13

Improving the use of liquid biofuels in internal combustion engines

R.J. Pearson and J.W.G. Turner,    University of Bath, UK

Abstract:

Liquid biofuels offer a range of attractive qualities, including the potential for increased energy independency and reduced greenhouse gas emissions. However, the market penetration and supply of biofuels may be constrained by a number of factors. In this chapter, the use of alcohols and biodiesel in internal combustion engines is discussed and the technology of flexible-fuel vehicles is described. It is shown how iso-stoichiometric ternary blends of gasoline, ethanol and methanol can serve as drop-in fuels for E85 flex-fuel vehicles. The chapter goes on to further explore ways in which biofuels can become part of a wider sustainable energy system based on carbon-neutral liquid fuels for the transport sector, enabling a gradual evolution of both vehicle and fuel technology.

Key words

alcohol; biodiesel; ethanol; flexible-fuel; fatty acid methyl ester (FAME); hydrated vegetable oil; hydrous ethanol; methanol; ternary blends

13.1 Introduction

Liquid biofuels provide an opportunity for nations to increase energy independency or reduce greenhouse gas emissions by supplying energy-dense fuels which are miscible with petroleum gasoline and diesel. Both bioethanol and fatty acid methyl ester (FAME)-based biodiesel can be used in low concentration blends in vehicles with no modifications. Only minor changes in fuel system material specifications together with a low-cost alcohol sensor are necessary for Vehicle Operating compatibility with high concentrations of ethanol. The low level alcohol blends, particularly in the form of 10 vol% blends of ethanol in gasoline (E10), are already displacing significant quantities of gasoline in countries such as the United States. Ethanol provides beneficial properties, including high resistance to autoignition, which can be exploited at high-load operating conditions in modern downsized pressure-charged spark-ignition engines. Its main drawback is its low volumetric energy density which, in the absence of significant changes to the fuel taxation system, is likely to limit its market penetration in the form of high concentration blends such as E85 due to increased fuel consumption under light-load operating conditions. Biodiesel penetration, in the form of FAME, is limited by the reluctance of manufacturers to warranty their vehicles for use with this fuel above very low concentration levels (5–7 vol%). This reluctance to warranty vehicles for operation on high levels of biodiesel could lead to gradual increases in the level of ethanol in gasoline or hydrated vegetable oil in diesel in order to meet legislation which effectively mandates the use of renewable fuels.

Ultimately the availability of sustainable feedstocks constrains the supply of biofuels and this limits the level at which they are able to displace fossil fuels. As a route to going beyond the biomass supply limit, the concept of Sustainable Organic Fuels for Transport (SOFT) has been developed, based on the synthesis of methanol from recycled CO2, water and renewable energy. In this way biofuels can be part of a wider energy system based on carbon-neutral liquid fuels for the transport sector, enabling a gradual evolution of both vehicle and fuel technology.

13.2 Competing fuels and energy carriers

The drive to reduce dependency on fossil fuels over recent years has focused attention on the use of alternative fuels for transport, particularly in the United States and Brazil, where the use of biofuels blended into fossil fuels has been adopted, in the latter case, since the mid-1970s. When they are made from feedstocks which satisfy appropriate sustainability criteria and do not give rise to the emission of significant levels of greenhouse gases in their cultivation, biofuels can help to alleviate concerns regarding energy security and climate change within the road transport sector which, globally, is over 90% dependent on oil (IEA, 2010). Through their miscibility with conventional gasoline and diesel fuels, bioethanol and biodiesel have been introduced into the fuel pool in significant quantities. The use of these alternative fuels has been possible without a quantum change in either the transport energy distribution infrastructure or the technology, and therefore cost, of the vehicles in which they are used. In the respect that they offer an evolutionary rather than a revolutionary transition, the adoption of liquid alternative fuels could serve as a more stable pathway than electrification or the use of molecular hydrogen in order to address issues of climate change security of energy supply for transport. Indeed, it is the opportunity for evolutionary transition which has led to the significant current presence of biofuels in the marketplace, as shown in Section 13.3.

13.2.1 On-board energy density and technology costs

Renewable liquid fuels provide, at low additional vehicle cost, the on-board energy storage levels required by vehicles for personal transport which are not dedicated solely to regular short-range routes. This is not the case for some other competing technologies. The very low net gravimetric and volumetric energy densities (including the mass/volume of the energy containment system) of current technology batteries are shown for lead-acid, nickel-metal hydride (NiMH) and lithium ion chemistries in Fig. 13.1. To match the range of a conventional gasoline-fuelled vehicle with a 50-litre fuel tank would require a useable battery capacity of approximately 100 kWh, accounting for the greater TTW efficiency of an electric vehicle. The mass of a fuel tank containing 50 litres of gasoline would be about 46 kg; that of a 100 kWh lithium ion battery would be in the range of 700–900 kg, depending on the technology and the permissible depth of discharge.

image
13.1 Net system volumetric and gravimetric energy densities for various on-board energy carriers (based on lower heating values).

Hydrogen is also fundamentally limited as an energy carrier in that it is the least dense element in the periodic table. Figure 13.1, which includes the system package volumes and masses, shows that, while the net on-board energy density of hydrogen comfortably exceeds that of current batteries, it is still very low compared with liquid fuels. Because of the extreme physical conditions required to package hydrogen, the bulky system volume and high storage tank mass become high fractions of the net volumetric and gravimetric energy levels of the vehicle on-board energy storage system (Amaseder & Krainz, 2006); Eberle, 2006). Details of high-pressure, cryogenic, and metal hydride hydrogen storage systems are discussed by Bossel (2006), Pearson et al. (2012a) and Pearson and Turner (2012).

Clearly, since alcohols are partially oxidized, they do not have the energy density of liquid hydrocarbon fuels but they are significantly better than current technology batteries and hydrogen storage systems, as shown for ethanol and methanol in Fig. 13.1. The fact that they are liquids means that vehicles can be fully re-fuelled in two or three minutes and the shape of the tanks containing them can be more easily adapted to available vehicle package space with simple vapour recovery systems and without the additional requirements for cooling systems required by electrochemical storage systems.

Key to enabling the widespread availability of sustainable transport is to provide solutions which customers can afford to purchase. Renewable liquid fuels satisfy this criterion by enabling the evolution of vehicles and fuel distribution infrastructures which are broadly compatible with current technologies. Figure 13.2 compares the vehicle bill of material costs for a variety of alternative fuel vehicles with a conventional vehicle powered by an internal combustion engine (ICE). A fixed ‘glider’ (vehicle rolling chassis, including the body) cost is assumed for all options (Pearson et al., 2012a). For the battery electric vehicle (BEV) options, a minimum state of charge (SoC) of 15% has been assumed; for the extended range electric vehicles (EREV) and proton exchange membrane fuel cell (PEM FC) options, a minimum SoC of 35% has been assumed. The battery cost assumed for volume production levels is $750/kWh (slightly better than current prices). An EREV can be thought of as a plug-in hybrid electric vehicle (PHEV) with a significant electric-only range and with the engine taking the role of an electrical generator operating largely independently of vehicle speed. For EREVs, the strategy is often to try to size the battery so that a large portion of the distance travelled by the vehicle can be done in EV mode.

image
13.2 Cost comparison of alternative energy vehicles. Assumed vehicle range (unless stated) = 550 km; battery cost = $750/kWh.

It is clear that, for a range-equivalent vehicle, the cost of the battery makes the BEV unaffordable to most customers. Reducing the vehicle range to 150 km from 550 km brings the costs down to a more accessible level, but this significantly range-compromised vehicle is still about 2.5 times more expensive than a conventional vehicle with a much higher utility level. This presents the customer with a very large negative price-performance differential. The EREV option, which enables lower storage capacity batteries to be used but requires both an electric motor and a fuel converter/generator (ICE assumed in this case), has a similar cost premium but is not encumbered by range compromise.

A significant portion of the high cost premium of a hydrogen-fuelled vehicle, whether using fuel cells or ICEs (the latter vehicle powertrain variant is not shown in Fig. 13.2), is the cost of the hydrogen storage system which, because the vehicle is also hybridized to manage the operating locus of the fuel cell in order to exploit its theoretically high efficiency levels, is additional to those of a 14 kWh battery (the same as that assumed for the EREV). It can be seen in Fig. 13.2 that even at the lowest fuel cell cost ($/kW), the bill of materials for hydrogen fuel cell electric vehicles is of such a level that widespread adoption is unlikely in the foreseeable future. This high vehicle price exacerbates the difficulty of justifying the expense of the necessary fuel-distribution infrastructure. Mintz et al. (2002) estimated the cost of providing a hydrogen infrastructure in the United States capable of re-fuelling 100 million fuel cell vehicles (40% of the US light duty vehicle fleet) as up to $650 × 109.

13.2.2 Environmental benefits

The debate regarding the environmental benefits of renewable fuels is complex and controversial. For a given feedstock and fuel, the specific production process, fertilizer used, transport and distribution, and, importantly, land-use change (direct and indirect) must be considered. In addition to well-to-wheel, or life cycle GHG emissions, parameters such as ‘carbon pay-back’ times have been calculated to quantify the time period over which a biofuel must be produced in order to offset the negative GHG impact of cultivating land which was formerly a carbon sink in the natural ecosystem. Whilst many studies have discussed these effects (Fargione et al., 2008; RFA, 2008; Searchinger et al., 2008; Bringezu et al., 2009; Zinoviev et al., 2010) they are not within the scope of the present work. However, the biofuels introduced into the EU as a result of the Renewable Energy Directive (EC, 2009a) are governed by sustainability criteria with a view to transport energy suppliers reducing the life cycle GHG emissions of their fuel by at least 6% by the end of 2020; there are also interim 2% and 4% targets to be met by the end of 2014 and 2017, respectively (EC, 2009c). Article 7b of the Fuel Quality Directive states that the GHG reductions for biofuels sold in the EU must be at least 35% (currently), rising to 50% in 2017, and 60% in 2018 for biofuels produced in installations in which production started after 1 January 2017 (EC, 2009c). Criteria for calculating these GHG benefits have been developed and default values for various fuels and pathways are defined in the Directive (EC, 2009c).

In addition to meeting supply targets discussed in Section 13.3.2, the US Renewable Fuel Standard (EPA, 2010a) also has requirements that fuels meet GHG emissions thresholds for compliance with each of four types of renewable fuel categories. California has its own initiative, the Low Carbon Fuel Standard (ARB, 2012), calling for a reduction in the carbon intensity of the transportation fuel pool used in the state of 10% by 2020. Figure 13.3 has been developed in order to indicate the well-to-wheel CO2 emissions of vehicles with a range of tank-to-wheel CO2 emissions as a function of the carbon intensity (g CO2/MJ) of the fuel being used. The vertical lines in figures (a) and (b) represent ethanol at E20 and E85 levels, respectively, which meets the EU GHG reduction targets for biofuels mentioned above. It can be seen that using low-carbon-intensity renewable fuels at high concentration levels enables low well-to-wheel CO2 emission levels to be achieved for a range of cars covering a range of operating efficiencies. These GHG emissions are similar to (E20), or substantially lower than (E85), the well-to-wheel emissions of an electric vehicle operating on electricity generated at the EU average carbon intensity (Pearson and Turner, 2012).

image
13.3 Well-to-wheel CO2 emissions (g CO2/km) as a function of fuel well-to-tank carbon intensity (g CO2/MJ) and vehicle tank-to-wheel CO2 emissions. (a) using E20; (b) using E85. Carbon intensity of default fossil fuel is that given in EC (2009c) of 83.8 g CO2/MJ. No TTW CO2 benefits due to the use of ethanol are assumed.

13.3 Market penetration of liquid biofuels

In the previous section it has been established that renewable liquid fuels are a potentially pragmatic route to de-carbonizing transport because they provide evolutionary transition mechanisms for both the vehicle technology and fuel distribution infrastructure. These fuels are already making a significant contribution to transport energy supply. Figure 13.4 shows that, globally, ethanol was the largest contributor to alternative road transport fuel in 2009 with consumption of 38.7 million tonnes of oil equivalent per annum (Mtoe/a) representing 29% of the alternative road transport energy supply but only 2.3% of the total global fuel consumption for this sector of 1701 Mtoe/a (IEA, 2010). Only natural gas is close to ethanol in terms of energy supply as an alternative road transport fuel. The 13 Mtoe/a of biodiesel production gives it almost 10% of alternative road transport energy supply, but represents only 0.7% of total road transport fuel consumption (IEA, 2010).

image
13.4 Alternative road transport fuels as a fraction of global total alternative fuel supply. Based on data in IEA (2010).

13.3.1 Biodiesel

The European Union (EU) dominated the production of biodiesel in 2009 with 54% of production, followed by the United States and Brazil with 13% and 9%, respectively (IEA, 2010). About 75% of EU biodiesel is made from rapeseed feedstock, with 13% coming from soybean, and 8% from palm oil (Haer, 2011). China and India produced 2% and 1% of the world’s biodiesel, respectively. The main market for biodiesel consumption is Europe, in particular Germany, which introduced the use of ‘pure’ biodiesel (B100) from rapeseed feedstock grown on fallow land in the 1990s. Tax relief on this fuel meant that initially the retail price was 5–15% higher than fossil diesel on an energy equivalent basis and some OEMs decreed that their existing light duty vehicles were compatible with B100 (Kramer and Anderson, 2012). These factors gave rise to 1,900 fuel stations offering biodiesel by 2006, with about 70% of sales being of B100. The imposition of taxes by the German government in 2008 in order to recoup lost revenue transformed the market to one where most biodiesel sales originated from blending with fossil diesel.

Simultaneously, vehicle manufacturers raised compatibility concerns regarding the use of B100 with modern common rail fuel injection systems and particulate filters and their regeneration strategies. Additionally, concerns regarding oil dilution and degradation, deposit formation, and materials compatibility limited the blend concentration of biodiesel for use in all vehicles in the EU to 7 vol% by volume (B7) (EC, 2010). Cooper (2011) has shown that the world cereal and grain production is far greater than that of vegetable oil. This ameliorates the threat of production of bioethanol on food prices relative to that of biodiesel. In this chapter the focus will henceforth be primarily on the production and use of alcohol fuels from biological and synthetic techniques, although some of the technology discussed regarding the onward synthesis of hydrocarbons has relevance to alternative diesel fuels.

13.3.2 Alcohol fuels

The potential of ethanol as a fuel in the internal combustion engine has been recognized for over 100 years (White, 1907) and its high octane index meant that it was initially used as a knock inhibitor in gasoline until tetraethyl lead came to dominate the market. The US produced 56% of the ethanol consumed by the road transport sector in 2009, with Brazil producing 33% and the EU 4% (IEA, 2010).

Brazil

Historically, Brazil has the world’s most mature market for bioethanol in road transport fuels. Although ethanol-gasoline blending was taking place in Brazil on a significant scale in the 1930s, it was the OPEC oil crisis of 1973 which prompted the large-scale introduction of ethanol made from sugar cane as part of a national alcohol programme (‘ProAlcool’) in 1975 (Soccol et al., 2005). The evolution of alcohol-fuelled vehicles in Brazil is well summarized by Kramer and Anderson (2012). The extent of the penetration of ethanol in the Brazilian fuel pool is such that it is not possible to purchase gasoline which does not contain bioethanol. The level of ethanol in gasoline (to form ‘gasohol’) is currently allowed to vary between 18 and 25 vol%, depending on the state of the global sugar market. Hydrous ethanol is also sold, consisting of at least 94.5 vol% ethanol, with the balance being a permitted mix of components consisting of water (mainly), hydrocarbons and other alcohols.

After initially blending ethanol with gasoline at around 20 vol%, dedicated E100 vehicles were introduced in 1979 and, by 1985, these vehicles represented 80% of light-duty vehicle production, assisted by ethanol prices which were sufficiently lower than the gasoline price to easily offset the volumetric energy density differential. Subsequent fluctuations in the global sugar market led to gasoline being cheaper than ethanol and the demand for dedicated E100 vehicles virtually disappeared by the mid-1990s. The introduction of flexible-fuel vehicles (FFVs), capable of using anything from gasohol to E100 (actually hydrous ethanol about E94 with 6 vol% water), around 2002 rejuvenated ethanol sales and enabled customers to exploit the price and tax advantages of E100 over gasoline whilst protecting them from the volatility of the sugar price by enabling operation on lower ethanol blends when desirable.

United States

In the US, the oil crisis of the 1970s and the drive to improve air quality in states such as California gave rise to an interest in methanol. Abundant availability of indigenous feedstocks (principally coal and natural gas) and low production costs drove the introduction of gaoline-M85 FFVs (the first ‘modern’ FFVs), capable of operating on any mixture ranging from conventional gasoline to a mixture of 85 vol% methanol/15% gasoline. A national M85 standard (ASTM D5797 – covering mixtures containing between 70 and 85% methanol in gasoline) was put in place. Political factors, combined with the drop in the oil price and a methanol shortage brought about by the sale of methanol stocks to make methyl-tert-butyl ether (MTBE) for use as an oxygenate in reformulated gasoline, required by the Clean Air Act Amendments (EPA, 2012), led to the reduction of interest in methanol in the late 1990s and the rise of bioethanol production. The use of reformulated gasoline is mandated in some areas of the US in order to reduce emissions of usnburned hydrocarbons from older vehicles and to help reduce smog (ground-level ozone) formation. MTBE had been used in gasoline at low levels since 1979 as an octane enhancer when tetra-ethyl lead was banned. MTBE itself was subsequently banned in many states and its role as an oxygenate component in gasoline has now been completely replaced by ethanol.

The role of ethanol as an oxygenate source in reformulated gasoline began to rise in the early 2000s when the use of MTBE as an oxygenate additive was phased out, as shown in Fig. 13.5(a). Ethanol consumption has risen strongly since 2005 due to the federal policies which encouraged its use, reaching over 7.3 billion gallons gasoline-equivalent (gge) in 2009 and representing over 90% of US alternative fuel energy demand in 2009. The rapid growth in consumption has been matched by the rapid growth in production, as shown in Fig. 13.6 which reveals that the US is now by far the largest producer of ethanol, its output having risen to 62% of the total of 22.4 billion gallons produced globally in 2011. Together, the US and Brazil accounted for 87% of global production in that year.

image
13.5 Alternative fuel consumption (normalized to gasoline-equivalent US gallons where 1 US gallon = 3.785 litres)in the US road transport sector (a) 2003-2009 and (b) 2003-2010/11. Based on data in Davis et al., (2011) and US DoE (2012a). (‘EtOH as ox’ and ‘MTBE as ox’ indicate the use of ethanol and MTBE as oxygenates in gasoline fuel.).
image
13.6 Evolution of ethanol production by country since 2007. Volume in gallons of ethanol. Based on data from (US DoE, 2012a).

Figure 13.5(b) focuses on the lower demand fuels in the US market; they are all dwarfed by the quantities of ethanol used as an oxygenate component (see Fig. 13.5(a)). The rise, decline, and recent huge jump in biodiesel consumption can be seen, together with the gradual fall of liquid petroleum gas (LPG) consumption. Natural gas consumption (compressed (C)NG and liquefied (L)NG in the figure) has shown a steady rise, and it might be postulated that the current low cost of natural gas resulting from the exploitation of shale gas reserves could lead to further significant increases. An alternative scenario is the conversion of natural gas to methanol for transport use (Turner et al., 2012a). It is also clear that electricity and hydrogen have no significant demand up to 2009, comprising 0.06% and less than 0.002% of US alternative fuel energy demand in 2009, respectively.

The original Renewable Fuels Standard (RFS) (EPA, 2007) was created under the Energy Policy Act of 2005 and required 7.5 billion US gallons of renewable fuel to be blended into gasoline by 2012. As a result of the Energy Independence and Security Act of 2007 (US Congress, 2007) the Renewable Fuel Standard was expanded (RFS2) (EPA, 2010a) to include diesel as well as gasoline, mandated an increase of renewable fuel from 9 billion gallons in 2008 to 36 billion gallons in 2022, and established new categories of renewable fuels, setting separate targets for volume and greenhouse gas (GHG) reduction for each fuel. Figure 13.7 shows the ramp up in fuel volumes required to meet the RSF2 stipulation. The target for ‘biomass-based diesel’ is a minimum of 1 billion gallons from 2012 to 2022 (with the exact target to be set by future rulemaking) whilst that for ‘cellulosic biofuel’ was set to rise from 0.5 billion gallons in 2012 to 16 billion gallons in 2022, most of which is expected to be in the form of cellulosic ethanol (EPA, 2010a). In total, 21 billion gallons of ‘advanced biofuels’ (including cellulosic biofuels) are required in 2022, leaving 15 billion gallons to be supplied by first generation fuels, mostly in the form of corn ethanol. Many problems have been encountered meeting the targets for cellulosic biofuels and the 2012 target has recently been revised downward to 8.65 million gallons – this will represent less than 0.006% of US fuel usage that year (more precisely, the figure of 0.006% represents cellulosic biofuels as a fraction of non-renewable gasoline and diesel use), compared with 9.23% overall for renewable fuels, comprised mostly of ‘corn ethanol’ (EPA, 2011a).

image
13.7 Renewable fuel volume requirements for RFS2 Note: any fuel which meets the requirement for cellulosic biofuel or biomass-based diesel is also valid for meeting the advanced biofuel requirement. (based on data in EPA, 2010a).

Anderson et al. (2012a) provide an excellent account of the introduction of ethanol into the gasoline fuel pool in the US, including details of the changes in the gasoline blend stock, namely the blend stock for oxygenate blending (BOB), into which the ethanol is blended. Between 2000 and 2010, US ethanol consumption grew from 1.6 billion gallons/year to 13 billion gallons/year; the latter figure represents a hypothetical nationwide uniform ethanol-gasoline blend level of almost 10 vol% (Anderson et al., 2012a). In fact, virtually all the ethanol used in US transport is used in the form of E10 (a mixture of 10 vol% of ethanol in ‘gasoline’) which has been available in the US since the 1980s but is now widespread due to the political and environmental developments described above.

A ‘blend wall’ has arisen due to all the low-level ethanol-gasoline blends, often referred to as ‘gasohol’, being close to the maximum ethanol content (10 vol%) which vehicle manufacturers will allow for a user to remain within their warranty provision. In order to address this issue, the EPA has granted two partial waivers (EPA, 2010b, 2011b) which allow, but do not mandate, the introduction of E15 for use in light-duty vehicles of model year 2001 and later. The approval process is inherently slow since it involves extensive testing and is open to challenge by vehicle manufacturers.

Anderson et al. (2012) examine various scenarios for ethanol introduction, including the most optimistic, where the RFS2 targets are met. This latter scenario would lead to notional uniform ethanol blend levels of E24 by 2022 and of E29 by 2035 (assuming a 2% annual growth post-2022). In the absence of significant growth in E85 sales, this could not be implemented with new waiver approvals since these levels of ethanol are well beyond the tolerance capability level of current conventional vehicles. New vehicle and engine specifications would be required, together with the maintenance of a protection grade fuel (E10 or E15) for existing vehicles. A discussion of octane targets for these blends is also included.

Flexible-fuel (or flex-fuel) vehicles (FFVs) are capable of using ethanol in concentration levels of up to 85 vol% (E85). However, despite the registration of over 9 million such cars at the end of 2011, incentivized by the rating of FFVs under CAFE legislation (EPA, 2010c), representing 4% of the US LDV fleet (Anderson et al., 2012a), only 1% of the total ethanol use has been in the form of E85 sales, as shown in Fig. 13.5(a) and (b). This has led to a reappraisal of the Alternative Motor Fuels Act (AMFA) credits given to FFVs in future CAFE and EPA fuel economy and emissions legislation (EPA, 2011c). The fuel economy rating of an FFV is calculated by taking the harmonic mean of the fuel economy on gasoline (or diesel) and the fuel economy of the alternative fuel divided by 0.15. Thus, a vehicle which achieves 25 miles/US gallon using gasoline and 15 miles/US gallon using E85 would be rated having a fuel economy of (2/((1/25) + (0.15/15)) = 40 miles/US gallon. This assumes that the FFV operates on the alternative fuel 50% of the time. US Congress has extended the FFV incentive (called the ‘dual-fuelled’ vehicle incentive) to model year (MY) 2019 but has provided for its phase-out between MY 2015 and MY 2019 by gradually reducing the allowed limit of the maximum fleet fuel economy increase for a manufacturer due to this credit from 1.2 miles/gallon. (MYs 1993–2014) to 0 miles/gallon after MY 2019 in 0.2 miles/gallon decrements over that period (EPA, 2011c). FFV credits for MY 2020 and beyond will reflect the ‘real-world’ percentage of usage of the alternative fuels (EPA, 2011c).

Ethanol in high-concentration form, E85 (now defined by ASTM D5798 as 51–83 vol% ethanol in gasoline), has suffered both from limited availability and uncompetitive pricing on an energy basis in the US. There are fewer E85 pumps in the US than there are EV charging stations (about 2,500 versus about 6,750 in 2012). The requirement for FFVs to run on any concentration of ethanol in gasoline, from 0% to 85%, also means that the vehicles are generally not configured to be capable of exploiting the high octane numbers of the higher level ethanol blends – this prevents such vehicles being able to offset the reduction in volumetric energy content of the fuel by increasing the thermal efficiency of the engine. Conversely, if FFVs were to be optimized to have high compression ratios in order to reduce their volumetric fuel consumption on E85 (as discussed in Section 13.4.2), they would give an unattractive increase in fuel consumption when operating on fuel with lower ethanol concentration such as E10 if the octane level of the BOB is maintained (Anderson et al., 2012a). Although this may be viewed as a mechanism to incentivize the FFV customer to use E85, the proposition is only reasonable if there are sufficient fuel stations offering the fuel.

European Union

Whilst the European gasoline specification EN228 allows up to 3% methanol in gasoline, there has never been any specification for high concentration of methanol in Europe analogous to the ASTM 5797 standard. In the EU, by the end of 2020, the Renewable Energy Directive (RED) and the Fuel Quality Directive (FQD) (EC, 2009a, 2009c) together require that 10% of transport energy be supplied in renewable form and that the overall GHG intensity of fuels should be reduced by 6%. With diesel penetration at approximately 50% across the EU, it is possible that, due to lack of sufficient supplies of sustainable vegetable oils for biodiesel manufacture and some issues of achieving emission compliance of modern vehicles using more than 7 vol% of biodiesel, the RED and FQD targets may need to be met by supplying base fuel in the form of E20, when the lower volumetric energy density of ethanol is considered (Cooper, 2011). However, recent attempts to introduce E10 into the German market did not go well due to some customer confusion (Kramer and Anderson, 2012).

In contrast to the CAFE regulations which make it attractive to manufacture FFVs in the US, the fiscal penalties for GHG emissions from cars sold in the EU is based only on tailpipe CO2 emissions. This gives vehicle manufacturers no incentive to spend even the small extra amount required in order to produce an FFV (ca. €100/vehicle). It has recently been suggested by a major transport energy supplier (Cooper, 2011) that attributing some CO2 benefit to manufacturers will provide a more compelling reason for OEMs to make FFVs and thus produce a greater outlet for ethanol as an automotive fuel.

Despite the lack of apparent incentive, manufacturers such as Saab, Ford, Volvo, Renault, and VW have introduced FFVs into their vehicle range. The EU vehicle tailpipe CO2 penalty system does, however, presently allow a 5% reduction in tailpipe CO2 to be claimed for any FFV that an OEM sells, provided one-third of the fuel stations in the country in which the vehicle is sold has at least one E85 refuelling pump (EC, 2009b). In Sweden there has been co-ordinated activity to install E85 pumps so that by 2009 50% of the network was covered (Bergström et al., 2007a), rising to 59% in 2011 (Kramer and Anderson, 2012), and in 2008, 22% of all new car sales were FFVs (Kramer and Anderson, 2012), driven by government fuel tax relief, which made E85 cheaper than gasoline on an equal energy basis, and vehicle use initiatives. Kramer and Anderson (2012) show that this fuel tax benefit has been variable since 2005. They also show that the drop in FFV sales which occurred in 2009 coincided with a period when E85 had a cost disadvantage of 30% relative to gasoline.

The benefits of liquid fuels compared with their gaseous counterparts are again highlighted by the 2006 Swedish legislation requiring that all fuel stations above a certain size offer at least one alternative fuel: most stations covered by the law installed E85 pumps and storage tanks which, at €40,000–45,000, offered a ten-fold lower installation cost compared with biogas storage and dispensing equipment (Kramer and Anderson, 2012).

In the rest of Europe, legislation and incentives for high concentration ethanol use are largely absent and thus fuel pump availability for E85 and FFV sales remain sparse. For Germany in 2010 FFVs represented 0.05% of the 2.9 million new vehicle sales (Kramer and Anderson, 2012). It is, however, perhaps worth noting that for a vehicle at the 2011 EU average of 135.7 g CO2/km, and at the highest proposed fine rate in 2015 of €95/(gCO2/km), this represents a saving to the manufacturer of €541 per car (with the benefit limited to 5.7 g CO2/km in this instance since the target would be achieved), which the authors contend is significantly greater than the additional costs of producing a vehicle which is flex-fuel capable E85 (Turner et al., 2012a).

China

Ethanol blends in gasoline have been used in five Chinese provinces since 2004; however, the use of methanol is favoured in order to avoid conflicts with food demand. Whilst China produced only 3% of the ethanol used in road transport globally in 2009 (US DoE, 2012a), it has dominated the production and use of methanol in this sector. The consumption of methanol in China, mainly in the form of M15, is around 3 million tonnes per annum (Niu and Shi, 2011) and is motivated by China’s large coal reserve which offers the potential of greater energy independence. Processes to convert coal to ethanol are also being investigated (Pang, 2011).

13.4 Use of liquid biofuels in internal combustion engines

13.4.1 Biodiesel

Whilst alcohol fuels present the fuel blender, additive supplier, and vehicle manufacturer with a tightly defined blend component having consistent properties, the chemical composition of biodiesel formed by transesterification of seed oils or animal fats to form fatty acid methyl esters (FAMEs) is dependent on the original feedstock source and the esterification process. This results in a wide variation in FAME compostition.

The cetane number, which measures the auto-ignitability of fuels for compression-ignition engines, varies over wide ranges for FAME components from the same feedstock origin. The cetane number of FAME is dependent on the distribution of fatty acids in the original oil or fat from which it was produced. Higher cetane numbers are caused by higher saturation levels in the fatty acid molecules and longer carbon chains (Geller and Goodrum, 2004). Bamgboyne and Hansen (2008) report cetane numbers for biodiesel fuels from a wide range of feedstocks. The cetane number of biodiesel derived from soyabean oil (soyabean methyl ester, or SME) has been found to vary between 45 (lower auto-ignitability) and 60 (higher auto-ignitability), whilst that of rapeseed methyl ester (RME) can vary between 48 and 61.2. Palm oil methyl ester (POME) has been measured to have cetane numbers between 59 and 70. This compares with a typical cetane number of a premium European diesel of around 60. The minimum cetane number required by conventional diesel fuel specifications in the EU (EN 590) and the US (ASTM D975) is 51 and 40, respectively.

The fact that FAME molecules are esters and are therefore oxygenated means that they will have a lower gravimetric energy density than petroleum diesel fuels. However, the large size of the molecules (in the range C12 to C22) means that the impact of the two oxygen atoms which comprise the ester functional group is much lower than that of the oxygen atom contained within the methanol (C1) or ethanol (C2) molecules. The degree of this energetic deficit is also ameliorated by the fact that the density of FAME is higher than that of petroleum diesel due to increased chain length and the presence of carbon–carbon double bonds. Thus, whilst the gravimetric energy density of FAME can be perhaps 12 per cent lower than a mineral diesel such as US No. 2 diesel fuel, its increased density can reduce the difference in volumetric energy content to about 7%. In B10 blends the impact on fuel economy of FAME has been found to be less than 1% (Gardiner et al., 2011).

The wide variations in the FAME composition and its consequent variable interaction with the base diesel in a blend can have markedly different effects on low temperature vehicle operability, with the fuel pour point and cold filter plugging point changing significantly (Saito et al., 2008). The fuel’s oxidation stability (McCormick et al., 2006; Miyata et al., 2004), its compatibility with the vehicle fuel injection equipment, its propensity to form deposits (Caprotti et al., 2007), and the effects of fuel dilution on the engine lubricant (Thornton, 2009) also vary significantly with the composition of the FAME. Bespoke additives are required for specific blend compositions, making the task of ensuring fuel compliance with the vehicle fleet a complex task. The issues are well summarized by Richards et al. (2007). In contrast, hydrogenated vegetable oils (HVOs) and biomass to liquid (BTL) fuels have compositions which are much closer to their petroleum diesel counterparts; however, both types of fuel are significantly more energetically intensive to produce than FAME-based fuels and therefore more expensive. The ability of FAME-based biodiesel to lubricate fuel pumps and fuel injectors (its ‘lubricity’) is superior to that of petroleum diesel; however, the lubricity of HVO is lower, requiring lubricity additives.

As with all fuels, adherence to rigorous quality standards is necessary for increased penetration of biodiesel into the market. Standards exist in the EU and US for biodiesel quality. The EU standard, EN 14214, controls the quality of FAME used either as a fuel itself or as a blending component in diesel fuel. This biodiesel standard specifies the minimum ester content (96.5 mass%) and maximum methanol (used in the production process), glyceride, and glycerol content. Density and viscosity ranges are specified, and the minimum cetane number, set at 51, is identical to that of petroleum diesel fuel (a discussion of the relevance of the specification parameters is given by Ferrari et al., 2011).

The ASTM D975 standard for conventional diesel fuel allows biodiesel concentration of up to 5 vol%. Such blends are approved for safe operation in any compression-ignition engine designed to be operated on petroleum diesel (US DOE, 2012b). Blends of up to 20% biodiesel in 80% petroleum diesel (B20), controlled by ASTM D7467 (B6–B20), are the largest outlet for biodiesel in the United States and provide a good compromise in balancing cost, emissions, cold-weather performance, and materials compatibility. Operation on B20 and lower-level blends does not in principle require engine modification. However, not all diesel engine manufacturers warrant their products for use with such blends. ASTM D6751 regulates the specification of B100 for use as both a fuel and a blending component in petroleum diesel.

Whilst there can be significant emissions benefits from the use of biodiesel, as discussed by Gardiner et al. (2011), it seems that, in the form of FAME, there is an industry-wide desire to limit its use to relatively low-level blends. The quality of conventional diesel fuel in the EU is controlled by the EN 590 standard which currently limits FAME content to 5 vol% (B5). An extension of the FAME limit to 7% (B7) is proposed, with no limit to levels of HVO or ‘diesel-like hydrocarbons made from biomass using the Fischer–Tropsch process’ (EC, 2009c). The varying properties of FAME as a function of its feedstock gives rise to conservatism on the part of the vehicle manufacturers to warrant their vehicles for use with higher blends and may be an obstacle to achieving compliance with the RED and FQD, requiring either additional bio-content to come from HVO or BTL fuels or from a disproportionate increase in the use of ethanol in gasoline blends.

13.4.2 Alcohol fuels for spark-ignition engines

The principal alcohols which have been used or are being considered for use as either a component in fuel blends with gasoline or as the dominant fuel blend constituent themselves are methanol, ethanol and butanol. Methanol (CH3OH) and ethanol (C2H5OH) have one and two carbon atoms respectively, and no isomers, whilst butanol (C4H9OH) has four carbon atoms and four structural isomers. The higher-order alcohols, that is, those with more than two carbon atoms (primarily propanol, butanol and pentanol (amyl alcohol)), formed in the fermentation process are sometimes known as fusel alcohols.

As the number of carbon atoms increases, the influence of the functional hydroxyl (OH) group on the physico-chemical properties of the molecule diminishes. The presence of the OH group in place of one of the hydrogen atoms of an alkane induces significant polarity in the molecule due to the two lone pairs of electrons present on the oxygen atom (Pearson and Turner, 2012). The concentration of negative charge around the oxygen atom produces a net positive charge on the rest of the alcohol molecule, which, in particular, is focused around the hydrogen atom attached to the oxygen atom of the hydroxyl group. This intra-molecular polarity generates strong inter-molecular forces, known as hydrogen bonds. These forces give the low-carbon-number alcohols higher boiling points and enthalpies of vaporization than would be expected for non-polar compounds of similar molecular mass. As in FAME-based biodiesel, however, the presence of the oxygen atom reduces the energy density of the fuel. Its impact is more pronounced the smaller the molecule, so that methanol has less than half the gravimetric energy density of gasoline (generally a mixture of C4 to C8 hydrocarbons) and about 40% of that of methane, the corresponding C1 alkane. More detail on the properties of alcohols as fuels can be found in Pearson and Turner (2012).

Volumetric energy density and stoichiometry

The impact of alcohol concentration on the volumetric energy density of gasoline blends with ethanol, methanol, and butanol is shown in Fig. 13.8. For alcohol concentration of 10 vol% in gasoline, the reduction in volumetric energy density relative to the gasoline is 1.5%, 3.3% and 5%, for butanol, ethanol and methanol, respectively, whereas for 85% alcohol concentration the respective reductions in energy density are 12.8%, 28% and 42.5% for these alcohols. With vehicles which use engines not optimized for use with alcohol fuels (which is the current situation, even with FFVs), the consequence of this reduced volumetric energy content is an approximately proportional increase in volumetric fuel consumption for operating cycles which are not sufficiently aggressive to exploit the higher octane numbers of the resulting fuel blends. A consequence of this is that it is difficult to achieve significant market penetration of fuels with high alcohol concentrations with a volumetric-based fuel taxation system, rather than one based on the energetic content of the fuel, or the non-renewable carbon component (Turner and Pearson, 2008).

image
13.8 Volumetric energy density variation of alcohol-gasoline blends.

As a corollary to the presence of the oxygen atom in alcohol fuels having an impact on reducing the volumetric energy content which reduces as the length of the carbon chain increases, Fig. 13.9 shows how the stoichiometric air–fuel ratios of gasoline-butanol mixtures are closer to that of gasoline than those of gasoline-ethanol and gasoline-methanol mixtures. This is one of the reasons that gasoline standards have oxygenate limits. The impact of the alcohol concentration on the oxygen mass fraction is shown in Fig. 13.9. The European EN228 standard for gasoline currently limits the mass fraction of oxygenates in the fuel to 2.7% – this limit corresponds to about 7 vol% of ethanol in the blend (E7) and 11 vol% of butanol (Bu16). Methanol is limited as a specific component to 3 vol%. The increase in this level to 3.7% specified in the Fuel Quality Directive (EC, 2009c) allows E10 and Bu15 blends. Methanol content is again limited to 3 vol%. For comparison, a 3.7% oxygenate limit allows approximately 22 vol% MTBE or ETBE in gasoline – these fuels can be made from methanol and ethanol, respectively, with a process energy overhead.

image
13.9 Stoichiometric air–fuel ratio of blends of alcohol-gasoline blends.

Vapour pressure

Vapour pressure is an important property of automotive fuels, influencing the start quality of an engine in cold ambient temperatures (if the vapour pressure is too low), and affecting the evaporative emissions from the vehicle (adversely if the vapour pressure is too high). Gasoline specifications stipulate allowable vapour pressures dependent on the season, geographical location and alcohol content – these are extensively reviewed by Andersen et al. (2010). Figure 13.10 shows how vapour pressures (calculated as Reid vapour pressure, or RVP) of methanol, ethanol and iso-butanol vary as a function of the volumetric concentration of the alcohol in a blend with a reference gasoline fuel (RF-02-03 (Turner et al., 2012b; Pearson et al., 2012b)). Both ethanol and methanol cause a peak in measured RVP at concentration levels between 5 and 10 vol%, increasing the value relative to the base gasoline by about 8 kPa and 23 kPa, respectively. This requires an adjustment in the vapour pressure of the BOB in order for the blend to remain below normal specified limits. The EU Fuel Quality Directive (EC, 2009c) allows a derogation from the maximum summer vapour pressure for low-level ethanol blends.

image
13.10 Calculated variation of vapour pressure of methanol, ethanol and isobutanol with alcohol volume fraction. For validation data, see Turner et al. (2012b) and Pearson et al. (2012b). The base gasoline has an RVP of 65 kPa. Courtesy of M. Davy, Loughborough University.

Figure 13.10 shows that, at high concentrations, the vapour pressure of the methanol and ethanol blends reduces below that of the gasoline blend stock: this occurs at about 80 and 45 vol% for methanol and ethanol, respectively, with the gasoline used in this case. At high concentrations this can cause low temperature cold-start difficulties in unmodified engines, but this issue is easily overcome in FFVs, particularly with direct fuel injection technology where injection into the hot compressed air is found close to top-dead-centre. Siewart and Groff (1987), Kapus et al. (2007), Marriott et al. (2008) and Hadler et al. (2011) discuss the successful use of fuel injection strategies to augment the quality of the engine start. For port-fuel-injected engines, measures such as heating the fuel rail can enable acceptable cold-start performance down to − 25 °C. Bergström et al. (2007a) report acceptable cold starts down to − 25 °C in the absence of additional technology on an engine with port-injection (PI) of fuel using a Swedish winter grade bioethanol (E75 with a Reid vapour pressure of 50). In the US, the ASTM D5798 standard for E85 can allow the ethanol concentration to be as low as 51% by volume ethanol – this provides a significantly easier blend for low-temperature starting.

Octane numbers

Table 13.1 shows the low-carbon number alcohols have a number of physico-chemical characteristics which are synergistic with modern engine designs, and in particular highly pressure-charge downsized engines. In particular, the high research octane numbers (RONs) of methanol, ethanol and iso-butanol are facilitators for improved combustion phasing and lower component protection over-fuelling (to control exhaust gas temperatures). The motor octane numbers (MONs) of these fuels are proportionally less high and thus their sensitivity (RON-MON) is very high relative to gasoline; nevertheless, these fuels would not be considered as having a low MON relative to most gasoline sold throughout the world.

Table 13.1

Properties of 95 RON gasoline, methanol, ethanol, and iso-butanol

Fuel QLHV [MJ/(kg fuel)] hfg [kJ/(kg fuel)] Density at NTP [kg/l] RON/MON Boilng point at 1 bar [deg. C] Stoich. AFR (mass)[-] Stoich. AFR(mole)[-] QLHV [MJ/ (kg stoich. air)] hfg [kJ/(kg stoich. air)] CO2emission[g/MJ]
Gasoline 95 43.2 350 0.74 95.3/85.9 37–167 14.31 47.12 3.02 24.5 72.85
Methanol 20.09 1100 0.79 108.7/88.6 65 6.44 7.14 3.12 170.9 68.44
Ethanol 26.95 925 0.79 108.6/89.7 79 8.96 14.29 3.01 103.3 70.98
Iso-butanol 33.08 585 0.81 106.3/90.4 118 11.14 28.57 2.97 52.5 71.90

Image

Note that the properties of ’95 RON gasoline’ vary considerably depending on local refinery streams and other blend components.

Moran and Taylor (1995) posited that, because the intake manifold temperature is not controlled during the RON test, the high heat of vaporization of the low-carbon number alcohols would cause a physical cooling effect which would have a direct bearing on the knock-limited compression ratio. In contrast, the MON test uses a heater to try to control the intake manifold temperature to 149 °C, which is well above the respective boiling points of these alcohols and is thus not affected by this physical cooling effect.

Stein et al. (2012) showed that both denatured ethanol (97% ethanol, 1% water, 2% gasoline) and hydrous ethanol (94.2% ethanol, 5.8% water) are extremely knock resistant, allowing very high engine loads to be achieved at low speeds in an engine with a 14.0:1 compression ratio.

Performance

Nakata et al. (2006) used a high compression ratio (13:1) naturally aspirated port-fuel injected spark-ignition engine and found that torque increased by 5% and 20% using E100 compared with the operation on 100 RON and 92 RON gasoline, respectively. The full improvements in torque due to being able to run MBT ignition timing were apparent for E50.

Marriott et al. (2008) show significant performance benefits for E85 (RON measured as 107.7) compared with a 104 RON gasoline when used in a naturally aspirated direct-injection gasoline engine. Peak torque generated by the engine increased by 5% and peak power by about 4% at the same enriched air–fuel ratio. An increase in volumetric efficiency of about 3% was measured at the peak torque operating point. Smaller but still significant performance benefits were available from operation at stoichiometric conditions when using E85 fuel. The majority of the combustion-related benefit in performance using E85 was determined to come from a reduction in the cumulative heat energy rejected to the engine coolant.

Korte et al. (2011) and OudeNijeweme et al. (2011) show significant synergies between the high octane properties of alcohol fuels and pressure-charged down-sized engines. Alcohol blends were made up using a 95 RON blend stock so that the E10 blend had a RON of 99 and the 16 vol% iso-butanol blend (Bu16) had a RON of 98. Note that this value is higher than that of a typical commercial E10 blend, even within Europe, due to the use of a 95 RON gasoline blend stock. Octane numbers of alcohol fuels with various blend stocks are discussed by Anderson et al. (2012a, 2012b).

Other blends used were E22, which matched the 102 RON of the high-octane forecourt specification gasoline included in the study, Bu68 (RON 104) and E85 (RON 106). Korte et al. (2011) conclude that all the high octane fuels used show a synergy with down-sizing under high-load conditions and that such fuels enable either higher levels of downsizing or increased compression ratios to be used, leading to additional CO2 improvements. In related work using substantially the same blends, Stansfield et al. (2012) showed that it is possible to use high-alcohol blends in an unmodified production vehicle which has not been sold as having flex-fuel capability. Here E85 gave appreciably the most power (approximately 20% more than 95 RON gasoline), with the other blends forming a fairly tight band lying between the two.

Efficiency

Using E100, a full-load thermal efficiency at 2,800 rev/min. of 39.6% was reported by Nakata et al. (2006), compared with 37.9% and 31.7% using the high- and low-octane gasoline, respectively. A thermal efficiency improvement of 3% was achieved using E100 over the 100 RON gasoline at a typical part-load operating point, where the engine was far from the area where knock becomes a limiting factor. The efficiency benefit using ethanol in the part-load region was attributed to the lower heat losses due to the reduction in combustion temperatures. At high loads, when the engine is knock-limited, spark timing must be retarded and this leads to higher exhaust gas temperatures. Typically a pre-turbine temperature limit is imposed. High-octane fuels enable more advanced combustion phasing which in turn requires lower levels of over-fuelling in order to remain within the temperature constraints.

Bergström et al. (2007a, 2007b) found that, using a production turbocharged ethanol-gasoline flex-fuel engine with port-fuel injection, the high knock resistance and concomitant lower exhaust gas temperatures experienced when running on E85 allowed the fuel enrichment level at full-load to be reduced to the extent that, for the same limiting peak pressures as those tolerated using gasoline fuel, stoichiometric operation across the engine speed range is possible. Kapus et al. (2007) found that for identical engine performance, the more favourable combustion phasing when operating on E85 at full load leads to less requirement for fuel enrichment giving a 24% improvement in efficiency compared with operation on 95 RON gasoline. Thermal efficiency improvements at full load of over 35% relative to 95 RON gasoline have been found using E100 in a direct-injection, turbocharged spark-ignition engine (Brewster, 2007) operating at high BMEP levels.

Korte et al. (2011) show detailed fuel flow rate maps for an engine with a peak BMEP of over 30 bar using the fuels described in Section 13.3.2. It was found that whilst an increase in fuel flow was required with the alcohol fuels in order to obtain the same load for part-load operation, when the load was above 20 bar BMEP, operation on E22 (rated at 102 RON) required a lower fuel flow rate than when 95 RON gasoline was used. In order to protect engine components from over-heating, ‘over-fuelling’ is applied. Around the peak power point, operation on E85 required approximately the same fuel flow rate as the 95 RON gasoline whilst the fuel consumption of E22 was about 20% lower. At the peak power point these fuel flow rates correspond to increases of thermal efficiency, relative to the 95 RON gasoline, of 35% and 38% for E22 and E85, respectively. The full-load torque curve of the engine could be achieved at stoichiometric air–fuel ratio using E85.

Dedicated alcohol engines

The greater dilution tolerance of methanol and ethanol was exploited by Brusstar et al. (2002) who converted a base 1.9-litre direct-injection turbocharged diesel engine to run on M100 and E100 by replacing the diesel injectors with spark plugs and fitting a low-pressure alcohol fuel injection system in the intake manifold. Running at the 19.5:1 compression ratio of the base diesel engine, the PI methanol variant increased the peak brake thermal efficiency from 40% to 42%, while parity with the diesel was achieved using ethanol. Cooled exhaust gas recirculation (EGR) enabled the engine to achieve close-to-optimum ignition timing at high loads, while high levels of EGR dilution were used to spread the high efficiency regions to extensive areas of the part-load operating map. This was possible due to the lower molar air–fuel ratios and higher flame speeds of the fuels providing higher combustion tolerance to cooled EGR.

Vancoille et al. (2012) have confirmed the effectiveness of this approach for methanol, investigating the viability of throttle-less load control using EGR or excess air. Experiments performed on a single-cylinder engine showed that the EGR and excess air dilution tolerances of methanol are significantly higher than for gasoline. Using a turbocharged multi-cylinder engine with a 19:1 compression ratio, similar to that used by Brusstar et al. (2002), they found it was possible to use EGR levels of up to 30 mass% at a range of engine speeds and loads with limited cycle-to-cycle combustion variation. A peak thermal efficiency of 42% was obtained. The engine operating maps from this work were used by Naganuma et al. (2012) to model the potential benefits of a dedicated methanol-fuelled vehicle. The predictions indicated that a vehicle using an engine optimized for operation on M100 would increase the average engine efficiency on the NEDC from 22.8% for operation on baseline gasoline to 26% for M100, reducing CO2 emissions from 167 g/km to 132 g/km. Still larger benefits are predicted for other vehicles (Naganuma et al., 2012).

Pollutant emissions, deposits, and lubricant dilution

Bergström et al. (2007b) show that, using a vehicle with a port-fuel-injected LEV2 emissions-capable engine on the US FTP 75 cycle, NOx emissions levels are reduced for all the ethanol blends tested without secondary air injection (SAI). CO emissions levels were increased by up to 50% for the highest concentration ethanol blends tested (E64 and E85) and unburned hydrocarbon emissions (uHC) were increased by about 75% using the E85 blend. They also show that when SAI is used, the uHC and CO emissions are significantly reduced. For uHC emissions, the levels for E85 were reduced by more than a factor of three and became independent of ethanol concentration. Bergström et al. (2007a) show that a similar technology engine calibrated to EU IV emissions compliance using gasoline would also be compliant using E85. Turner et al. (2012c) report emissions results for an EU V vehicle with a boosted DI engine, using a bespoke E85 blend and the iso-stoichiometric ternary blends of gasoline, ethanol and methanol described in Section 13.5.3. While emissions analysers are known to have a lower response to oxygenated species than to pure hydrocarbons, the level of reduction in uHC shown by Turner et al. (2012c) for their blends is of the order that, correcting using a ratio of 1.154 from the earlier work of OudeNijeweme et al. (2011) (measured by them for E85), still resulted in uHC emissions no worse than for gasoline.

Of the various types of uHC emissions, those of the aldehydes in particular are a potential concern, and especially those of formaldehyde. This is because in the metabolism of formaldehyde the crucial step is that from formic acid to carbon dioxide and water, which depends on folic acid and gives rise to a wide variation in fatal dose depending on the victim’s age, body mass and, in the case of women, whether they are pregnant or not. With alcohol combustion, aldehydes are intermediate species of the oxidation of the fuel: methanol primarily yields formaldehyde and ethanol acetaldehyde, together with lower amounts of formaldehyde. Early work, presumably with older engine and emissions control technology than is commonly used today, suggested that generally some changes to catalyst formulation may be necessary to ensure long-term catalyst durability with methanol (Nichols et al., 1988), but this is not expected to be an issue with present-day technology: flex-fuel vehicles have been shown to be capable of meeting limits for formaldehyde when operated on E85 (West et al., 2007) and are expected to be able to do so for other alcohols such as n-butanol (Gingrich et al., 2009). It is known that aldehyde emissions can be successfully neutralized by the type of three-way catalysts typically used to control emissions of modern spark-ignition engines (Menrad et al., 1988; Wagner and Wyszyński, 1996; Shenghua et al., 2007). Gasoline may actually yield greater challenges with respect to aldehydes on legislated drive cycles in the future: Benson et al. (1995) reported acetaldehyde emissions being three times higher in a flex-fuel vehicle when operated on gasoline than E85, and in another study West et al. (2007) reported that gasoline had higher aldehyde emissions than E85 in a vehicle operated on the US06 drive cycle (which requires higher driving loads than is typically the case for other drive cycles). Furthermore, some potential technologies to improve the knock limit of down-sized engines such as cooled EGR have been found to increase aldehyde emissions from gasoline (Gingrich et al., 2009). Pearson and Turner (2012) discuss the potential health effects arising from aldehyde emissions from the combustion of the alcohols as well as giving an overview of other safety aspects of the fuels.

Cairns et al. (2009) found, using a boosted DI engine, that E22 blends produced the highest levels of smoke emissions in their study of ethanol and butanol blends with gasoline. Negligible smoke emissions were found using E85 fuel. Additionally they found that fuel injector deposits were highest using an E10 blend whereas the use of E85 at part load produced almost no deposits. Bergström et al. (2007a) found significant deposits accumulating on a port injector using E85 due to the presence of polyisobutylene components in the gasoline.

Price et al. (2007) measured particulate matter (PM) number concentrations for alcohol-gasoline blends in a direct injection engine. They found that PM number was lowest for E85, with 95 RON gasoline, E30 and M30 blends giving similar PM number concentrations, and M85 giving the highest level. They suggested that with the high-alcohol blends, the potential for PM to form at high load is due to the fact that the local AFR is above the saturation point of the mixture, leading to droplet burning which is only partially offset by the oxygen bound in the fuel. Later work by Chen et al. (2011) showed that under stoichiometric conditions high-ethanol-blend fuels gave higher PM emissions than low-blend ones when operated in a DI engine under homogeneous stoichiometric conditions. Under rich conditions, the PM output from an E10 ethanol blend was much lower, however, and they suggested that this is what gives much lower PM from vehicles operating on alcohol blends when they are operated on transient drive cycles. This suggests that any stratification in the combustion chamber will have some effect on PM formation with high-ethanol-blend fuels. Conversely, Stansfield et al. (2012) reported data from vehicle tests showing that the two high-alcohol blends that they tested, E85 and Bu68, both failed the EU V particulate mass limit, although it should be noted that the production vehicle that they were using had not been sold as being flex-fuel capable.

Bergström et al. (2007a, 2007b) and Hadler et al. (2011) describe the potential for high levels of oil dilution in the engine lubricant using ethanol, particularly if the engine is repeatedly cold-started in a cold climate and run for a short time. Bergström et al. (2007a) state that the build-up of fuel is not harmful as long as the engine is given time to reach normal operating conditions, preferably within ten cold start events.

13.5 Vehicle and blending technologies for alcohol fuels and gasoline

13.5.1 Flexible-fuel vehicles

Ethanol/gasoline flex fuel vehicles have been in existence for many years, the first practical example being the Ford Model T which could operate on gasoline or ethanol (thought at that time to be an attractive feature for potential customers in rural areas). Kramer and Anderson (2012) discuss the market penetration of many forms of alternative fuels and the related vehicles on a worldwide basis, including ethanol. They discuss and define the different types of vehicle and show that some flexibility in the use of different fuels is important to vehicle purchasers for market penetration of both fuels and vehicles. The fact that when ethanol was developed as a transport fuel for Brazil, and that dedicated fuel vehicles were first used which subsequently restricted the uptake of ethanol, versus the acceleration of the market when true flex-fuel vehicles were offered, is cited by them as a clear example of the desirability of fuel flexibility onboard the vehicle. It is interesting to note that in Brazil hydrous ethanol is offered as a fuel and can be used in flex-fuel vehicles which may also have a proportion of gasoline in the tank, easing some of the concern about phase separation (because the vehicles can be made sufficiently flexible in operation to overcome this, and agitation of the tank when the vehicle moves will also tend to keep the various components mixed together). While gasoline starting systems have been used to date on Brazilian ‘total flex’ vehicles, there is a move towards replacing these with heated injectors to permit cold starting even on hydrous E100. On this subject, it is interesting to note that in Brazil, even gasoline (or ‘gasohol’) contains 22–25 vol% ethanol, meaning that the minimum level of alcohol in fuel is equivalent to that which could be realized in the entire fuel pool in the US (or Europe) were all available ethanol eventually blended into it (Anderson et al., 2012a).

In order to ascertain the exact proportion of gasoline and ethanol in the fuel tank so that the engine management parameters can be set accordingly, generally Brazilian FFVs have used a so-called ‘virtual’ sensor to assist in adjusting the fuelling rate. Knowledge of the ratio between the two fuels is necessary because of the different volumetric heating values and stoichiometries of the various fuels supplied to the marketplace. More recently, physical alcohol sensors have become commonplace, especially in markets with challenging emissions and onboard diagnostics requirements. The following section discusses the different fuel sensing technologies and various background reasons for their deployment.

Virtual sensors are so called because during a re-fuelling event, they utilize the fuel tank level sensor to obtain an approximate value for the volume of fuel added, and subsequently employ an algorithm to allow the engine management system (EMS) to calculate two approximate ‘new’ values of the stoichiometric AFR, assuming that the fuel added was either gasoline or E85 only. Because prior to key-off the EMS knew the value of stoichiometry that the vehicle was operating at that moment, since it has calculated two possible new values for stoichiometric AFR, on restart, it monitors the O2 sensors and looks for a perturbation in their signal. Once it has ascertained whether the AFR is swinging in the rich or lean direction, it adjusts the operating parameters gradually until it locks on to the new value, and the vehicle has been ‘conditioned’. Figure 13.11 shows a representation of a virtual sensor detection system. The initial swings to the new AFR can be quite rapid, but an extended period of conditioning is necessary because commercial E85 fuels can feature a reduced ethanol content of 51% and still be within the ASTM D5798 specification, and low ethanol contents have habitually been used in winter to facilitate easier starting and better low-temperature behaviour. Vehicles certified to EU IV level emissions standards in Europe employed such a virtual sensor, an obvious attraction of which is that they are very cost-effective, as they require no additional hardware.

image
13.11 Representation of a virtual sensor system to detect the concentration of ethanol in the fuel system of a gasoline/alcohol flex-fuel vehicle. Reproduced from Turner et al. (2013).

For diagnosis of emissions system performance and reliability under EU V regulations, there is a requirement to diagnose all sensors used by the EMS, meaning that there is a requirement for all FFVs certified to this emissions level to employ an additional sensor capable of directly measuring the concentration of alcohol in the fuel system. This is termed a ‘physical’ sensor. Physical sensors generally measure the electric permittivity (or the resistance) of the fuel. Since these values vary widely between non-polar pure hydrocarbons of the type generally comprising gasoline, and compounds with high polarity (such as the alcohols), they can be used to indicate the alcohol concentration directly. Figure 13.12 shows a representation of a fuel system employing a physical alcohol concentration sensor.

image
13.12 Representation of a physical sensor system to detect the concentration of ethanol in the fuel system of a gasoline/alcohol flex-fuel vehicle. Reproduced from Turner et al. (2013).

Kramer and Anderson (2012) point out that in the US, FFVs have been widely introduced at no on-cost to the consumer, and from this it is reasonable to assume that the on-cost to the vehicle manufacturer, while not negligible, is small enough to be offset against the benefits of producing the vehicle from a corporate average fuel economy (CAFE) standards point of view. In other markets, such as Sweden, the vehicles carry only a modest price penalty, which can quickly be recouped given favourable fuel prices. Bergström et al. (2007a) and Hadler et al. (2011) give a detailed account of the engine and fuel system design modifications required in order to produce an FFV from a dedicated gasoline vehicle platform. In Brazil, the vehicle production cost is slightly more because of the need for a small gasoline-fuelled starting system (itself mainly necessary since hydrous ethanol can be put into the primary tank of these vehicles). Future vehicles could adopt heated components in the fuel system to vaporize the fuel and, through the consequent ability to delete the gasoline-based starting system, reduce the price to closer to that of a simple gasoline vehicle. This would mean that the payback time for the customer would be concomitantly shorter (Kramer and Anderson, 2012). Note that beyond the use of a gasoline start system, Brazilian market vehicles still use flex-fuel technology; the next section will discuss vehicles which do not have the feature of flexible fuelling from a single tank.

13.5.2 Tri-flex-fuel vehicles

As mentioned previously, it is possible to create vehicles capable of operating on any volumetric proportion of methanol, ethanol and gasoline. Pearson et al. (2009a) report one such vehicle which used a physical alcohol sensor to infer the approximate bulk alcohol concentration to set the ignition timing and a wide-range oxygen sensor to provide arbitrary control of the injector pulse-width using the oxygen sensor. Emissions capability of this vehicle was demonstrated by operating it on different ratios of the three fuels.

If a vehicle has originally been developed to operate on high-blend methanol fuels and gasoline, it is relatively straightforward to modify it to accept ethanol as well, as discussed by Nichols (2003). However, if the vehicle has not been developed to function satisfactorily on high proportions of methanol in an alcohol-gasoline system, it is still possible to introduce it into the fuel pool as a major fuel component. This can be achieved through the adoption of targeted blends of gasoline, ethanol and methanol with specific ratios targeting the stoichiometric AFRs of any ethanol-gasoline binary mixture that a flex-fuel vehicle has been developed to operate on. The following section discusses this possibility.

13.5.3 Iso-stoichiometric ternary blends

Gasoline, ethanol and methanol (GEM) are all miscible together (with or without cosolvents to avoid phase separation, which varies with temperature). Ternary, or three-component, blends of them can be configured to have the same target stoichiometric air–fuel ratios (AFRs) as any binary gasoline-ethanol blend. Work that has been conducted to date has concentrated on such ‘GEM’ ternary blends with a target stoichiometric AFR equivalent to that of E85, i.e. 9.7, but equally GEM blends equivalent to E10, E22, etc., could be arranged. For a fixed stoichiometric AFR, the relationship between the three components is defined by linear volumetric relationships as shown for E85-equivalent stoichiometry in Fig. 13.13. Importantly, when configured in such a manner, all such iso-stoichiometric blends not only have near-identical volumetric lower heating value (LHV), it has also been found that they have practically the same octane numbers and extremely close enthalpies of vaporization (to ± 2%, based on the mass ratios in the blend). In Fig. 13.13 one can see that as the volume percentage of ethanol is reduced, so the rate of increase of the methanol proportion is faster than that of the gasoline proportion, because as one volume unit of ethanol is removed, a volume unit of the binary gasoline-methanol mixture with the same stoichiometric AFR as ethanol has to be used to replace it. The necessary volume ratio of gasoline:methanol to give the required 9:1 stoichiometic AFR is 32.7:67.3, as discussed by Turner et al. (2011).

image
13.13 Relationship between blend proportions of gasoline, ethanol and methanol in iso-stoichiometric GEM ternary blends configured with a stoichiometric AFR of 9.7. Reproduced from Turner et al. (2013).

Note that Fig. 13.13 clearly shows that E85 contains no methanol and that its binary equivalent for a gasoline and methanol mixture (where no ethanol is present) occurs at 44 and 56 vol%, respectively. This is the left-hand limit for this stoichiometry. It is interesting to note that the ratio where the proportion of gasoline and methanol are equal occurs at approximately 42.5 vol% ethanol, which is coincidentally half the volume which would be present in E85.

Test results showing that these iso-stoichiometric blends are effectively drop-in fuels, on a vehicle operational level, to E85/gasoline FFVs with both virtual and physical sensor systems are presented by Turner et al. (2011, 2012b). Preliminary investigations into material compatibility have been reported by Turner et al. (2012c), together with results showing that the primary exhaust pollutants can be readily controlled, all E85-equivalent ternary blends essentially out-performing commercial 95 RON gasoline in this respect, although particulate and aldehyde emissions were not measured in these tests. In the US, the use of iso-stoichiometric GEM blends may be especially attractive since there are already 9 million E85/gasoline FFVs in use, a result of their improved performance in terms of gasoline usage under US CAFE regulations. In any location, a progressive subsequent rollout by methanol percentage and region could allow introduction of methanol as a major transport fuel component contributing to meaningfully reduced greenhouse gas emissions, improved energy security and better air quality before a vehicle specification change to M100 has to be adopted (e.g., as proposed by the US Open Fuel Standard). This approach was discussed in Turner et al. (2012c).

Note that, should their use be beneficial with regard to the utilization of all available feedstocks, it is possible to produce ternary blends of other alcohols with gasoline, or even quaternary (and higher) blends. Examples of these may be mixtures of gasoline, methanol and butanol with or without ethanol respectively. A more detailed description of the physico-chemical properties of some of these mixtures, with additional remarks on iso-stoichiometric blends of gasoline, butanol and methanol (GBM), and hydrous GEM and GBM blends, is given by Pearson et al. (2012b).

13.6 Future provision of renewable liquid fuels

The rise in global oil demand over the next 25 years is likely to originate entirely from the transport sectors of emerging economies as increasing prosperity drives greater levels of personal mobility and freight. The development of renewable liquid transport fuels which are not feedstock constrained could enable the continued provision of full-range affordable vehicles and will mitigate, and potentially eliminate, the wealth transfer and exposure to price volatility associated with high dependency on feedstock which have a finite supply. Here the provision of liquid and gaseous fuels which are synthesized from CO2, water and renewable energy are discussed. Without the provision of synthetic carbon-neutral liquid fuels, biofuels are likely to appear as merely palliatives in the processes of alleviating issues regarding security of energy supply and reducing transport CO2 emissions.

13.6.1 The biomass limit

The worldwide sustainable potential of biogenic wastes and residues has been estimated at approximately 50 EJ/year (1 EJ = 1 × 1018 J). Estimates of the global sustainable potential of energy crops have a huge spread, ranging from 30 EJ to 120 EJ/year, depending mainly on the assumptions made regarding food security and retaining biodiversity. These combined values put the total potential sustainable bioenergy supply in 2050 between 80 and 170 EJ/year (Pearson and Turner, 2012). Current global energy use is about 500 EJ/year. Thus the mid-point value of the total sustainable bioenergy supply is around one quarter of this and less than one tenth of the projected global energy use in 2050 (WBGU, 2008). The global transport energy demand in 2007 was about 100 EJ (EIA, 2010) and, extrapolating the EIA value of 143 EJ for liquid fuels in 2035 (EC, 2011), is projected to grow to about 170 EJ in 2050. Assuming that half the available sustainable biomass feedstock was available for biofuel production at a conversion efficiency of 50% (Bandi et al., 1995) limits the substitution potential of biofuels to about 20% of the 2050 energy demand. The IEA make the slightly more optimistic prediction that 32 EJ of biofuels will be used globally in 2050, providing 27% of transport fuel (IEA, 2011b). For individual countries, the biomass potential could be significantly higher or lower, depending on their population densities and sustainable agricultural potential.

A further limiting factor is the requirement to feed the increasing global population, with its shift toward westernized diets demanding much greater amounts of land and water (UN, 2006). These issues may constrain biofuel production to the use of the wastes and residues quantified above.

13.6.2 Sustainable Organic Fuels for Transport (SOFT)

In this section, the concept of Sustainable Organic (meaning carbon-containing) Fuels for Transport (SOFT) is briefly introduced as a means of circumventing the biomass limit to the penetration of biofuels in transport. The concept is proposed as a long-term solution to supplying carbon-neutral liquid fuels which could eventually supply the bulk of transport energy demand beyond the 20–30% which can be sustainably supplied by biofuels. These fuels, like biofuels, would be miscible with current petroleum-based fuels so that an evolutionary transition from one organic liquid fuel to another could occur. Whilst Section 13.2 described the problems of using molecular hydrogen as a transport fuel, here it is proposed to ‘package’ the hydrogen in a more convenient manner by combining it with re-cycled CO2 to synthesize energy-dense liquid fuel. In this way hydrogen is used in the fuel rather than as the fuel. Since the concept is based on the reduction of CO2 and water to synthesize the fuel and its subsequent use results in oxidation returning the fuel to these components, the process is not feedstock-limited. Importantly, if all processes are powered with carbon-free energy and the CO2 used to make the fuel is captured directly from the atmosphere, then the combustion of this fuel would result in zero net increase in the atmospheric CO2 concentration. The concept of synthesizing fuel from water and recycled CO2 was first proposed in the 1970s by Steinberg (1977) and there have been many other proposals in the meantime (Bandi et al., 1995; Stucki et al., 1995; Weimer et al., 1996; Olah et al., 2009, 2011; Jensen et al., 2007; Pearson and Turner, 2007; Pearson et al., 2009b, 2012a; Littau, 2008; Jiang et al., 2010; Graves et al., 2011).

Recycling CO2

Synthetic carbon-neutral liquid fuels mimic the overall function of photosynthesis by reducing water and CO2 to make hydrocarbon-based products. The energy input to the process can either be provided directly in the form of sunlight (to produce ‘solar fuels’) or may be off-peak renewable energy (see Section 13.6.3). By combining hydrogen with CO2 it can be chemically liquefied into a high energy density hydrocarbon fuel. Clearly, if the captured CO2 stems from the combustion of fossil energy resources, this approach is not renewable and will still result in an increase in atmospheric CO2 concentration. Rather than a recycling process, it amounts to CO2 re-use and offers the potential of a notional reduction in emissions of approximately 50% (Graves et al., 2011). In a closed-cycle fuel production process, ideally there is no net release of non-renewable CO2. If the hydrogen generation process is via the electrolysis of water, this represents by far the greatest energy input to the process. For this reason the fuels produced in this way may be referred to as ‘electrofuels’, as they are essentially vectors for the storage and distribution of electricity generated from renewable energy. When the feedstocks are water and CO2 from the atmosphere, the fuel production and use cycle is materially closed and therefore sustainable. Such a cycle offers security of feedstock supply on a par with that of the ‘hydrogen economy’, as the timescale for mixing of CO2 in the atmosphere is sufficiently short to ensure a homogeneous distribution. With access to sufficient water and renewable energy, the process has the potential to provide fuel from indigenous resources for most nations.

CO2 and water are the end products of any combustion process involving materials containing carbon and hydrogen. Further reactions to form carbonates are exothermic processes. The capture of CO2 in inorganic carbonates and other media is a burgeoning area of research and some of the literature is described in detail in Graves et al. (2011) and Pearson et al. (2012a).

Whereas the adoption of battery electric or hydrogen fuel cell vehicles requires paradigm shifts in the costs of the vehicles themselves or their fuel distribution infrastructure, or both (Pearson et al., 2012a), the development of carbon-neutral liquid fuels enables a contiguous transition to sustainable transport. Drop-in fuels such as gasoline, diesel and kerosene can be produced from CO2 (via CO) and H2 via Fischer–Tropsch (FT) synthesis, but the simplest and most efficient liquid fuel to make is methanol (Pearson and Turner, 2012). Indeed the option to make gasoline is retained even if methanol is produced initially since the former can be made via the Exxon-Mobil methanol-to-gasoline process. In addition to being the simplest fuel to synthesize from CO2 and water feedstocks, methanol provides much greater biomass feedstock diversity, as it can be made from anything which is (or ever was) a plant.

Fuel synthesis

Once hydrogen and CO2 are available, the simplest and most direct route to producing a high-quality liquid fuel is the catalytic hydrogenation of CO2 to methanol via the reaction:

CO2+3H2CH3OH+H2OΔH2980=49.9kJ/mol.methanol

image

During the production of methanol via direct hydrogenation of CO2, by far the largest component of the process energy requirement is the hydrogen production (Pearson and Tuner, 2012; Pearson et al., 2012a). This is true of any electrofuel using hydrogen as an intermediate or final energy carrier. Assuming an electrolyser efficiency of 80% and a CO2 extraction energy of 250 kJ/mol.CO2 (representing about a 10% rational thermodynamic efficiency relative to the minimum thermodynamic work requirement of 20 kJ/mol.CO2) gives a higher heating value (HHV) ‘electricity-to-liquid’ efficiency of about 45%, including multi-pass synthesis of the methanol and re-compression of the unconverted reactants (Pearson and Turner, 2012; Pearson et al., 2012a). In the late 1990s, Specht et al. (1998a, 1998b) measured total process CO2 capture energy levels of 430 kJ/mol, in a demonstration plant using an electrodialysis process to recover the absorbed CO2, representing a rational efficiency of less than 5%. Despite this low CO2 capture and concentration efficiency, the measured overall fuel production efficiency was 38%.

Without policy intervention, the intermittent use of alkaline electrolysers, due to their limited current densities, is likely to be too expensive (Graves et al., 2011) to produce fuel under present market economics. Improvements to this technology are at an advanced stage of development (Graves et al., 2011; Ganley, 2009) and other promising technologies are emerging. Graves et al. (2011) describe the use of high temperature co-electrolysis of CO2 and H2O giving close to 100% electricity-to-syngas efficiency for use in conventional FT reactors. This ultra-efficient high temperature electrolysis process using solid oxide cells combined with a claimed CO2 capture energy (from atmospheric air) as low as 50 kJ/mol (Lackner, 2009) leads to a prediction of an electricity-to-liquid efficiency of 70% (HHV basis). With a constant power supply this high overall efficiency enables the production of synthetic gasoline at $2/gallon ($0.53/litre) using electricity available at around $0.03/kWh (Graves et al., 2011). Doty et al. (2010) state that off-peak wind energy in areas of high wind penetration in the US averaged $0.0164/kWh in 2009 and the lowest 6 hours of the day averaged $0.0071/kWh. With more pessimistic values for the cost of CO2 capture such as the $1,000/tonne quoted by House et al. (2011), the gasoline cost component due to the supply of the carbon feedstock alone might be as high as $7.5/gallon (about €1.30/litre). With 20% electrolyser capacity the cost of fuel synthesis could be as high as $4/gallon at $0.03/kWh electricity (higher current density electrolysers could reduce this to $2.2/gallon) (Graves et al., 2011). For perspective, a cost of $11.5/gallon is around €2.05/litre. Currently gasoline retail costs in the EU range from €1.14/litre to €1.67/litre including duties and taxes (which can be as high as €0.6/litre).

In a reconfigured system which bases fuel duty and taxation on non-renewable life cycle carbon intensity, a fuel made from air-extracted CO2 and water might be commercially attractive in the medium term, i.e. in advance of the point where sequestration of air-captured CO2 becomes economically feasible. Recycling of the CO2, rather than sequestrating it after it has been removed from the atmosphere, cannot result in any net greenhouse gas (GHG) reduction. Its inclusion in a closed carbon cycle to make transport fuels, however, can potentially have the effect of rendering carbon-neutral the fastest growing GHG emissions sector as well as providing a spur to the development of air capture technology for CO2 which may subsequently be used for sequestration purposes.

13.6.3 Renewable fuels within an integrated renewable energy system

Renewable energy sources, such as those based on wind and solar power, are limited in their ability to meet current and future energy demands not by the resource potential, which for wind and solar is many times current demand, but by the intermittent nature of supply. To escape this conundrum, large-scale storage systems are required. Biomass is a form of large-scale storage of solar energy but, whilst it may be part of a sustainable system, it cannot underpin it. One possibility for large-scale energy storage is to use off-peak renewable energy to synthesize chemical energy carriers. Chemical energy storage systems, based on the conversion of renewable energy into a gaseous or liquid energy carrier, enable the stored energy to be either re-used for power generation or transferred to other energy sectors such as transport, where the de-carbonization issue is more problematic, and there is an ever-present demand to supply a high-value energy carrier.

In addition to the ready synthesis of methanol from CO2 and hydrogen, methane can also be made from the same feedstocks using the Sabatier process. It then has the advantage that it can be stored in the gas grid, which, for most developed countries, far exceeds the capacity of existing renewable energy storage media (e.g., pumped hydro) or proposed systems such as large flywheels or redux fuel cells. Sterner (2009), Specht et al. (2009), and Breyer et al. (2011) describe such a concept, where the synthesized methane is stored and readily retrievable to smooth out the supply of renewable energy through conversion back to electrical energy via combustion in conventional power stations. This process is given the name ‘renewable power methane’ (RPM), and its operation within a renewable energy system based on wind, solar and biomass has been modelled over a period of one week on a one hour resolution based on a winter load demand. The renewable-power-to-methane efficiency is predicted to be 48% (Sterner, 2009; Specht et al., 2009) using measured energy values for capture and concentration of CO2 from air of 430 kJ/mol, which, as discussed in Section 13.6.2, is a realistic value achievable with current cost-effective ‘off-the-shelf’ technology.

Because the limit on the amount of renewable energy that can be stored would now be significantly beyond any expected day-to-day variation in energy utilization, the upper limit on the level of renewables in which it is economically attractive to invest is effectively raised. While the ‘round-trip’ losses in converting renewable electricity to methane and back to electricity are significant and would amount to an expected process efficiency of about 20% (electricity-to-electricity), the entire process is practical, achievable and, most importantly, represents an evolutionary path to a fully sustainable energy economy. The production of renewable electricity and RPM for power generation reserve (and its related use in the heat sector, where it would displace fossil natural gas) can also be integrated with the production of liquid fuels, in the form of both biofuels and electrofuels, for use directly in transport (Pearson et al., 2012a). A schematic representation of such a system combining the power, heat and transport sectors is shown in Fig. 13.14, where the renewable liquid fuels are represented by methanol (CH3OH), ethanol (C2H5OH) and hydrocarbon fuels (n(-CH2-)). The latter can be synthesized from methanol using such processes as the methanol-to-gasoline (MTG) process. In this system, where it can be sustainably produced, biomass is able to contribute to the transport sector via the production of ethanol (or biodiesel), or the energy sector via direct combustion for power generation, as shown in Fig. 13.14.

image
13.14 Integrated power, heat and transport system featuring large-scale energy storage capacity and combining renewable methane and liquid fuels and using renewable energy and biomass.

13.7 Conclusion

The production of low-carbon and, ultimately, carbon-neutral liquid fuels is the most pragmatic way in which to either de-carbonize transport or increase its sustainability and thus increase the security of energy supply to the transport sector. This route also allows the continued provision of globally compatible, affordable transport via the retention of low-cost internal combustion engines supplied both on and off the vehicle by low-cost liquid fuel systems.

Low-level biofuel blends, particularly in the form of E10, are already displacing significant quantities of gasoline in countries such as the United States. Ethanol can provide fuels with high resistance to auto-ignition which are synergistic with the trend towards downsized pressure-charged spark-ignition engines. Without significant changes to the fuel taxation system, the low volumetric energy density of ethanol is likely to limit the market penetration of high concentration blends such as E85. Biodiesel penetration is limited by the reluctance of manufacturers to warrant their vehicles for use with higher levels of FAME. These factors, together with the limitations on vegetable oil feedstocks, may lead to higher levels of ethanol in gasoline in order to satisfy energy security, renewability or GHG reduction targets. Butanol and methanol may also have a part to play.

Low-carbon-number alcohols can be used for personal mobility and light-duty applications, and synthetic hydrocarbons for applications where maximum energy density is crucial (such as aviation). A range of technologies is described to enable the transition from the current vehicle fleet to equivalent-cost vehicles capable of using sustainable methanol. All transport energy can be supplied, using biofuels up to the biomass limit, and beyond it using carbon-neutral liquid fuels made using renewable energy, water and re-cycled CO2 from the atmosphere – this approach uses hydrogen in the fuel rather than using it as the fuel. The role of biofuels in this transitional route and end-game would prevent them being regarded as a dead-end by vehicle manufacturers and politicians alike, thus ensuring their continued production and development.

13.8 Acknowledgements

The authors wish to thank the following people for the benefit of discussions with them during the compilation of this work: Arthur Bell (SASOL), Martin Davy (Loughborough University), Eelco Dekker (BioMCN), Peter Edwards (University of Oxford), Matt Eisaman (Brookhaven National Laboratories), Stefan de Goede (SASOL), Ben Iosefa (Methanex), Leon diMarco, Richard Stone (University of Oxford), Andre Swartz (SASOL), Gordon Taylor (GT-Systems), Sebastian Verhelst (University of Ghent), Chris Woolard (University of Cape Town), and Paul Wuebben (CRI). The authors would also like to acknowledge the funding of HGCA and DEFRA during the ‘HOOCH’ project during which some of the concepts discussed above were developed.

13.9 References and further reading

1. Amaseder F, Krainz G. Liquid hydrogen storage systems developed and manufactured for the first time for customer cars. SAE paper number 2006-01-0432 Detroit, MI: SAE 2006 World Congress; 3–6 April.

2. Andersen VF, Anderson JE, Wallington TJ, Mueller SA, Nielsen OJ. Vapor pressures of alcohol–gasoline blends. Energy Fuels. 2010;24:3647–3654.

3. Anderson JE, DiCicco DM, Ginder JM, et al. High octane number ethanol–gasoline blends: quantifying the potential benefits in the United States. Fuel. 2012a;97:585–594.

4. Anderson JE, Leone TG, Shelby MH, et al. Octane numbers of ethanol-gasoline blends: Measurements and novel estimation method from molar composition. 2012b; SAE paper no. 2012-01-1274.

5. ARB. Low carbon fuel standard program. California Environmental Protection Agency – Air Resources Board. Available from: http://www.arb.ca.gov/fuels/lcfs/lcfs.htm; (accessed 19 February 2012).

6. Bamgboyne AJ, Hansen AC. Prediction of cetane number of biodiesel fuel from the fatty acid methyl ester (FAME) composition. Int Agrophysics. 2008;22:21–29.

7. Bandi A, Specht M, Weimer T, Schaber K. CO2 recycling for hydrogen storage and transportation – electrochemical CO2 removal and fixation. Energy Conversion and Management. 1995;36(6–9):899–902.

8. Benson JD, Koehl WJ, Burns VR, et al. Emissions with E85 and gasolines in flexible/variable fuel vehicles – The auto/oil air quality improvement research program. SAE paper no. 952508 1995.

9. Bergström K, Melin S-A, Jones CC. The new ECOTEC turbo BioPower engine from GM Powertrain – utilizing the power of nature’s resources. In: 28th International Vienna Motor Symposium, Vienna, 26–27 April. 2007a.

10. Bergström K, Nordin H, Konigstein A, Marriott CD, Wiles MA. ABC – Alcohol based combustion engines – challenges and opportunities. In: 16th Aachener Kolloquium Fahrzeug- und Motorentechnik, Aachen, 8–10 October. 2007b.

11. Best-europe. World’s first ethanol powered diesel car. Available from: www.best-europe.org/Pages/ContentPage.aspx?id=488; (accessed 17 December 2008).

12. Blumberg PN, 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. 2009;1(1):1186–1195.

13. Bossel U. Does a hydrogen economy make sense? Proc, IEEE. 2006;84(10):1826–1837.

14. Brewster S. Initial development of a turbo-charged direct injection E100 combustion system. SAE paper no. 2007-01-3625.

15. Breyer Ch, Rieke S, Sterner M, Schmid J. Hybrid PV-wind-renewable methane power plants – a potential cornerstone of global energy supply. In: 26th European Photovoltaic Solar Energy Conference, Hamburg, Germany, 5–9 September. 2011.

16. Bringezu S, Schutz H, O’Brien M, Kauppi L, Howarth RW, McNeely J. Towards sustainable production and use of resources: assessing Biofuels. In: International Panel for Sustainable Resource Management, United Nations Environment Programme. Available from: http://www.unep.org/pdf/Assessing_Biofuels-full_report-Web.pdf; (accessed 18 February 2012).

17. Bruetsch RI, Hellman KH. Evaluation of a passenger car equipped with a direct injection neat methanol engine. 1992; SAE paper no. 920196.

18. Brusstar MJ, Gray CL. High efficiency with future alcohol fuels in a stoichiometric medium duty spark ignition engine. SAE paper no. 2007-01-3993.

19. Brusstar M, Stuhldreher M, Swain D, Pidgeon W. High efficiency and low emissions from a port-injected engine with neat alcohol fuels. SAE paper no. 2002-01-2743.

20. Brusstar M, Haugen D, Gray C. Environmental and human health considerations for methanol as a transportation fuel. In: 17th Int. Symp. on Alt. Fuels, Taiyuan, China, 14 October. 2008.

21. Cairns A, Stansfield P, Fraser N, et al. A study of gasoline-alcohol blended fuels in an advanced turbocharged DISI engine. SAE paper no. 2009-01-0138.

22. Caprotti R, Breakspear A, Klaua T, Weiland P, Graupner O, Bittner M. RME behaviour in current and future diesel fuel FIEs. SAE paper no. 2007-01-3982.

23. Casten S, Teagan P, Stobart R. Fuels for fuel cell-powered vehicles. SAE paper no. 2000-01-0001.

24. Chen L, Xu F, Stone R, Richardson D. Spray imaging, mixture preparation and particulate matter emissions using a GDI engine fuelled with stoichiometric gasoline/ethanol blends. Paper Number C1328/002 In: Internal Combustion Engines: Improving Performance, Fuel Economy, and Emissions, Institution of Mechanical Engineers Conference, London, 29–30 November. 2011;43–52.

25. Cooper J. Future fuels for transport: regulatory and supply drivers. In: Internal Combustion Engines: Improving Performance, Fuel Economy, and Emissions, Institution of Mechanical Engineers Conference, London, 29–30 November. 2011.

26. Davis SC, Diegel SW, Boundy RG. Transportation Energy Data Book. Edition 30 Oak Ridge, TN: Center for Transportation Analysis, Energy and Transportation Science Division, Oak Ridge National Laboratory; 2011; Prepared for the Vehicle Technologies Program, Office of Energy Efficiency and Renewable Energy, US Department of Energy, ORNL-6986.

27. Doty GN, McCree DL, Doty JM, Doty FD. Deployment prospects for proposed sustainable energy alternatives in 2020. Paper ES 2010-90376 In: Proceedings of ES2010 Energy Sustainability 2010, Phoenix, AZ, 17–22 May.

28. Eberle U. GM’s research strategy:Towards a hydrogen-based transportation system. Hamburg: FuncHy Workshop; 2006; September.

29. Eberle U, Arnold G, von Helmolt R. Hydrogen storage in metal-hydrogen systems and their derivatives. J Power Sources. 2006;154:456–460.

30. EC. On the promotion of the use of energy from renewable sources and amending and subsequently repealing Directives 2001/77/EC and 2003/30/EC. In: Directive 2009/28/EC of the European Parliament and of the Council, 23 April. 2009a.

31. EC. Setting emission performance standards for new passenger cars as part of the Community’s integrated approach to reduce CO2 emissions from light-duty vehicles. In: Regulation (EC) 443/2009 of the European Parliament and of the Council, 23 April. 2009b.

32. EC. Amending Directive 98/70/EC as regards the specification of petrol, diesel, and gas-oil and introducing a mechanism to monitor and reduce greenhouse gas emissions and amending Council Directive 1999/32/EC as regards the specification of fuel used by inland waterway vessels and repealing Directive 93/12/EEC. In: Directive 2009/30/EC of the European Parliament and of the Council of 23 April 2009. 2009c.

33. EC. Comité Européen de Normalisation, Automotive fuels – diesel Requirements and test methods. 2010; E 590. February.

34. EC. Roadmap to a single European transport area – Towards a competitive and resource efficient transport system. In: White Paper COM(2011) 144 final, European Commission, 28 March.

35. Edwards PP. Private communication. 2012.

36. EIA. International energy outlook 2010: Transport sector energy consumption. US Energy Information Administration. Available from: http://www.eia.doe.gov/oiaf/ieo/transportation.html; (accessed 20 April 2011).

37. EPA. Regulation of fuels and fuel additives: Renewable Fuel Standard program: final rule. 40 CFR Part 80. Federal Register Environmental Protection Agency. vol. 72(no. 83):23900–24014 01/05. Available from: http://www.epa.gov/otaq/renewablefuels/rfs-finalrule.pdf; 2007; (accessed 19 February 2012).

38. EPA. Regulation of fuels and fuel additives: changes to Renewable Fuel Standard program: final rule. 40 CFR Part 80. Federal Register Environmental Protection Agency. vol. 75(no. 58):14670–14904 26/03. Available from: http://www.epa.gov/otaq/renewablefuels/rfs-finalrule.pdf; 2010a; (accessed 19 February 2012).

39. EPA. Partial grant and partial denial of Clean Air Act waiver application submitted by Growth Energy to increase the allowable ethanol content of gasoline to 15 percent; Decision of the administrator. Federal Register Environmental Protection Agency. 2010b; vol. 75(no. 213):68094–68150 04/11. Available from: http://www.gpo.gov/fdsys/pkg/FR-2010-11-04/pdf/2010-27432.pdf; 2010b; (accessed 19 February 2012).

40. EPA. Light-duty vehicle greenhouse gas emissions standards and corporate average fuel economy standards: final rule. Environmental Protection Agency/Department of Transportation/National Highway Traffic Safety Administration. 2010c;vol. 75(no. 88):25324–25728 40 CFR Parts 85, 86, and 600; 49 CFR Parts 531, 533, 536. Federal Register, 07/05.

41. EPA. EPA finalizes 2012 Renewable Fuel Standards. In: EPA Regulatory Announcement, Office of Transportation and Air Quality, EPA-420-F-11-044, December. Available from: http://www.epa.gov/otaq/fuels/renewablefuels/documents/420f11044.pdf; 2011a; (accessed 19 February 2012).

42. EPA. Partial grant of Clean Air Act waiver application submitted by Growth Energy to increase the allowable ethanol content of gasoline to 15 percent; Decision of the administrator. Federal Register Environmental Protection Agency. 2011b; vol. 76(no. 17):4662–4683 26/01. Available from: http://www.gpo.gov/fdsys/pkg/FR-2011-12-01/pdf/2011-30358.pdf; 2011b; (accessed 19 February 2012).

43. EPA. 2017 and later model year light-duty vehicle greenhouse gas emissions and corporate average fuel economy standards. Environmental Protection Agency/Department of Transportation/National Highway Traffic Safety Administration. 2011c;vol. 76(no. 231):74854 40 CFR Parts 85, 86, and 600; 49 CFR Parts 523, 531, 533, 536, and 537. Federal Register, 25728, 01/12.

44. EPA. MTBE in fuels. 2012; Available from: http://www.epa.gov/mtbe/gas.htm; (accessed 19 February 2012).

45. Fargione J, Hill J, Tilman D, Polasky S, Hawthorne P. Land clearing and the biofuel carbon debt. Science Express. 2008;319:1235–1238.

46. Ferrari RA, Turtelli Pighinelli ALM, Park KJ. Biodiesel production and quality. In: dos Santos Bernardes MA, ed. Biofuel’s Engineering Process Technology. Rijeka, Croatia: InTech; 2011; In: http://www.intechopen.com/books/biofuel-s-engineering-process-technology/biodiesel-production-and-quality.

47. Furey RL. Volatility characteristics of gasoline-alcohol and gasoline-ether blends. 1985; SAE paper number 852116.

48. Ganley JC. High temperature and pressure alkaline electrolysis. Int J Hydrogen Energy. 2009;34(9):3604–3611.

49. Gardiner T, Gaade J, Head B, Hygate C, Xu HM, Abdullah NR. Research into requirements for worldwide sustainable biodiesel capability. In: Internal Combustion Engines: Improving Performance, Fuel Economy and Emissions, Institution of Mechanical Engineers Conference, London, 29–30 November. 2011.

50. Geller DP, Goodrum JW. Effects of specific fatty acid methyl esters on diesel fuel lubricity. Fuel. 2004;83:2351–2356.

51. Gingrich J, Khalek I, Alger T, Mangold B. Consideration of emissions standards for a dilute spark-ignited engine operating on gasoline, butanol, and E85. In: SIA International Conference: The Spark-Ignition Engine of the Future, Strasbourg, France, 3–4 December. 2009.

52. Graves C, Ebbesen SD, Mogensen M, Lackner K. Sustainable hydrocarbon fuels by recycling CO2 and H2O with renewable or nuclear energy. Renewable and Sustainable Energy Reviews. 2011;15:1–23.

53. Hadler J, Szengel R, Middendorf H, Sperling H, Groer H-G, Tilchner L. The 1.4 l 118 kW TSI for E85 mode – expansion of the most economical line of petrol engine from Volkswagen. In: 32nd International Vienna Motor Symposium, Vienna, 5–6 May. 2011.

54. Haer G. Delivering now: US biodiesel market update and outlook 2012. In: IHS World Methanol Conference, San Diego, CA, 6–7 December. 2011.

55. Hikino K. Development of methanol engine with autoignition for low NOx and better fuel economy. SAE paper number 891842 In: Milwaukee, WI: SAE International Off-Highway & Powerplant Congress and Exposition; 1989.

56. House KZ, Baclig A, Ranjan M, van Nierop E, Wilcox J, Herzhog H. Economic and energetic analysis of capturing CO2 from ambient air. PNAS Early Edition 2011; In: www.pnas.org/cgi/doi/10.1073/pnas.1012253108.

57. Hunwartzen I. Modification of CFR test engine unit to determine octane numbers of pure alcohols and gasoline-alcohol blends. SAE paper no. 820002 In: Society of Automotive Engineers International Congress and Exposition, Detroit, MI, 22–26 February. 1982.

58. IEA. Advanced Motor Fuels Annual Report. Paris: International Energy Agency; 2010; Available from: http://www.iea-amf.vtt.fi/pdf/annual_report_2010.pdf; (accessed 02 January 2012).

59. IEA. World Energy Outlook Paris: International Energy Agency; 2011a; Available from: http://www.iea.org/Textbase/npsum/weo2011sum.pdf; (accessed 05 December 2011).

60. IEA. Technology road map: biofuels for transport. 2011b; International Energy Agency. Available from: http://www.iea.org/papers/2011/biofuels_roadmap.pdf; (accessed 01 February 2012).

61. Jackson MD, Unnasch S, Sullivan C, Renner RA. Transit bus operation with methanol fuel. 1985; SAE paper no. 850216.

62. Jackson N. Low carbon vehicle strategies – options and potential benefits. In: Cost-Effective Low Carbon Engines Conference, I.Mech.E., London, November. 2006.

63. Jensen SH, Larsen PH, Mogensen M. Hydrogen and synthetic fuel production from renewable energy sources. Int J Hydrogen Energy. 2007;32:3253–3257.

64. Jiang Z, Xiao T, Kuznetsov VL, Edwards PP. Turning carbon dioxide into fuel. Phil Trans R Soc A. 2010;368:3343–3364.

65. Kapus PE, Fuerhapter A, Fuchs H, Fraidl GK. Ethanol direct injection on turbocharged SI engines – potential and challenges. SAE paper number 2007-01-1408.

66. Korte V, OudeNijeweme D, Bisordi A, et al. Downsizing and biofuels: synergies for significant CO2 reductions. In: 20th Aachen Colloquium Automobile and Engine Technology, Aachen, Germany, 10–12 October. 2011.

67. Kramer U, Anderson JE. Prospects for flexible- and bi-fuel light duty vehicles: consumer choice and public attitudes. In: 2012 MIT Energy Initiative Symposium: Prospects for Flexible- and Bi-Fuel Light Duty Vehicles, Cambridge, MA, 19 April.

68. Lackner KS. Capture of carbon dioxide from ambient air. Eurpoean Physical Journal – Special Topics. 2009;176(1):93–106.

69. Li W, Zhong L, Xie K. The development of methanol industry and methanol fuel in China. In: 6th Annual Methanol Forum, Dubai, 3–5 November. 2008.

70. Littau K. An “atmospherically healthy” recipe for carbon-neutral fuels: synthetic fuel made from sunlight, CO2, and water. In: CTSI Clean Technology & Sustainable Industries Conference & Trade Show, Boston, MA, 1–5 June. 2008.

71. Marriott CD, Wiles MA, Gwidt JM, Parrish SE. Development of a naturally aspirated spark ignition direct-injection flex-fuel engine. 2008 SAE paper no. 2008-01-0319.

72. McCormick RL, Alleman TL, Waynick JA, Westbrook SR, Porter S. Stability of biodiesel and biodiesel blends: interim report. Technical Report NREL/TP-540-39721 National Renewable Energy Laboratory 2006; April.

73. Menrad H, Bernhardt W, Decker G. Methanol vehicles of Volkswagen – a contribution to better air quality. 1988; SAE paper number 881196.

74. Mintz M, Folga S, Molburg J, Gillette J. Cost of some hydrogen fuel infrastructure options. In: Argonne National Laboratory Transportation Technology R&D Center, Transportation Research Board, 6 January. 2002.

75. Miyata I, Takei Y, Tsurutani K, Okada M. Effects of bio-fuels on vehicle performance – degradation mechanism analysis of bio-fuels. 2004; SAE paper no. 2004-01-3031.

76. Moran DP, Taylor AB. An evaporative and engine-cycle model for fuel octane sensitivity prediction. 1995; SAE paper no. 952524.

77. Moreira JR, Coelho ST, Velazquez SMSG, Apolinario SM, Melo EH, Elmadjian PHB. BEST project – Contribution of ethanol usage in public urban transport. In: 2008; Available from: www.aea.org.br/twiki/pub/AEA/PAPERS/PAP0018-17.09-13h30-AnditrioIPE.pdf; (accessed 1 November 2011).

78. Naganuma K, Vancoillie J, Verhelst S, et al. Drive cycle analysis of load control strategies for methanol fuelled ICE vehicle. SAE paper number 2012-01-1606 In: SAE Powertrains, Fuels and Lubricants Meeting, Malmo, Sweden, 18–20 September.

79. Nakata K, Utsumi S, Ota A, Kawatake K, Kawai T, Tsunooka T. The effect of ethanol fuel on a spark ignition engine. SAE paper no. 2006-01-3380.

80. National Standard of the People’s Republic of China. Methanol gasoline (M85) for motor vehicles. Standardization Administration of the PRC, GB/T23799-2009, ICS 75.160.20. E31. Issued 18 May; Implemented 1 December.

81. Nichols RJ. The methanol story: a sustainable fuel for the future. J Sci Ind Research. 2003;62:97–105.

82. Nichols RJ, Clinton EL, King ET, Smith CS, Wineland RJ. A view of flexible fuel vehicle aldehyde emissions. 1988; SAE paper number 881200.

83. Niu J, Shi L. Methanol used as vehicle fuel will become a main alternative fuel in China. In: XIX ISAF International Symposium on Alcohol Fuels, Verona, Italy, 10–14 October. 2011.

84. Ogawa T, Kajiya S, Kosaka S, Tajima I, Yamamoto M. Analysis of oxidative deterioration of biodiesel fuel SAE paper 2008-01-2502.

85. Olah GA, Goeppert A, Prakash GKS. Beyond Oil and Gas: The Methanol Economy. 2 edn Weinheim: Wiley-VCH; 2009.

86. Olah GA, Prakash GK, Goeppert A. Anthropogenic chemical carbon cycle for a sustainable future. J Am Chem Soc. 2011;133(33):12881–12898.

87. OudeNijeweme D, Stansfield P, Bisordi A, et al. Significant CO2 reductions by utilizing the synergies between a downsized SI engine and biofuels. In: Internal Combustion Engines: Improving Performance, Fuel Economy and Emissions, Institution of Mechanical Engineers Conference, London, 29–30 November. 2011.

88. Owen K, Coley T. Automotive Fuels Reference Book. 2nd edn Warrendale, PA: Society of Automotive Engineers; 1995.

89. Pang P. China coal chemical development and outlook. In: IHS World Methanol Conference, San Diego, CA, 6–7 December. 2011.

90. Pearson RJ, Turner JWG. Exploitation of energy resources and future automotive fuels. SAE paper number 2007-01-0034 In: SAE Fuels and Emissions Conference, Cape Town, South Africa, January.

91. Pearson RJ, Turner JWG. Renewable fuels: an automotive perspective. In: Oxford: Elsevier; 2012;305–342. Sayigh A, ed. Comprehensive Renewable Energy. vol. 5.

92. Pearson RJ, Turner JWG, Peck AJ. Gasoline-ethanol-methanol tri-fuel vehicle development and its role in expediting sustainable organic fuels for transport. In: 2009 I.Mech.E. Low Carbon Vehicles Conference, London, 20–21 May. 2009a;89–110.

93. Pearson RJ, Turner JWG, Eisaman MD, Littau KA. Extending the supply of alcohol fuels for energy security and carbon reduction. SAE paper no. 2009-01-2764 In: SAE Powertrains, Fuels and Lubricants meeting, San Antonio, TX, 2–4 November. 2009b.

94. Pearson RJ, Eisaman MD, Turner JWG, et al. Energy storage via carbon-neutral fuels made from CO2, water, and renewable energy. Special Issue of Proc IEEE: ‘Addressing the intermittency challenge: Massive energy storage in a sustainable future. 2012a;100(2):440–460.

95. Pearson RJ, Turner JWG, Bell A, de Goede S, Woolard C, Davy M. Iso-stoichiometric fuel blends: characterization of physico-chemical properties for mixtures of gasoline, ethanol, methanol, and water. Journal of Automotive Engineering, Part D 2012b; in preparation.

96. Pickard W, Shen A, Hansing J. Parking the power: strategies and physical limitations for bulk energy storage in supply–demand matching on a grid whose input power is provided by intermittent sources. Renewable and Sustainable Sustainable Energy Reviews. 2009;13:1934–1945.

97. Price P, Twiney B, Stone R, Kar K, Walmsley H. Particulate and hydrocarbon emissions measurements from a spray guided direct injection spark ignition engine with oxygenate fuel blends. SAE paper no. 2007-01-0472.

98. RFA. The Gallagher Review of the Indirect Effects of Biofuels Production St-Leonards-on-Sea, UK: Renewable Fuels Agency; 2008; July.

99. Richards P, Ried J, Tok L-H, MacMillan I. The emerging market for biodiesel and the role of fuel additives. SAE paper no. 2007-01-2033 (JSAE paper no. 20077232).

100. Rodgers RD, Voth GA. Ionic liquids. Acc Chem Res. 2007;40:1077–1078.

101. Saito K, Kobayashi S, Tanaka S. Storage stability of FAME blended diesel fuels. SAE paper no. 2008-01-2505.

102. Searchinger T, Heimlich R, Houghton RA, et al. Use of US croplands for biofuels increases greenhouse gases through emissions from land-use change. Science Express. 2008;319:1238–1240.

103. SEKAB. Ethanol also for lorries. press information, 17 March. Available from: www.sekab.com; 2008.

104. Shenghua L, Clemente ERC, Tiegang H, Yanjv W. Study of spark ignition engine fueled with methanol/gasoline blends. Applied Thermal Engineering. 2007;27:1904–1910.

105. Siewart RM, Groff EG. Unassisted cold starts to − 29 °C and steadystate tests of a direct-injection stratified-charge (DISC) engine operated on neat alcohols. 1987; SAE paper no. 872066.

106. Soccol R, Vandenberghe LPS, Costa B, et al. Brazilian biofuel programme: an overview. J of Sci Ind Res. 2005;64:897–904.

107. Specht M, Bandi A, Elser M, Staiss F. Comparison of CO2 sources for the synthesis of renewable methanol. In: Inui T, Anpo M, Izui K, Yanagida S, Yamaguchi T, eds. Advances in Chemical Conversion for Mitigating Carbon Dioxide. Amsterdam: Elsevier Science; 1998a;363–367. Studies in Surface Science vol. 114.

108. Specht M, Staiss F, Bandi A, Weimer T. Comparison of the renewable transport fuels, liquid hydrogen and methanol, with gasoline – energetic and economic aspects. Int J Hydrogen Energy. 1998b;23(5):387–396.

109. Specht M, Baumgart F, Feigl B, et al. Storing bioenergy and renewable electricity in the natural gas grid. In: FVEE – AEE Topics 2009: 69–78. Available from: http://www.solar-fuel.net/fileadmin/user_upload/Publikationen/Wind2SNG_ZSW_IWES_SolarFuel_FVEE.pdf; 2009; (accessed 28 February 2012).

110. Stansfield P, Bisordi A, OudeNijeweme D, Williams J, Gold M, Ali R. The performance of a modern vehicle on a variety of alcohol-gasoline fuel blends. SAE Int J Fuels Lubr. 2012;5(2):813–822.

111. Stein RA, Polovina D, Roth K, et al. Effect of heat of vaporization, chemical octane, and sensitivity on knock limit for ethanol–gasoline blends. SAE Int J Fuels Lubr. 2012;5(2):823–843.

112. Steinberg M. Production of synthetic methanol from air and water using controlled thermonuclear reactor power – I Technology and energy requirement. Energy Conversion. 1977;17:97–112.

113. Sterner M. Bioenergy and renewable power methane in integrated 100% renewable energy systems. University of Kassel 2009; Dr.-Ing Thesis, September.

114. Stucki S, Schuler A, Constantinescu M. Coupled CO2 recovery from the atmosphere and water electrolysis: feasibility of a new process for hydrogen storage. Int J Hydrogen Energy. 1995;20(8):653–663.

115. Thornton MJ. Impacts of biodiesel fuel blends oil dilution on light-duty diesel engine operation. SAE paper no. 2009-01-1790.

116. Turner JWG, Pearson RJ. The application of energy-based fuel formulae to increase the efficiency relevance and reduce the CO2 emissions of motor sport. SAE paper no. 2008-01-2953 In: SAE Motorsports Engineering Conference and Exposition, Concord, NC, December.

117. Turner JWG, Pearson RJ, Holland B, Peck R. Alcohol-based fuels in high performance engines. SAE paper number 2007–01-0056 In: SAE Fuels and Emissions Conference, Cape Town, South Africa, January. 2007a.

118. Turner JWG, Peck A, Pearson RJ. Flex-fuel vehicle development to promote synthetic alcohols as the basis for a potential negative-CO2 energy economy. 2007b; SAE paper no. 2007-01-3618.

119. Turner JWG, Pearson RJ, Purvis R, Dekker E, Johansson K, ac Bergström K. GEM ternary blends: removing the biomass limit by using iso-stoichiometric mixtures of gasoline, ethanol and methanol. In: The 10th International Conference on Engines and Vehicles, Capri, Naples, Italy, 11–16 September. 2011; SAE paper number 2011-24-0113.

120. Turner JWG, Pearson RJ, Dekker E, et al. Evolution of alcohol fuel blends towards a sustainable transport energy economy. In: 2012 MIT Energy Initiative Symposium: Prospects for Flexible- and Bi-Fuel Light Duty Vehicles, Cambridge, MA, 19 April. 2012a.

121. Turner JWG, Pearson RJ, McGregor MA, et al. GEM ternary blends: testing iso-stoichiometric mixtures of gasoline, ethanol and methanol in a production flex-fuel vehicle fitted with a physical alcohol sensor. SAE paper number 2012-01-1279 In: Detroit, MI: SAE 2012 World Congress; 2012b; 24–26 April.

122. Turner JWG, Pearson RJ, Bell A, De Goede S, Woolard C. GEM ternary blends of gasoline, ethanol and methanol: investigations into exhaust emissions, blend properties and octane numbers. In: SAE Powertrains, Fuels and Lubricants Meeting, Malmo, Sweden, 18–20 September. 2012c; SAE paper number 2012–01-1586.

123. Turner JWG, Pearson RJ, Dekker E, et al. Extending the role of alcohols as transport fuels using iso-stoichiometric ternary blends of gasoline, ethanol and methanol. Applied Energy. 2013;102:72–86.

124. UN. World agriculture: towards 2030/2050. Rome: FAO, United Nations; 2006; Available from: http://www.fao.org/fileadmin/user_upload/esag/docs/; (accessed 20 December 2011).

125. Urban CM, Timbario TJ, Bechtold RL. Performance and emissions of a DDC 8 V-71 engine fueled with cetane improved methanol. 1989; SAE paper no. 892064.

126. US Congress. Energy Independence and Security Act of 2007, Public Law 110-140, 110th Congress, DOCID: f:publ140.110.

127. US DoE. Energy efficiency and renewable energy. Alternative Fuels and Advanced Vehicles Data Center. Available from: http://www.afdc.energy.gov/afdc/data/fuels.html; 2012 (accessed 20 June 2012).

128. US DoE. ‘Biodiesel blends’, Alternative Fuels and Advanced Vehicles Data Center. Available from: http://www.afdc.energy.gov/fuels/biodiesel_blends.html; (accessed 20 June 2012).

129. Vancoille J, Verhelst S, Demuynck J, Galle J, Sileghem L, Van De Ginste M. Experimental evaluation of lean-burn and EGR as load control strategies for methanol engines. 2012; SAE paper no. 2012-01-1283.

130. von Helmolt R, Eberle U. Fuel cell vehicles: status 2007. J Power Sources. 165:833–843.

131. Wagner T, Wyszyński ML. Aldehydes and ketones in engine exhaust emissions – a review. Proc Instn Mech Engrs Journal of Automotive Engineering, Part D. 1996;210:109–122.

132. WBGU. World in transition: future bioenergy and sustainable land use: summary for policy makers. 2008; October. Available from: http://www.cbd.int/doc/biofuel/wbgu-bioenergy-SDM-en-20090603.pdf; (accessed 15 December 2011).

133. Weimer T, Schaber K, Specht M, Bandi A. Methanol from atmospheric carbon dioxide: a liquid zero emission fuel for the future. Energy Conversion and Management. 1996;37(6–8):1351–1356.

134. West BH, López AJ, Theiss TJ, Graves RL, Storey JM, Lewis SA. Fuel economy and emissions of the ethanol-optimized Saab 9-5 Biopower. In: SAE Powertrain & Fluid Systems Conference and Exhibition, Chicago, IL, October. SAE paper number 2007-01-3994.

135. White TL. Alcohol as a fuel for the automotive motor. 1907; SAE paper no. 070002.

136. Wuebben P, Unnasch S, Pellegrin V, Quigg D, Urban B. Transit bus operation with a DDC 6 V-92TAC engine operating on ignition-improved methanol. 1990; SAE paper no. 902161.

137. Zinoviev S, Muller-Langer F, Das P, et al. Next-generation biofuels: survey of emerging technologies and sustainability issues. ChemSusChem. 2010;3:1106–1133.

13.10 Appendix: List of abbreviations

AFR air:fuel ratio

ATDC after top-dead-centre

BXX blend of XX% by volume of biodiesel (FAME) in diesel

BEV battery electric vehicle

BMEP brake mean effective pressure

BOB blend-stock for oxygenate blending

BTL biomass-to-liquids

BuXX blend of XX% by volume of butanol in gasoline

CAFE corporate average fuel economy

CFR Co-operative Fuels Research

CNG compressed natural gas

CoV coefficient of variation

CTL coal-to-liquids

DI direct (fuel) injection

DME dimethyl ether

ECU electronic control unit

EGR exhaust gas recirculation

EMS engine management system

EREV extended-range electric vehicle

ETBE ethyl-tert butyl ether

EXX blend of XX% by volume of ethanol in gasoline

EU European Union

FAME fatty acid methyl ester

FFV flexible-fuel vehicle

GEM gasoline-ethanol-methanol

GHG greenhouse gas

GTL gas-to-liquids

HC hydrocarbon

HFCEV hydrogen fuel cell electric vehicle

HHV higher heating value

ICE internal combustion engine

ICEV internal combustion engine vehicle

IMEP indicated mean effective pressure

LCOE levelized cost of energy

LHV lower heating value

LNG liquid natural gas

LPG liquid petroleum gas

MBT minimum advance for best torque

MFB mass fraction burned

MON motor octane number

MTBE methyl-tert butyl ether

MTG methanol-to-gasoline

Mtoe million tonnes of oil equivalent

MXX blend of XX% by volume of methanol in gasoline

MY model year

NEDC New European Drive Cycle

NiMH nickel metal hydride

NMEP net mean effective pressure

NOx oxides of nitrogen

OPEC Organization of Petroleum Exporting Countries

PEM FC proton exchange membrane fuel cell

PI port (fuel) injection

PHEV plug-in hybrid electric vehicle

POME palm oil methyl ester

RFS renewable fuel standard

RME rapeseed methyl ester

RON research octane number

SI spark-ignition

SME soya bean methyl

TTW tank-to-wheel

US United States

WTT well-to-tank

WTW well-to-wheel

..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset