properties being similar to diesel fuel. Generally, the low cetane number of alcohols increases ignition delay and so leads to a higher combustion temperature and then higher NOx emissions [77]. However, the higher heat of vaporization in alcohols reduces the cylinder flame temperature and so reduces NOx emissions [91]. It can be concluded from the different emissions analyses in the literature that engine operating conditions and injection strategies could affect the emission characteristics. A lower carbon-to-hydrogen ratio and the presence of oxygen in alternative oxygenated fuels lead to decreasing soot emissions from internal combustion engines [69]. According to Pinzi et al. [91], there is an increasing interest in employing long-chain alcohols, such as butanol and pentanol, in the transportation industry due to their potential production from residual biomass via fermentative methods. There is evidence that incorporating alcohol into diesel fuel allows for the well-known smoke-NOx trade-off to be overcome in steady-state compression ignition engine running. However, the influence of long-chain alcohols under engine transient conditions is unknown, and their behaviour during the Worldwide Harmonized Light Vehicles Test Procedure (WLTP) cycle has not been recorded. The purpose of a recent study was to fill the research gaps mentioned above by characterising the noise and exhaust emissions (CO, total hydrocarbon content or THC, NOx, and particulate matter or PM mass, PM number, and PM distribution) of a diesel engine running on ultralow sulphur diesel (ULSD) fuel and its mixtures with 1-pentanol and 1-butanol under stationary and transient conditions (WLTP). Dekati electrical pressure low impactor ELPI + paired with Dekati Fine Particle Sampler FPS-4000 monitored transient PM quantity, mass, and size distribution while Horiba Mexa 7100D assessed transient gaseous emissions. Increasing the long-chain alcohol concentration in gasoline blends reduces PM quantity and mass greatly in both stationary and WLTP tests, which is mostly due to the oxygen content of 1-butanol and 1-pentanol.
The addition of long-chain alcohols reduces NOx emissions in stationary operation whereas the reverse tendency was observed under WLTP. The use of greater alcohol/ULSD fuel combinations appears to marginally increase noise levels. Overall, it is possible to infer that the use of greater alcohol/ULSD fuel mixes appears to be a better substitute for plain ULSD fuel [92].
Today, the scientific community is showing increasing interest in studying n-butanol as part of the mixture. Compared with ethanol, the safety of biobutanol in fuel transportation, processing, and storage [93] as well as its higher cetane index [94], higher calorific value [95], lower volatilization [96], higher flash point, improved lubricity [90], and better compatibility with diesel fuel (especially at low temperatures) have contributed to this interest. A summary of the some physical and chemical properties of alcohols fuels are listed in Table 1. Fernández-Rodrguez et al. [97] investigated and discovered, as in Figure 5 below, which shows scientific publications published in the previous few years linked to various fuels in diesel engines, ethanol in diesel engines and butyl alcohol in diesel engines. Schematic Diagram 5 demonstrates that the scientific community is aware of the eye-catching opportunities and situations resulting from the use of biofuels in diesel engines in road freight transportation, agricultural applications, such as tractors, harvesters, and self-propelled sprinklers, as well as cogeneration applications. Furthermore, Figure 5 indicates the growing interest in n-butanol as a mixing element for diesel engines, as evidenced by a growing number of papers in recent years.
Table 2 shows the main physical properties of C2H5OH and C4H9OH. However, the physiochemical properties of n-CH4H9OH are highly similar to those of diesel than to those of C2H5OH; it still cannot replace diesel at 100% [37]. The literature reports that alcohol–diesel blends are tested up to 40% butanol content (volume basis) while not requiring engine modifications [98,99].

3. Effects of Bioalcohols on Emissions

The physical properties of the particles in this study are related to the quality, quantity, and size of the particles, the microstructure and agglomerates of the particles, and the nanostructure of the primary particles. Particulate matter is composed of soot and other liquids or solids, and soot accounts for the largest part of diesel particulate mass. Soot is usually a solid that contains carbon as a byproduct of fuel combustion. Various processes include fuel pyrolysis, nucleation, coalescence/cohesion, surface growth, coalescence, and oxidation. The first five steps from fuel pyrolysis to agglomeration must be carried out in sequence, and the oxidation process can be carried out at any time during the soot formation process [103].
In diesel engines, alcohol, especially methanol, has attracted more and more attention as an oxygen-rich and renewable fuel. Methanol fuel can be mixed with diesel in the engine; it can be mixed with diesel (85% methanol and 15% diesel) or other percentages. Huang et al. [104] examined the main combustion characteristics of the methanol/diesel fuel mixture by analysing the pressure in the cylinder of a compression ignition engine. The results show that an increase of the methanol mass fraction in the methanol/diesel fuel mixture leads to an increase in the heat release rate of the premixed combustion stage and a decrease in the combustion duration and diffusion combustion stage.
Sayin et al. [105] investigated the influence of injection sites on exhaust emissions from direct injection (DI) single-cylinder four-stroke diesel engines running on a diesel/methanol mixture. The results demonstrate that using methanol-based diesel fuel reduces flue gas opacity, CO, and UHC emissions by 5–22%, 33–52%, and 26–50%, respectively, while reducing CO2 and NOx emissions by 14–68%. Depending on engine operating circumstances, this might range from 22 to 69%. Chao et al. [106] investigated the influence of methanol-containing additives (MCA) on carbonyl compound emissions from diesel engines. According to the findings, applying 10% and 15% MCA increased the emission factors of carbonyl compounds (CBC), acrolein, and isovaleraldehyde by at least 91%.
Zhang et al. [107] investigated the effects of controlled and unregulated exhaust emissions in four-cylinder direct-injection diesel engines powered by Euro V diesel and fumigated methanol. Fumigated methanol was injected under various engine circumstances to fulfil the engine load of 10%, 20%, and 30%. Methanol significantly increases HC, CO, and NO2 emissions while decreasing NOx. The use of methanol and diesel in combination is an efficient way to minimize PM and NOx emissions from diesel cars in operation [108].
Cheng et al. [109] discovered that fumigated methanol can be utilized to reduce diesel consumption in part-load diesel engines. The greatest methanol utilization was discovered to be 43% of total fuel consumption. At low loads, BFC diminishes as the amount of fumigated methanol increases. It increases with the amount of fumigated methanol in the case of heavy exposure [110]. Lee et al. [108] developed an improved multidimensional model and detailed chemical kinetics mechanism to study the combustion and emission characteristics of controlled compression ignition (RCCI) engines operating with methanol/diesel. It has been found that burning methanol/diesel fuel can reduce emissions and improve fuel efficiency. Sayin et al. [111] studied the effects of injection pressure and injection time on the performance and emission characteristics of DI diesel engines using methanol (5%, 10%, and 15%) as mixed diesel fuel. The results show that as the proportion of methanol in the fuel mixture increases, BSFC, specific braking energy consumption (BSEC), and NOx emissions increase as BTE, smog, CO, and THC decrease. It was also found that the increase in injection pressure and timing resulted in a decrease in smoke, CO, and THC emissions while increasing NOx emissions.
Zhang et al. [112] studied the effect of methanol fumigation on the combustion and particulate emissions of diesel engines under various engine loads and fumigation levels. It was found that the fumigation of methanol would increase the ignition delay, but it would not significantly affect the combustion duration. Aeration with methanol can effectively reduce the particle mass and particle number concentration. Geng et al. [113] studied particulate matter emissions (PM) using the same engine; the results showed that PM number and concentrations were significantly reduced at low and medium loads and increased when the engine was operating at high loads. Wang et al. [114] assessed the working range and combustion behaviour of diesel engines treated with methanol and studied these in experimental tests. It was found that the operating range of methanol/diesel differs in load and methane displacement percentage; this is achieved in a load range of 6% to 100%. The scope of operation is limited by discovery: partial combustion, misfire, roaring combustion, and explosion.
Nagafi and Yusaf [115] studied the effect of a methanol-diesel mixture on diesel engine performance. In this study, a diesel engine was tested in which diesel fuel was mixed with methanol at specific mixing ratios of 10:90, 20:80, and 30:70. Experimental results show that the effective power and torque of diesel are less than that of diesel fuel mixed with methanol. The optimal mixing ratio to produce the lowest exhaust temperature is 10% to 90% methanol. Diesel engine exhaust temperature is higher than any mixed fuel mixture. A mixture of 30% methanol and 70% diesel reaches the minimum BSFC value. Compared with any mixing ratio, the BSFC of diesel is much lower. Under almost all operating conditions where diesel fuel is mixed with methanol, BTE improved.
Zhang et al. [112] studied the effect of methanol and ethanol bleed on gas and particulate emissions in a four-cylinder DI diesel engine. Under various engine operating conditions, methanol or ethanol was injected to complete the engine loads of 10% and 20%. Diesel and fumigated methanol or ethanol can reduce nitrogen oxides and diesel particles and mass, and methanol fumigation is more effective, such as in fumigation with ethanol to reduce particles. Sayin et al. [116] studied the effects of methanol–diesel (M5, M10) and ethanol–diesel (E5, E10) mixtures on the performance and exhaust emissions of a single-cylinder four-stroke diesel engine with a naturally aspirated engine. Direct injection results show that when methanol–diesel and ethanol–diesel mixtures are used, BSFC and NOx emissions increase while BTE, smog, CO, and THC emissions decrease.
After consulting the available literature on the effect of alcohol fuel (as most studies have proved and summarised in Table 3 and Table 4) on the PM emissions of internal combustion engines, we can draw the following main conclusions (presented in Table 5):
  • Both lower alcohols and higher alcohols can significantly change the composition and structure of fine dust, which leads to an average reduction of about 50% (weight) of fine dust, 60% (weight) of fine dust (TNC), and 30% (by weight) under various conditions of fine dust size (GMD). Under the same alcohol consumption (by volume or weight), lower alcohols, especially methanol and ethanol, are more effective than higher alcohols in reducing PM. This is because compared with the characteristics of higher alcohols, the difference in characteristics of lower alcohols significantly affects the formation of PM (especially when the carbon content is low, and the oxygen content is high).
  • Mixed alcohol has a greater impact on PM reduction than fumigated alcohol, especially when running at low engine temperature (i.e., low speed or low torque). Evaporation of fumigation alcohol can cause poor combustion. In addition, part of the fumigated alcohol droplets will enter the low-temperature cooling area of the combustion chamber wall/tank, resulting in unburned fuel. In addition, spraying alcohol with a low-pressure fuel injection will produce large drops of alcohol, reducing the chance of complete combustion and further particle formation.
  • Because alcohol has a lower weight, quantity, size, elemental carbon (about −60%) and flash point (about −8%) as well as higher fine dust oxidation reaction activity (about increased performance and service life) it can be used in a catalytic converter, especially a DPF.
Table 3. Reviewed investigations available in the literature about the PM physicochemical properties of internal combustion engines.
Table 3. Reviewed investigations available in the literature about the PM physicochemical properties of internal combustion engines.
Reference FuelAlcohol UsedFuelling TypeEngine TestedOperating ConditionsTest ResultsRef.
DieselMethanolFumigation (achieving 10, 20, and
30% of desired engine load by
fumigated methanol)
4-cylinder, NA,
diesel engine
Japanese 13-mode test cycleDecrease in PM mass (up to
−21.2%, on average of different
conditions for 30% fumigated
methanol) with increase in
methanol ratio
[117]
DieselEthanolBlend (mixing 5, 10, 15, and 20%
of ethanol by volume with diesel)
6-cylinder, heavy-duty, turbocharged
diesel engine
European 13-mode test cycleDecrease in dry soot but increase
in SOF resulting in no obvious
effect of ethanol on the PM mass
[118]
DieselButanolBlend (mixing 5, 10, 15, and 20%
of butanol by volume with diesel)
Single-cylinder, DI
stationary diesel
engine
25, 50, and 75% of
maximum engine power at a
speed of 3000 rpm
Decrease in PM mass (up to
−25%) with increase in butanol
ratio
[119]
Dieseln-PentanolBlend (mixing 10, 20, and 30% of
n-pentanol by volume with diesel)
4-cylinder, NA, DI
diesel engine
Five engine loads (BMEP of
0.84, 2.06, 4.15, 6.20, and
7.10 bar) at a speed of 1800
rpm
Decrease in PM mass (up to about
−70%) with increase in pentanol
ratio (on average of five loads,
−42% reduction for 30%
pentanol blend)
[120]
Table 4. Summary of the main parameters of the reviewed investigations available in the literature about the PM physicochemical properties of SI and CI engines.
Table 4. Summary of the main parameters of the reviewed investigations available in the literature about the PM physicochemical properties of SI and CI engines.
Fuel UsedEngine TestedParameter InvestigatedRef.
Gasoline; dieselSI and CIPM mass; TNC; GMD; nanostructure of particles; OC-EC; ions; metal and elements[121]
Gasoline; dieselSI and CIMicro- and nanostructure of particles[122]
EthanolSI and CIPM mass; PAHs[123]
MethanolSI and CIPM mass; PAHs[123]
DieselCIPM mass; TNC; GMD; morphology, micro- and nanostructures of particles; OC-EC; metals and elements; ions; oxidation reactivity[107]
Alcohol (methanol, ethanol, butanol, and
pentanol)
CISoot; PAHs[124]
Table 5. Summaty of the reviewed investigations available in the literature about the PM physicochemical properties of diesel engines.
Table 5. Summaty of the reviewed investigations available in the literature about the PM physicochemical properties of diesel engines.
Reference FuelAlcohol UsedFuelling TypeEngine TestedOperating ConditionsTest ResultsRef.
DieselCH3OHBlend (mixing 11.5% methanol by volume with 88.5% diesel)4-cylinder, turbocharged diesel engineOne engine load (BMEP of 12 bar) at a speed of 1400 rpmDecrease in the radius of gyration (−25%), fractal dimension (about −10%), number of primary particles (−51%), and primary particle diameter (7.5–42.5 nm for methanol and 9.5–50 nm for diesel)[86]
DieselC2H5OHBlend (mixing 10 and 20% of ethanol
with diesel)
4-cylinder,
supercharged
diesel engine
One engine load (full load) at
full engine speed
Decrease in the agglomerates and
primary particles sizes with an increase in ethanol ratio
[114]
Dieseln-ButanolBlend (mixing 10 and 20% of butanol
by volume with diesel)
4-cylinder,
turbocharged
diesel engine
One engine load (78.4 Nm) at
a speed of 2000 rpm
Decrease in the agglomerates and
primary particles sizes
[104]
Dieseln-PentanolBlend (mixing 15 and 30% of pentanol
by volume with diesel)
4-cylinder,
turbocharged
diesel engine
One engine load of 130 Nm at
a speed of 2000 rpm
No significant differences in the overall morphology of particles (near-spherical shape) and agglomerates (branched-chain structures) for all the fuels tested; decrease in primary particles size (up to − 8.2%) with an increase in pentanol ratio[105]
In summary, in the future, zero-emission nonfossil energy vehicles will be fully introduced into today’s fleet. However, the specific time is not clear. The combination of alcohol fuels, especially low-alcohol fuels, and emission catalysts may be effective. Methods to reduce the consumption of fossil fuels and emissions while reducing particulates, as illustrated in Figure 6, thereby improving the performance and service life of emission catalysts, must be developed.

4. Conclusions

There is no extraordinary super fuel that can meet all the requirements for the economy, maximum thermal efficiency, and engine performance while remaining clean enough to protect the environment. Each fuel has its advantages and disadvantages, and the choice of specific fuel depends on various parameters. The physical properties of fuel are shown in Table 1. If we start with the advantages of alcohol fuels, they can be summarized as follows:
  • Compared to methanol, ethanol has a higher energy density, is nontoxic, and is available in large quantities, reducing racers’ vulnerability to price increases and supply disruptions. Methanol can be made from organic materials, such as biomass and domestic waste; at some point, it may even be made from coal. The top five countries in the world’s coal reserves are indeed the US, Russia, Australia, China, and India, and these reserves will be very abundant in the next few years. As compared with gasoline, the molar ratio of alcohol combustion products to reactants is higher, so higher combustion pressure is generated in the combustion chamber of the internal combustion engine, thereby improving energy production and thermal efficiency [125].
  • Because alcohol fuel has higher volumetric efficiency, alcohol has better combustion characteristics and performance, making methanol the preferred fuel for racing cars. The acceleration time decreases as the power increases.
  • Alcohol has better visibility in the event of a fire. It is used for evacuation, rescue, minor suffocation, cold flames, and low heat, resulting in minor burns, minor smoke damage, and easy removal of residues. Water can be used to extinguish fires, and it is easier to use powder foam.
  • Alcohol fuel has lower vapour emissions. When alcohol fuel is used, less harmful byproducts are released into the atmosphere.
  • Because the carbon content in alcohol fuel is very low, the combustion of internal combustion engines will produce a small amount of soot, which is discharged into the atmosphere.
  • Alcohol fuel is liquid, so it only needs to be modified slightly to use the same transportation and infrastructure management as traditional fuels.
This review paper focused on presenting the performance, combustion, and emission characteristics of CI engines using alternative alcohol-based fuels. The effects of alcohol fuels on the performance, combustion characteristics, and emissions, such as NOx, CO, HC, and CO2, were also investigated. This work also examined the PM emissions of CI engines when using conventional fuels and when alcohol was introduced. In general, to reduce PM emissions or reduce the harm to human health and the environment, it is necessary to determine their chemical properties. Since the internal combustion engine is the source of particulate matter in the atmosphere, it is better to understand the chemical properties of particulate matter, such as organic carbon, elemental carbon, inorganic ions, metallic elements, polycyclic aromatic hydrocarbons (PAH), and oxidation reactions in internal combustion engines and integrated circuit motors. TC consists of OC and EC, which is harmful to the environment and human health. OC is considered to be carcinogenic for humans because it consists of some nonvolatile hydrocarbons [37]. Due to the European Union affecting air pollution, the global carbon cycle, and OC’s ability to absorb solar radiation, which leads to energy redistribution and climate change [126,127,128,129], some technologies are needed to reduce CO and EC emissions from electronic injection engines.
Biofuels, such as bioalcohols, are promising fuels that could be partially used in CI engines aiming to improve some combustion and performance characteristics of engines as well as to either reduce or control the exhaust emission and the particulate matter characterization of diesel engines. These biofuels can be used without requiring major modifications to the engine and could contribute to the dependence on fossil diesel. The compiled and detailed information about alcohol-based fuels is beneficial for conventional and also hybrid diesel vehicles to reach the forthcoming emissions regulations.

Author Contributions

Conceptualization, M.D.T., F.A.T., A.M.A. and O.D.; methodology, M.D.T., F.A.T. and O.D.; formal analysis, M.D.T. and A.M.A.; investigation, M.D.T.; writing—original draft preparation, M.D.T., F.A.T. and A.M.A.; writing—review and editing, M.D.T., F.A.T. and O.D.; project administration, P.K.T., Y.B. and A.M.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Acknowledgments

In this section, you can acknowledge any support given which is not covered by the author contribution or funding sections. This may include administrative and technical support, or donations in kind (e.g., materials used for experiments).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. International Energy Agency (IEA) World Energy Balances: Overview. Available online: https://www.iea.org/reports/world-energy-balances-overview/world (accessed on 21 February 2022).
  2. Sadeghinezhad, E.; Kazi, S.N.; Badarudin, A.; Oon, C.S.; Zubir, M.N.M.; Mehrali, M. A Comprehensive Review of Bio-Diesel as Alternative Fuel for Compression Ignition Engines. Renew. Sustain. Energy Rev. 2013, 28, 410–424. [Google Scholar] [CrossRef]
  3. International Energy Agency (IEA) Global Energy Review 2021: Oil. Available online: https://www.iea.org/reports/global-energy-review-2021/oil (accessed on 14 March 2022).
  4. Arbab, M.I.; Masjuki, H.H.; Varman, M.; Kalam, M.A.; Imtenan, S.; Sajjad, H. Fuel Properties, Engine Performance and Emission Characteristic of Common Biodiesels as a Renewable and Sustainable Source of Fuel. Renew. Sustain. Energy Rev. 2013, 22, 133–147. [Google Scholar] [CrossRef]
  5. Doustdar, O.; Wyszynski, M.L.; Mahmoudi, H.; Tsolakis, A. Enhancing the Properties of Fischer-Tropsch Fuel Produced from Syngas over Co/SiO2 Catalyst: Lubricity and Calorific Value. IOP Conf. Ser. Mater. Sci. Eng. 2016, 148, 012092. [Google Scholar] [CrossRef]
  6. Prado, C.M.R.; Antoniosi Filho, N.R. Production and Characterization of the Biofuels Obtained by Thermal Cracking and Thermal Catalytic Cracking of Vegetable Oils. J. Anal. Appl. Pyrolysis 2009, 86, 338–347. [Google Scholar] [CrossRef]
  7. Sadeghinezhad, E.; Kazi, S.N.; Sadeghinejad, F.; Badarudin, A.; Mehrali, M.; Sadri, R.; Reza Safaei, M. A Comprehensive Literature Review of Bio-Fuel Performance in Internal Combustion Engine and Relevant Costs Involvement. Renew. Sustain. Energy Rev. 2014, 30, 29–44. [Google Scholar] [CrossRef]
  8. Duarte Souza Alvarenga Santos, N.; Rückert Roso, V.; Teixeira Malaquias, A.C.; Coelho Baêta, J.G. Internal Combustion Engines and Biofuels: Examining Why This Robust Combination Should Not Be Ignored for Future Sustainable Transportation. Renew. Sustain. Energy Rev. 2021, 148, 111292. [Google Scholar] [CrossRef]
  9. Ali, O.M.; Mamat, R.; Abdullah, N.R.; Abdullah, A.A. Analysis of Blended Fuel Properties and Engine Performance with Palm Biodiesel–Diesel Blended Fuel. Renew. Energy 2016, 86, 59–67. [Google Scholar] [CrossRef] [Green Version]
  10. Ashraful, A.M.; Masjuki, H.H.; Kalam, M.A.; Rizwanul Fattah, I.M.; Imtenan, S.; Shahir, S.A.; Mobarak, H.M. Production and Comparison of Fuel Properties, Engine Performance, and Emission Characteristics of Biodiesel from Various Non-Edible Vegetable Oils: A Review. Energy Convers. Manag. 2014, 80, 202–228. [Google Scholar] [CrossRef]
  11. Yusri, I.M.; Mamat, R.; Najafi, G.; Razman, A.; Awad, O.I.; Azmi, W.H.; Ishak, W.F.W.; Shaiful, A.I.M. Alcohol Based Automotive Fuels from First Four Alcohol Family in Compression and Spark Ignition Engine: A Review on Engine Performance and Exhaust Emissions. Renew. Sustain. Energy Rev. 2017, 77, 169–181. [Google Scholar] [CrossRef] [Green Version]
  12. Awad, O.I.; Mamat, R.; Ali, O.M.; Sidik, N.A.C.; Yusaf, T.; Kadirgama, K.; Kettner, M. Alcohol and Ether as Alternative Fuels in Spark Ignition Engine: A Review. Renew. Sustain. Energy Rev. 2018, 82, 2586–2605. [Google Scholar] [CrossRef]
  13. No, S.-Y. Application of Liquid Biofuels to Internal Combustion Engines; Zentrum für Sonnenenergie-und Wasserstoff-Forschung Baden-Württemberg (ZSW), Ed.; Green Energy and Technology; Springer: Singapore, 2019; ISBN 978-981-13-6736-6. [Google Scholar]
  14. Sridhar, R.; Jeevahan, J.; Chandrasekaran, M. Effect of the Addition of 1-Pentanol on Engine Performance and Emission Characteristics of Diesel and Biodiesel Fuelled Single Cylinder Diesel Engine. Int. J. Ambient Energy 2020, 41, 58–63. [Google Scholar] [CrossRef]
  15. Shahir, S.A.; Masjuki, H.H.; Kalam, M.A.; Imran, A.; Ashraful, A.M. Performance and Emission Assessment of Diesel–Biodiesel–Ethanol/Bioethanol Blend as a Fuel in Diesel Engines: A Review. Renew. Sustain. Energy Rev. 2015, 48, 62–78. [Google Scholar] [CrossRef]
  16. Solubility of Things Alcohol Solubility-Methanol, Ethanol, Propanol Etc. in Solubility OF Things. Available online: https://www.solubilityofthings.com/water/alcohols (accessed on 2 February 2022).
  17. Cheung, C.S.; Zhu, L.; Huang, Z. Regulated and Unregulated Emissions from a Diesel Engine Fueled with Biodiesel and Biodiesel Blended with Methanol. Atmos. Environ. 2009, 43, 4865–4872. [Google Scholar] [CrossRef]
  18. Pellegrini, L.A.; Soave, G.; Gamba, S.; Langè, S. Economic Analysis of a Combined Energy–Methanol Production Plant. Appl. Energy 2011, 88, 4891–4897. [Google Scholar] [CrossRef]
  19. Bermúdez, J.M.; Fidalgo, B.; Arenillas, A.; Menéndez, J.A. CO2 Reforming of Coke Oven Gas over a Ni/ΓAl2O3 Catalyst to Produce Syngas for Methanol Synthesis. Fuel 2012, 94, 197–203. [Google Scholar] [CrossRef] [Green Version]
  20. Specht, M.; Bandi, A. “The Methanol Cycle”—Sustainable Supply of Liquid Fuels; Zentrum für Sonnenenergie-und Wasserstoff-Forschung Baden-Württemberg (ZSW): Stuttgart, Germany, 1999. [Google Scholar]
  21. Wang, C.; Chen, W.; Wang, W.; Wu, Y.; Chi, R.; Tang, Z. Experimental Study on Methanol Recovery through Flashing Vaporation in Continuous Production of Biodiesel via Supercritical Methanol. Energy Convers. Manag. 2011, 52, 1454–1458. [Google Scholar] [CrossRef]
  22. Shamsul, N.S.; Kamarudin, S.K.; Rahman, N.A.; Kofli, N.T. An Overview on the Production of Bio-Methanol as Potential Renewable Energy. Renew. Sustain. Energy Rev. 2014, 33, 578–588. [Google Scholar] [CrossRef]
  23. Leduc, S.; Lundgren, J.; Franklin, O.; Dotzauer, E. Location of a Biomass Based Methanol Production Plant: A Dynamic Problem in Northern Sweden. Appl. Energy 2010, 87, 68–75. [Google Scholar] [CrossRef]
  24. Renó, M.L.G.; Lora, E.E.S.; Palacio, J.C.E.; Venturini, O.J.; Buchgeister, J.; Almazan, O. A LCA (Life Cycle Assessment) of the Methanol Production from Sugarcane Bagasse. Energy 2011, 36, 3716–3726. [Google Scholar] [CrossRef]
  25. Holmgren, K.M.; Berntsson, T.; Andersson, E.; Rydberg, T. System Aspects of Biomass Gasification with Methanol Synthesis–Process Concepts and Energy Analysis. Energy 2012, 45, 817–828. [Google Scholar] [CrossRef]
  26. Clausen, L.R.; Houbak, N.; Elmegaard, B. Technoeconomic Analysis of a Methanol Plant Based on Gasification of Biomass and Electrolysis of Water. Energy 2010, 35, 2338–2347. [Google Scholar] [CrossRef] [Green Version]
  27. Hall, J. Potential Air Quality Benefits of Methanol as a Vehicle Fuel. Energy 1985, 10, 733–736. [Google Scholar] [CrossRef]
  28. Dürre, P. Biobutanol: An Attractive Biofuel. Biotechnol. J. 2007, 2, 1525–1534. [Google Scholar] [CrossRef] [PubMed]
  29. U.S. Department of Energy. Ethanol Fuel Basics. Available online: https://afdc.energy.gov/fuels/ethanol_fuel_basics.html (accessed on 5 February 2022).
  30. Renewable Fuels Association (RFA) The Impact of Accidental Ethanol Releases on the Environment. Available online: https://ethanolrfa.org/ethanol-101/environment (accessed on 15 March 2022).
  31. Liquid Transportation Fuels from Coal and Biomass; National Academies Press: Washington, DC, USA, 2009; ISBN 978-0-309-13712-6.
  32. García, V.; Päkkilä, J.; Ojamo, H.; Muurinen, E.; Keiski, R.L. Challenges in Biobutanol Production: How to Improve the Efficiency? Renew. Sustain. Energy Rev. 2011, 15, 964–980. [Google Scholar] [CrossRef]
  33. Lee, S.Y.; Park, J.H.; Jang, S.H.; Nielsen, L.K.; Kim, J.; Jung, K.S. Fermentative Butanol Production by Clostridia. Biotechnol. Bioeng. 2008, 101, 209–228. [Google Scholar] [CrossRef]
  34. Grana, R.; Frassoldati, A.; Faravelli, T.; Niemann, U.; Ranzi, E.; Seiser, R.; Cattolica, R.; Seshadri, K. An Experimental and Kinetic Modeling Study of Combustion of Isomers of Butanol. Combust. Flame 2010, 157, 2137–2154. [Google Scholar] [CrossRef]
  35. Niven, R.K. Ethanol in Gasoline: Environmental Impacts and Sustainability Review Article. Renew. Sustain. Energy Rev. 2005, 9, 535–555. [Google Scholar] [CrossRef]
  36. Singh, G.N.; Bharj, R.S. Study of Physical-Chemical Properties for 2nd Generation Ethanol-Blended Diesel Fuel in India. Sustain. Chem. Pharm. 2019, 12, 100130. [Google Scholar] [CrossRef]
  37. Babu, V.; Murthy, M.; Rao, A.P. Butanol and Pentanol: The Promising Biofuels for CI Engines—A Review. Renew. Sustain. Energy Rev. 2017, 78, 1068–1088. [Google Scholar] [CrossRef]
  38. Wu, H.; Nithyanandan, K.; Zhou, N.; Lee, T.H.; Lee, C.F.; Zhang, C. Impacts of Acetone on the Spray Combustion of Acetone–Butanol–Ethanol (ABE)-Diesel Blends under Low Ambient Temperature. Fuel 2015, 142, 109–116. [Google Scholar] [CrossRef]
  39. Chen, Z.; Wu, Z.; Liu, J.; Lee, C. Combustion and Emissions Characteristics of High N-Butanol/Diesel Ratio Blend in a Heavy-Duty Diesel Engine and EGR Impact. Energy Convers. Manag. 2014, 78, 787–795. [Google Scholar] [CrossRef]
  40. Tian, Z.; Zhen, X.; Wang, Y.; Liu, D.; Li, X. Combustion and Emission Characteristics of N-Butanol-Gasoline Blends in SI Direct Injection Gasoline Engine. Renew. Energy 2020, 146, 267–279. [Google Scholar] [CrossRef]
  41. Gao, K.; Boiano, S.; Marzocchella, A.; Rehmann, L. Cellulosic Butanol Production from Alkali-Pretreated Switchgrass (Panicum Virgatum) and Phragmites (Phragmites Australis). Bioresour. Technol. 2014, 174, 176–181. [Google Scholar] [CrossRef] [PubMed]
  42. Procentese, A.; Raganati, F.; Olivieri, G.; Elena Russo, M.; Marzocchella, A. Pre-Treatment and Enzymatic Hydrolysis of Lettuce Residues as Feedstock for Bio-Butanol Production. Biomass Bioenergy 2017, 96, 172–179. [Google Scholar] [CrossRef]
  43. Killol, A.; Reddy, N.; Paruvada, S.; Murugan, S. Experimental Studies of a Diesel Engine Run on Biodiesel N-Butanol Blends. Renew. Energy 2019, 135, 687–700. [Google Scholar] [CrossRef]
  44. Wei, H.; Feng, D.; Pan, M.; Pan, J.; Rao, X.; Gao, D. Experimental Investigation on the Knocking Combustion Characteristics of N-Butanol Gasoline Blends in a DISI Engine. Appl. Energy 2016, 175, 346–355. [Google Scholar] [CrossRef]
  45. Rajesh Kumar, B.; Saravanan, S. Effects of Iso-Butanol/Diesel and n-Pentanol/Diesel Blends on Performance and Emissions of a DI Diesel Engine under Premixed LTC (Low Temperature Combustion) Mode. Fuel 2016, 170, 49–59. [Google Scholar] [CrossRef]
  46. Zheng, J.; Tashiro, Y.; Wang, Q.; Sonomoto, K. Recent Advances to Improve Fermentative Butanol Production: Genetic Engineering and Fermentation Technology. J. Biosci. Bioeng. 2015, 119, 1–9. [Google Scholar] [CrossRef]
  47. No, S.-Y. Utilization of Pentanol as Biofuels in Compression Ignition Engines. Front. Mech. Eng. 2020, 6, 15. [Google Scholar] [CrossRef]
  48. Lee, Y.J. Utilisation of Bio-Ethanol as Automobile Fuels and Technology Development. Available online: https://www.konetic.or.kr/main/REPORT/REPORT_VIEW.asp?PARENT_NUM=1055&MENU1=4024 (accessed on 12 May 2022).
  49. Yun, Y. Alcohol Fuels: Current Status and Future Direction. In Alcohol Fuels-Current Technologies and Future Prospect; Yun, Y., Ed.; IntechOpen: London, UK, 2020; ISBN 978-1-78985-043-7. [Google Scholar]
  50. Renewable Fuels Association (RFA) Annual Ethanol Production: U.S. and World Ethanol Production. Available online: https://ethanolrfa.org/markets-and-statistics/annual-ethanol-production (accessed on 12 May 2022).
  51. Cascone, R. Biobutanol: A Replacement for Bioethanol? Chem. Eng. Prog. 2008, 104, 4. [Google Scholar]
  52. CropEnergies Bioethanol Production Processes. Available online: https://www.cropenergies.com/en/products/ethanol (accessed on 17 March 2022).
  53. Kandasamy, M.; Hamawand, I.; Bowtell, L.; Seneweera, S.; Chakrabarty, S.; Yusaf, T.; Shakoor, Z.; Algayyim, S.; Eberhard, F. Investigation of Ethanol Production Potential from Lignocellulosic Material without Enzymatic Hydrolysis Using the Ultrasound Technique. Energies 2017, 10, 62. [Google Scholar] [CrossRef]
  54. Gregg, J.; Bolwig, S.; Hansen, T.; Solér, O.; Ben Amer-Allam, S.; Pladevall Viladecans, J.; Klitkou, A.; Fevolden, A. Value Chain Structures That Define European Cellulosic Ethanol Production. Sustainability 2017, 9, 118. [Google Scholar] [CrossRef] [Green Version]
  55. Cotana, F.; Cavalaglio, G.; Pisello, A.; Gelosia, M.; Ingles, D.; Pompili, E. Sustainable Ethanol Production from Common Reed (Phragmites Australis) through Simultaneuos Saccharification and Fermentation. Sustainability 2015, 7, 12149–12163. [Google Scholar] [CrossRef] [Green Version]
  56. Eggert, H.; Greaker, M. Promoting Second Generation Biofuels: Does the First Generation Pave the Road? Energies 2014, 7, 4430–4445. [Google Scholar] [CrossRef]
  57. Zhang, X.; Fu, J.; Lin, G.; Jiang, D.; Yan, X. Switchgrass-Based Bioethanol Productivity and Potential Environmental Impact from Marginal Lands in China. Energies 2017, 10, 260. [Google Scholar] [CrossRef] [Green Version]
  58. Hansdah, D.; Murugan, S.; Das, L.M. Experimental Studies on a DI Diesel Engine Fueled with Bioethanol-Diesel Emulsions. Alex. Eng. J. 2013, 52, 267–276. [Google Scholar] [CrossRef] [Green Version]
  59. Chen, Q.; Jin, Y.; Zhang, G.; Fang, Y.; Xiao, Y.; Zhao, H. Improving Production of Bioethanol from Duckweed (Landoltia Punctata) by Pectinase Pretreatment. Energies 2012, 5, 3019–3032. [Google Scholar] [CrossRef] [Green Version]
  60. Rajendran, K.; Rajoli, S.; Taherzadeh, M. Techno-Economic Analysis of Integrating First and Second-Generation Ethanol Production Using Filamentous Fungi: An Industrial Case Study. Energies 2016, 9, 359. [Google Scholar] [CrossRef] [Green Version]
  61. Connolly, D.; Mathiesen, B.V.; Ridjan, I. A Comparison between Renewable Transport Fuels That Can Supplement or Replace Biofuels in a 100% Renewable Energy System. Energy 2014, 73, 110–125. [Google Scholar] [CrossRef]
  62. Tsolakis, A.; Bogarra, M.; Herreros, J. Environmental Impacts of Road Vehicles. In Environmental Impacts of Road Vehicles: Past, Present and Future; Harrison, R.M., Hester, R.E., Eds.; Issues in Environmental Science and Technology; Royal Society of Chemistry: Cambridge, UK, 2017; pp. 1–24. ISBN 978-1-78262-892-7. [Google Scholar]
  63. Lapuerta, M.; García-Contreras, R.; Agudelo, J.R. Lubricity of Ethanol-Biodiesel-Diesel Fuel Blends. Energy Fuels 2010, 24, 1374–1379. [Google Scholar] [CrossRef]
  64. Blumberg, P.N.; Bromberg, L.; Kang, H.; Tai, C. Simulation of High Efficiency Heavy Duty SI Engines Using Direct Injection of Alcohol for Knock Avoidance. SAE Int. J. Engines 2008, 1, 1186–1195. [Google Scholar] [CrossRef]
  65. Gong, C.; Deng, B.; Wang, S.; Su, Y.; Gao, Q.; Liu, X. Combustion of a Spark-Ignition Methanol Engine during Cold Start under Cycle-by-Cycle Control. Energy Fuels 2008, 22, 2981–2985. [Google Scholar] [CrossRef]
  66. Li, J.; Gong, C.-M.; Su, Y.; Dou, H.-L.; Liu, X.-J. Effect of Injection and Ignition Timings on Performance and Emissions from a Spark-Ignition Engine Fueled with Methanol. Fuel 2010, 89, 3919–3925. [Google Scholar] [CrossRef]
  67. Gong, C.-M.; Huang, K.; Jia, J.-L.; Su, Y.; Gao, Q.; Liu, X.-J. Regulated Emissions from a Direct-Injection Spark-Ignition Methanol Engine. Energy 2011, 36, 3379–3387. [Google Scholar] [CrossRef]
  68. Brusstar, M.J.; Gray, C.L. High Efficiency with Future Alcohol Fuels in a Stoichiometric Medium Duty Spark Ignition Engine. In Proceedings of the Powertrain & Fluid Systems Conference and Exhibition, San Diego, CA, USA, 29 October 2007. [Google Scholar]
  69. Rajesh Kumar, B.; Saravanan, S. Use of Higher Alcohol Biofuels in Diesel Engines: A Review. Renew. Sustain. Energy Rev. 2016, 60, 84–115. [Google Scholar] [CrossRef]
  70. Rajesh Kumar, B.; Muthukkumar, T.; Krishnamoorthy, V.; Saravanan, S. A Comparative Evaluation and Optimization of Performance and Emission Characteristics of a DI Diesel Engine Fueled with N-Propanol/Diesel, n-Butanol/Diesel and n-Pentanol/Diesel Blends Using Response Surface Methodology. RSC Adv. 2016, 6, 61869–61890. [Google Scholar] [CrossRef]
  71. Herreros, J.M.; Schroer, K.; Sukjit, E.; Tsolakis, A. Extending the Environmental Benefits of Ethanol–Diesel Blends through DGE Incorporation. Appl. Energy 2015, 146, 335–343. [Google Scholar] [CrossRef] [Green Version]
  72. Hergueta, C.; Tsolakis, A.; Herreros, J.M.; Bogarra, M.; Price, E.; Simmance, K.; York, A.P.E.; Thompsett, D. Impact of Bio-Alcohol Fuels Combustion on Particulate Matter Morphology from Efficient Gasoline Direct Injection Engines. Appl. Energy 2018, 230, 794–802. [Google Scholar] [CrossRef]
  73. Hellier, P.; Purton, S.; Ladommatos, N. Molecular Structure of Photosynthetic Microbial Biofuels for Improved Engine Combustion and Emissions Characteristics. Front. Bioeng. Biotechnol. 2015, 3. [Google Scholar] [CrossRef] [Green Version]
  74. Kumar, S.; Cho, J.H.; Park, J.; Moon, I. Advances in Diesel–Alcohol Blends and Their Effects on the Performance and Emissions of Diesel Engines. Renew. Sustain. Energy Rev. 2013, 22, 46–72. [Google Scholar] [CrossRef]
  75. Sukjit, E.; Herreros, J.M.; Dearn, K.D.; García-Contreras, R.; Tsolakis, A. The Effect of the Addition of Individual Methyl Esters on the Combustion and Emissions of Ethanol and Butanol -Diesel Blends. Energy 2012, 42, 364–374. [Google Scholar] [CrossRef]
  76. Sukjit, E. Synergistic Effects of Alcohol Based Renewable Fuels: Fuel Properties and Emissions. Ph.D. Thesis, University of Birmingham, Birmingham, UK, 2013. [Google Scholar]
  77. Hellier, P.; Ladommatos, N.; Yusaf, T. The Influence of Straight Vegetable Oil Fatty Acid Composition on Compression Ignition Combustion and Emissions. Fuel 2015, 143, 131–143. [Google Scholar] [CrossRef] [Green Version]
  78. Cann, A.F.; Liao, J.C. Pentanol Isomer Synthesis in Engineered Microorganisms. Appl. Microbiol. Biotechnol. 2010, 85, 893–899. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  79. Campos-Fernández, J.; Arnal, J.M.; Gómez, J.; Dorado, M.P. A Comparison of Performance of Higher Alcohols/Diesel Fuel Blends in a Diesel Engine. Appl. Energy 2012, 95, 267–275. [Google Scholar] [CrossRef]
  80. Imdadul, H.K.; Masjuki, H.H.; Kalam, M.A.; Zulkifli, N.W.M.; Alabdulkarem, A.; Rashed, M.M.; Teoh, Y.H.; How, H.G. Higher Alcohol–Biodiesel–Diesel Blends: An Approach for Improving the Performance, Emission, and Combustion of a Light-Duty Diesel Engine. Energy Convers. Manag. 2016, 111, 174–185. [Google Scholar] [CrossRef]
  81. Li, L.; Wang, J.; Wang, Z.; Xiao, J. Combustion and Emission Characteristics of Diesel Engine Fueled with Diesel/Biodiesel/Pentanol Fuel Blends. Fuel 2015, 156, 211–218. [Google Scholar] [CrossRef]
  82. Rajesh kumar, B.; Saravanan, S. Effect of Exhaust Gas Recirculation (EGR) on Performance and Emissions of a Constant Speed DI Diesel Engine Fueled with Pentanol/Diesel Blends. Fuel 2015, 160, 217–226. [Google Scholar] [CrossRef]
  83. Fayad, M.A.; Tsolakis, A.; Fernández-Rodríguez, D.; Herreros, J.M.; Martos, F.J.; Lapuerta, M. Manipulating Modern Diesel Engine Particulate Emission Characteristics through Butanol Fuel Blending and Fuel Injection Strategies for Efficient Diesel Oxidation Catalysts. Appl. Energy 2017, 190, 490–500. [Google Scholar] [CrossRef]
  84. Fayad, M.A.; Herreros, J.M.; Martos, F.J.; Tsolakis, A. Role of Alternative Fuels on Particulate Matter (PM) Characteristics and Influence of the Diesel Oxidation Catalyst. Environ. Sci. Technol. 2015, 49, 11967–11973. [Google Scholar] [CrossRef] [Green Version]
  85. Sukjit, E.; Herreros, J.M.; Piaszyk, J.; Dearn, K.D.; Tsolakis, A. Finding Synergies in Fuels Properties for the Design of Renewable Fuels–Hydroxylated Biodiesel Effects on Butanol-Diesel Blends. Environ. Sci. Technol. 2013, 47, 3535–3542. [Google Scholar] [CrossRef]
  86. Choi, B.; Jiang, X. Individual Hydrocarbons and Particulate Matter Emission from a Turbocharged CRDI Diesel Engine Fueled with n -Butanol/Diesel Blends. Fuel 2015, 154, 188–195. [Google Scholar] [CrossRef]
  87. Fayad, M.A.; Fernández-Rodríguez, D.; Herreros, J.M.; Lapuerta, M.; Tsolakis, A. Interactions between Aftertreatment Systems Architecture and Combustion of Oxygenated Fuels for Improved Low Temperature Catalysts Activity. Fuel 2018, 229, 189–197. [Google Scholar] [CrossRef] [Green Version]
  88. Campos-Fernandez, J.; Arnal, J.M.; Gomez, J.; Lacalle, N.; Dorado, M.P. Performance Tests of a Diesel Engine Fueled with Pentanol/Diesel Fuel Blends. Fuel 2013, 107, 866–872. [Google Scholar] [CrossRef]
  89. Xing-cai, L.; Jian-guang, Y.; Wu-gao, Z.; Zhen, H. Effect of Cetane Number Improver on Heat Release Rate and Emissions of High Speed Diesel Engine Fueled with Ethanol–Diesel Blend Fuel. Fuel 2004, 83, 2013–2020. [Google Scholar] [CrossRef]
  90. Lapuerta, M.; García-Contreras, R.; Campos-Fernández, J.; Dorado, M.P. Stability, Lubricity, Viscosity, and Cold-Flow Properties of Alcohol−Diesel Blends. Energy Fuels 2010, 24, 4497–4502. [Google Scholar] [CrossRef]
  91. Pinzi, S.; Redel-Macías, M.D.; Carmona-Cabello, M.; Cubero, A.; Herreros, J.M.; Dorado, M.P. Influence of 1-Butanol and 1-Pentanol Addition to Diesel Fuel on Exhaust and Noise Emissions under Stationary and Transient Conditions. Fuel 2021, 301, 121046. [Google Scholar] [CrossRef]
  92. Lapuerta, M.; Hernández, J.J.; Fernández-Rodríguez, D.; Cova-Bonillo, A. Autoignition of Blends of N-Butanol and Ethanol with Diesel or Biodiesel Fuels in a Constant-Volume Combustion Chamber. Energy 2017, 118, 613–621. [Google Scholar] [CrossRef] [Green Version]
  93. Sarathy, S.M.; Oßwald, P.; Hansen, N.; Kohse-Höinghaus, K. Alcohol Combustion Chemistry. Prog. Energy Combust. Sci. 2014, 44, 40–102. [Google Scholar] [CrossRef]
  94. No, S.-Y. Application of Biobutanol in Advanced CI Engines—A Review. Fuel 2016, 183, 641–658. [Google Scholar] [CrossRef]
  95. Lapuerta, M.; Sánchez-Valdepeñas, J.; Sukjit, E. Effect of Ambient Humidity and Hygroscopy on the Lubricity of Diesel Fuels. Wear 2014, 309, 200–207. [Google Scholar] [CrossRef]
  96. Lapuerta, M.; Rodríguez-Fernández, J.; Fernández-Rodríguez, D.; Patiño-Camino, R. Cold Flow and Filterability Properties of N-Butanol and Ethanol Blends with Diesel and Biodiesel Fuels. Fuel 2018, 224, 552–559. [Google Scholar] [CrossRef]
  97. Fernández-Rodríguez, D.; Lapuerta, M.; German, L. Progress in the Use of Biobutanol Blends in Diesel Engines. Energies 2021, 14, 3215. [Google Scholar] [CrossRef]
  98. Wichmann, H.-E. Diesel Exhaust Particles. Inhal. Toxicol. 2007, 19, 241–244. [Google Scholar] [CrossRef] [PubMed]
  99. Han, Y.; Cao, J.; Chow, J.C.; Watson, J.G.; An, Z.; Jin, Z.; Fung, K.; Liu, S. Evaluation of the Thermal/Optical Reflectance Method for Discrimination between Char- and Soot-EC. Chemosphere 2007, 69, 569–574. [Google Scholar] [CrossRef] [PubMed]
  100. Huang, J.; Wang, Y.; Li, S.; Roskilly, A.P.; Yu, H.; Li, H. Experimental Investigation on the Performance and Emissions of a Diesel Engine Fuelled with Ethanol–Diesel Blends. Appl. Therm. Eng. 2009, 29, 2484–2490. [Google Scholar] [CrossRef]
  101. Abou-Rachid, H.; Marrouni, K.E.; Kaliaguine, S. DFT Studies of the Hydrogen Abstraction from Primary Alcohols by O2 in Relation with Cetane Number Data. J. Mol. Struct. THEOCHEM 2003, 631, 241–250. [Google Scholar] [CrossRef]
  102. Al-Hasan, M.I.; Al-Momany, M. The effect of iso-butanol-diesel blends on engine performance. Transport 2008, 23, 306–310. [Google Scholar] [CrossRef] [Green Version]
  103. Choi, S.; Myung, C.L.; Park, S. Review on Characterization of Nano-Particle Emissions and PM Morphology from Internal Combustion Engines: Part 2. Int. J. Automot. Technol. 2014, 15, 219–227. [Google Scholar] [CrossRef]
  104. Huang, Z.; Lu, H.; Jiang, D.; Zeng, K.; Liu, B.; Zhang, J.; Wang, X. Combustion Behaviors of a Compression-Ignition Engine Fuelled with Diesel/Methanol Blends under Various Fuel Delivery Advance Angles. Bioresour. Technol. 2004, 95, 331–341. [Google Scholar] [CrossRef]
  105. Sayin, C.; Ilhan, M.; Canakci, M.; Gumus, M. Effect of Injection Timing on the Exhaust Emissions of a Diesel Engine Using Diesel–Methanol Blends. Renew. Energy 2009, 34, 1261–1269. [Google Scholar] [CrossRef]
  106. Chao, H.R.; Lin, T.C.; Chao, M.R.; Chang, F.H.; Huang, C.I.; Chen, C.B. Effect of Methanol-Containing Additive on the Emission of Carbonyl Compounds from a Heavy-Duty Diesel Engine. J. Hazard. Mater. 2000, 73, 39–54. [Google Scholar] [CrossRef] [PubMed]
  107. Zhang, Z.H.; Cheung, C.S.; Chan, T.L.; Yao, C.D. Experimental Investigation on Regulated and Unregulated Emissions of a Diesel/Methanol Compound Combustion Engine with and without Diesel Oxidation Catalyst. Sci. Total Environ. 2010, 408, 865–872. [Google Scholar] [CrossRef] [PubMed]
  108. Li, Y.; Jia, M.; Liu, Y.; Xie, M. Numerical Study on the Combustion and Emission Characteristics of a Methanol/Diesel Reactivity Controlled Compression Ignition (RCCI) Engine. Appl. Energy 2013, 106, 184–197. [Google Scholar] [CrossRef]
  109. Cheng, C.H.; Cheung, C.S.; Chan, T.L.; Lee, S.C.; Yao, C.D. Experimental Investigation on the Performance, Gaseous and Particulate Emissions of a Methanol Fumigated Diesel Engine. Sci. Total Environ. 2008, 389, 115–124. [Google Scholar] [CrossRef] [PubMed]
  110. Wang, Q.; Wei, L.; Pan, W.; Yao, C. Investigation of Operating Range in a Methanol Fumigated Diesel Engine. Fuel 2015, 140, 164–170. [Google Scholar] [CrossRef]
  111. Sayin, C.; Ozsezen, A.N.; Canakci, M. The Influence of Operating Parameters on the Performance and Emissions of a DI Diesel Engine Using Methanol-Blended-Diesel Fuel. Fuel 2010, 89, 1407–1414. [Google Scholar] [CrossRef]
  112. Zhang, Z.H.; Tsang, K.S.; Cheung, C.S.; Chan, T.L.; Yao, C.D. Effect of Fumigation Methanol and Ethanol on the Gaseous and Particulate Emissions of a Direct-Injection Diesel Engine. Atmos. Environ. 2011, 45, 2001–2008. [Google Scholar] [CrossRef]
  113. Geng, P.; Yao, C.; Wei, L.; Liu, J.; Wang, Q.; Pan, W.; Wang, J. Reduction of PM Emissions from a Heavy-Duty Diesel Engine with Diesel/Methanol Dual Fuel. Fuel 2014, 123, 1–11. [Google Scholar] [CrossRef]
  114. Wang, X.; Wang, Y.; Bai, Y. Oxidation Behaviors and Nanostructure of Particulate Matter Produced from a Diesel Engine Fueled with N-Pentanol and 2-Ethylhexyl Nitrate Additives. Fuel 2021, 288, 119844. [Google Scholar] [CrossRef]
  115. Najafi, G.; Yusaf, T.F. Experimental Investigation of Using Methanol–Diesel Blended Fuels in Diesel Engine. In Proceedings of the Proceedings of the Fourth International Conference on Thermal Engineering: Theory and Applications, Abu Dhabi, United Arab Emirates, 12–14 January 2009; pp. 1–5. [Google Scholar]
  116. Sayin, C. Engine Performance and Exhaust Gas Emissions of Methanol and Ethanol–Diesel Blends. Fuel 2010, 89, 3410–3415. [Google Scholar] [CrossRef]
  117. Zhang, Z.H.; Balasubramanian, R. Influence of Butanol–Diesel Blends on Particulate Emissions of a Non-Road Diesel Engine. Fuel 2014, 118, 130–136. [Google Scholar] [CrossRef]
  118. Wei, L.; Cheung, C.S.; Huang, Z. Effect of N-Pentanol Addition on the Combustion, Performance and Emission Characteristics of a Direct-Injection Diesel Engine. Energy 2014, 70, 172–180. [Google Scholar] [CrossRef]
  119. Wei, J.; Zeng, Y.; Pan, M.; Zhuang, Y.; Qiu, L.; Zhou, T.; Liu, Y. Morphology Analysis of Soot Particles from a Modern Diesel Engine Fueled with Different Types of Oxygenated Fuels. Fuel 2020, 267, 117248. [Google Scholar] [CrossRef]
  120. Li, M.; Wang, Z.; Li, L.; Chen, L.; Li, R. Particulate Status of Diesel Engine Fueled with Ethanol/Diesel Blends. Nongye Jixie Xuebao/Transactions Chinese Soc. Agric. Mach. 2013, 44, 28–32. [Google Scholar] [CrossRef]
  121. Ghadikolaei, M.A. Effect of Alcohol Blend and Fumigation on Regulated and Unregulated Emissions of IC Engines—A Review. Renew. Sustain. Energy Rev. 2016, 57, 1440–1495. [Google Scholar] [CrossRef]
  122. Wang, X.; Wang, Y.; Bai, Y.; Wang, P.; Zhao, Y. An Overview of Physical and Chemical Features of Diesel Exhaust Particles. J. Energy Inst. 2019, 92, 1864–1888. [Google Scholar] [CrossRef]
  123. Hua, Y.; Liu, F.; Wu, H.; Lee, C.F.; Li, Y. Effects of Alcohol Addition to Traditional Fuels on Soot Formation: A Review. Int. J. Engine Res. 2020, 22, 1395–1420. [Google Scholar] [CrossRef]
  124. Song, C.L.; Zhou, Y.C.; Huang, R.J.; Wang, Y.Q.; Huang, Q.F.; Lü, G.; Liu, K.M. Influence of Ethanol–Diesel Blended Fuels on Diesel Exhaust Emissions and Mutagenic and Genotoxic Activities of Particulate Extracts. J. Hazard. Mater. 2007, 149, 355–363. [Google Scholar] [CrossRef] [PubMed]
  125. BP Statistical Review of World Energy 2021. Available online: https://www.bp.com/content/dam/bp/business-sites/en/global/corporate/pdfs/energy-economics/statistical-review/bp-stats-review-2021-full-report.pdf (accessed on 5 July 2022).
  126. Bond, T.C.; Doherty, S.J.; Fahey, D.W.; Forster, P.M.; Berntsen, T.; Deangelo, B.J.; Flanner, M.G.; Ghan, S.; Kärcher, B.; Koch, D.; et al. Bounding the Role of Black Carbon in the Climate System: A Scientific Assessment. J. Geophys. Res. Atmos. 2013, 118, 5380–5552. [Google Scholar] [CrossRef]
  127. Zhang, Z.H.; Balasubramanian, R. Effects of Oxygenated Fuel Blends on Carbonaceous Particulate Composition and Particle Size Distributions from a Stationary Diesel Engine. Fuel 2015, 141, 1–8. [Google Scholar] [CrossRef]
  128. Winberry W., T.; Jungclaus, G. Compendium of Methods for the Determination of Toxic Organic Compounds in Ambient Air Second Edition Compendium Method TO-13A Determination of Polycyclic Aromatic Hydrocarbons (PAHs) in Ambient Air Using Gas Chromatography/Mass Spectrometry (GC/MS); U.S. Environmental Protection Agency: Cincinnati, OH, USA, 1999.
  129. Ravindra, K.; Sokhi, R.; Van Grieken, R. Atmospheric Polycyclic Aromatic Hydrocarbons: Source Attribution, Emission Factors and Regulation. Atmos. Environ. 2008, 42, 2895–2921. [Google Scholar] [CrossRef]
Figure 1. Raw materials used in the manufacturing of bioethanol in several nations (modified figure from Ref. [49]).
Figure 1. Raw materials used in the manufacturing of bioethanol in several nations (modified figure from Ref. [49]).
Energies 16 00597 g001
Figure 2. Biochemical conversion of biomass to alcohol fuels (modified from Ref. [30]).
Figure 2. Biochemical conversion of biomass to alcohol fuels (modified from Ref. [30]).
Energies 16 00597 g002
Figure 3. Bioethanol production process diagram [51].
Figure 3. Bioethanol production process diagram [51].
Energies 16 00597 g003
Figure 4. Biobutanol production process [53].
Figure 4. Biobutanol production process [53].
Energies 16 00597 g004
Figure 5. Bibliometric diagram ISI web of knowledge [97].
Figure 5. Bibliometric diagram ISI web of knowledge [97].
Energies 16 00597 g005
Figure 6. Effect of alcohol fuels on the PM mass, number, and size via several aspects, including fuel, PM physical, and PM chemical properties aspects.
Figure 6. Effect of alcohol fuels on the PM mass, number, and size via several aspects, including fuel, PM physical, and PM chemical properties aspects.
Energies 16 00597 g006
Table 1. Physical and chemical properties of neat tested fuels [90].
Table 1. Physical and chemical properties of neat tested fuels [90].
PropertyDieselMethanolEthanolButanolPentanol
Molecular formulaC12H22CH3OHC2H5OHC4H9OHC5H11OH
Molecular weight (kg·kmol−1)166.332.0446.774.11288.15
Cetane number52581718.2
Density at 15 °C (kg·m−3) 834.8791.3789.4811.5814.8
Kinematic viscosity at 40 °C (cSt)2.6270.581.132.172.74
Lower heating value (MJ·kg−1)45.9719.5826.8333.8134.65
Self-ignition temperature (°C)254–300463420345300
Vapour pressure (mmHg)0.41275576
Latent heat of evaporation (kJ/kg)270–3751162.64918.42581.4656
Carbon (wt%)86.4437.4852.1464.8668.13
Hydrogen (wt%)13.5612.4813.0213.5113.72
Oxygen (wt%)049.9334.7321.6218.15
Water content (mg·kg−1)41.7 11.4629.7
Boiling point (°C)180–36064.778.3117.4137.9
Table 2. Summary of ethanol and n-butanol properties [90].
Table 2. Summary of ethanol and n-butanol properties [90].
PropertyMethodsEthanol (C2H5OH)Butanol (C4H9OH)
Purity (% v/v) 99.799
Density at 15 °C (Kg/m3)EN ISO 3675789.4809.7
Kinematic viscosity at 40 °C (cSt)EN ISO 31041.132.22
Lower heating value (MJ/kg)UNE 5112326.8333.09
C (wt%) 52.1464.82
H (wt%) 13.1313.60
O (wt%) 34.7321.59
Water content (ppm wt)EN ISO 1293720241146
Molecular weight (kg/kmol) 46.0774.12
Boiling point (°C) 78.3117.5
Stoichiometric fuel/air ratio 1/9.011/11.19
Cold filter plugging point (°C)EN 116<−51<−51
Lubricity (WSD) (µm)EN ISO 12156-11057591
Cetane number 8 [100,101]17 [94,102]
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Tanwar, M.D.; Torres, F.A.; Alqahtani, A.M.; Tanwar, P.K.; Bhand, Y.; Doustdar, O. Promising Bioalcohols for Low-Emission Vehicles. Energies 2023, 16, 597. https://doi.org/10.3390/en16020597

AMA Style

Tanwar MD, Torres FA, Alqahtani AM, Tanwar PK, Bhand Y, Doustdar O. Promising Bioalcohols for Low-Emission Vehicles. Energies. 2023; 16(2):597. https://doi.org/10.3390/en16020597

Chicago/Turabian Style

Tanwar, Manju Dhakad, Felipe Andrade Torres, Ali Mubarak Alqahtani, Pankaj Kumar Tanwar, Yashas Bhand, and Omid Doustdar. 2023. "Promising Bioalcohols for Low-Emission Vehicles" Energies 16, no. 2: 597. https://doi.org/10.3390/en16020597

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop