You are on page 1of 24

5.

16 Sustainable Aviation Fuels: Production, Use and Impact


on Decarbonization
Solange I Mussattoa, Ingrid Lopes Mottab, Rubens Maciel Filhob, Luuk van der Wielenc,d, Rafael Capazc,e, Joaquim Seabrae,
Patricia Osseweijerc, John Posadac, Marcelo de Freitas Gonçalvesf, Pedro Rodrigo Scorzag, and Giuliano Dragonea,
a
Department of Biotechnology and Biomedicine, Technical University of Denmark, Kongens Lyngby, Denmark; bLaboratory of
Optimization, Design and Advanced Control, School of Chemical Engineering, University of Campinas, Campinas, Brazil; cDepartment
of Biotechnology, Delft University of Technology, Delft, The Netherlands; dBernal Institute, University of Limerick, Limerick, Ireland;
e
Faculty of Mechanical Engineering, University of Campinas, Campinas, Brazil; fEmbraer S.A., São José dos Campos, Brazil; and
g
GOL Airlines, São Paulo, Brazil
r 2022 Elsevier Ltd. All rights reserved.

5.16.1 Current technologies for the production of aviation biofuels 348


5.16.1.1 Oil-to-jet (OtJ) 349
5.16.1.2 Alcohol-to-jet (AtJ) 349
5.16.1.3 Gas-to-jet (GtJ) 349
5.16.1.4 Sugar-to-jet (StJ) 350
5.16.1.5 Lignin-to-jet (LtJ) 350
5.16.2 Integration of thermochemical routes in biorefineries to produce aviation biofuels 350
5.16.2.1 Why and how to integrate thermochemical and biochemical plants 350
5.16.2.2 Thermochemical pathways to produce aviation biofuel with potential for integration in biorefineries 351
5.16.2.2.1 Gasification-based routes 351
5.16.2.2.1.1 BtL 352
5.16.2.2.1.2 Alcohol synthesis and AtJ 353
5.16.2.2.2 Pyrolysis-based routes 354
5.16.2.2.2.1 Catalytic fast pyrolysis 354
5.16.2.2.2.2 Bio-oil physical and chemical upgrading 354
5.16.2.2.3 HTL-based routes 355
5.16.2.3 On-going initiatives to integrate thermochemical and biochemical plants to produce aviation biofuels 356
5.16.3 Process integration, scale-up, and techno-economic aspects related to the production of aviation
biofuels 356
5.16.3.1 The role of integral biorefineries 358
5.16.4 Aviation biofuels and sustainability: The decarbonization challenge 358
5.16.4.1 SAF and the decarbonization of the aviation sector 358
5.16.4.2 Performance of aviation biofuels on GHG reduction emissions 359
5.16.4.2.1 GHG reduction accounting under different approaches 359
5.16.4.2.2 Land use change (LUC) 361
5.16.4.2.3 GHG reduction according to the reduction scheme for international aviation (CORSIA) 362
5.16.4.3 Other sustainability aspects of aviation biofuels beyond carbon footprint 362
5.16.4.4 GHG abatement costs: The link between economic and environmental performance 363
5.16.4.5 Social aspects of aviation biofuels 364
5.16.5 Prospects for a wider production and use of aviation biofuels: An industrial point of view 364
5.16.5.1 The commitment of international aviation industry with the climate change 364
5.16.5.1.1 Dependence on liquid fuels 365
5.16.5.2 Harmonization on jet fuel requirements 365
5.16.5.3 Approval process for new alternative fuels 365
5.16.5.4 New technologies, more feedstocks 367
5.16.5.5 SAF availability 367
5.16.6 Conclusion and vision ahead 368
Acknowledgments 368
References 368

348 Comprehensive Renewable Energy, Second Edition, Volume 5 doi:10.1016/B978-0-12-819727-1.00057-1


Sustainable Aviation Fuels: Production, Use and Impact on Decarbonization 349

5.16.1 Current technologies for the production of aviation biofuels

The aviation sector is responsible for approx. 12% of the carbon dioxide (CO2) emissions from all the transportation sector (Alves
et al., 2017). Therefore, the development of new alternative aviation fuels is considered the key nowadays to reduce the CO2
emissions, impacting not only the sustainability of the aviation sector, but of all the transportation sector. In this sense, aviation
companies are actively supporting projects that aim to develop a sustainable and economically competitive jet fuel substitute. The
target is to neutralize carbon emissions by 2030 and to reduce them by at least 50% in 2050 (Santos et al., 2018).
Several catalytic routes are currently available for the conversion of renewable materials into sustainable aviation fuels (SAF),
including oil-to-jet, alcohol-to-jet, gas-to-jet, and sugar-to-jet, among others. However, only a few of them (e.g., gas-to-jet and oil-
to-jet) have reached commercial/pre-commercial scale, while most, still remain in a relatively early stage of development (Wang
et al., 2020). A general overview of the main conversion technologies used for SAF production is given below.

5.16.1.1 Oil-to-jet (OtJ)


In the OtJ route, a wide range of oil-rich biomass can be converted into drop-in alternative jet fuels by different thermochemical
pathways: hydroprocessed esters and fatty acids (HEFA), catalytic hydrothermolysis and hydroprocessed depolymerized cellulosic
jet (Wei et al., 2019).
The production of HEFA fuel, also known as hydroprocessed renewable jet (HRJ), involves the hydrotreatment of triglycerides
from used cooking oils, animal fats or vegetable oils at high temperature and pressure with the aid of catalysts under excess
reacting gas (e.g., hydrogen) atmosphere. This process consists of an initial catalytic hydrogenation step, where the unsaturated
fatty acids and triglycerides are converted into saturated fatty acids, followed by deoxygenation reactions to remove oxygenated
compounds and transform the saturated fatty acids into strait chain paraffinic molecules (alkanes). These molecules are subse-
quently converted into the desired products, with carbon chains ranging from C8 to C16, by cracking and isomerization reactions.
A final step of distillation allows the separation of the mixture into paraffinic kerosene (HEFA SPK), green diesel, naphtha and
light gases (Why et al., 2019).
The catalytic hydrothermolysis (CH) or hydrothermal liquefaction (HTL) process turns triglycerides-based feedstocks (e.g., algal
oil, jatropha oil, camelina oil and tung oil) into naphtha, diesel fuel and jet fuel through a series of steps, including pretreatment
of the feedstock through conjugation, cyclization and cross-linking reactions, CH conversion at mild temperatures (200–400 1C)
and pressures (50–400 bar) in the presence of water, hydrogenation, catalytic decarboxylation and hydrotreating, and finally,
fractionation (Li et al., 2010a).
In the hydroprocessed depolymerized cellulosic jet (HDCJ) fuel production, bio-oil obtained from fast pyrolysis or from HTL
of lignocellulosic materials is upgraded to bio-jet fuel by catalytic hydrodeoxygenation (HDO), hydrogenation and further
fractionation (Xu et al., 2013). Hydrocarbon fuel produced by this emerging technology has not yet been approved by the
American Society for Testing and Materials (ASTM); however, it is expected to acquire a greater prominence in the near future, as
the costs of relevant equipment are decreasing (Cortez et al., 2015). In this sense, several pilot plants are currently under
construction or operational in a demonstration stage (Bauen and Nattrass, 2018).

5.16.1.2 Alcohol-to-jet (AtJ)


The AtJ route refers to the catalytic conversion process for kerosene production from various alcohols (e.g., methanol, ethanol or
butanol). These alcohols can be obtained from biogenic feedstocks (e.g., sugar, starch) via thermochemical and biochemical
(fermentative) processes, and serve as platform molecules for further manufacture of long-chain hydrocarbons (Pechstein et al.,
2018). The typical alcohol conversion route involves a series of upgrading reactions, including a first dehydration of the alcohol to
generate olefins, followed by their oligomerization in the presence of catalysts to produce longer chain hydrocarbons (middle
distillate) and later, hydrogenation of the middle distillate to obtain jet fuel-ranged hydrocarbons. The process is completed with a
final fractionation/distillation of the synthetic paraffin product (Geleynse et al., 2018). Catalytically converting alcohols to higher
hydrocarbon fractions suitable for transport applications is an established technology (e.g., methanol-to-gasoline process).
However, the conversion of ethanol and isobutanol is currently at the demonstration stage (Bauen and Nattrass, 2018).

5.16.1.3 Gas-to-jet (GtJ)


The GtJ route is based on the production of synthesis gas (syngas), a gaseous mixture of hydrogen (H2) and carbon monoxide
(CO), which is converted into liquid hydrocarbons. Different types of feedstocks, such as natural gas, coal or biomass, are first
gasified (or reformed) to produce syngas, and then converted to liquid synthetic paraffinic kerosene (SPK) through either the
Fischer-Tropsch (FT) catalytic process, or by gas fermentation (Wang and Tao, 2016). Depending on the feedstock used, the
processing pathway is also termed as “anything to liquid” (xtL), where “x” can be replaced by “B” (biomass), “C” (coal) or “G”
(gas) (Rye et al., 2010).
350 Sustainable Aviation Fuels: Production, Use and Impact on Decarbonization

The FT process is currently used on a commercial scale by companies such as Sasol, Shell, PetroSA, and Oryx GTL using fossil
feedstocks (natural gas and coal). On the other hand, the Fischer-Tropsch-Biomass-to-Liquid (FT-BtL) process is still at pilot and
demonstration stages, with first commercial plants in development (Bauen and Nattrass, 2018).
In the gas fermentation process, the syngas obtained via gasification is fermented via the reductive acetyl-coenzyme A (acetyl-
CoA) pathway (also known as Wood-Ljungdahl pathway) by acetogenic bacteria, producing a mixture of alcohols (e.g., ethanol
and 2,3-butanediol), which is then upgraded into jet fuel through the AtJ route described above (Pechstein et al., 2018). This
hybrid thermochemical/biochemical process has been presented as an alternative way to convert biomass-derived syngas into
liquid biofuels with high carbon efficiency, high yields, and at lower costs than the FT-BtL process (Griffin and Schultz, 2012).

5.16.1.4 Sugar-to-jet (StJ)


Sugars derived from renewable feedstocks, such as maize, sugarcane or lignocellulosic materials, can be converted into bio jet fuels
via the “Direct Sugar to Hydrocarbons” (DSHC) or the “Aqueous Phase Reforming” (APR) process (Wei et al., 2019).
In the DSHC process, genetically modified yeasts are used to convert the carbon source directly into synthetic isoparaffin (SIP)
fuel by fermentation, instead of producing an alcohol intermediate (AtJ route) (Rumizen, 2018). An example is given by the
company Amyris Inc., which has developed a technology to produce trans-b-farnesene by fermentation of sugars from Brazilian
sugarcane. b-farnesene is an unsaturated olefin with 15 carbon atoms (sesquiterpene), with many uses as a petroleum substitute
and specialty chemical (Benjamin et al., 2016).
The upgrading of water-soluble carbohydrates to non-oxygenated hydrocarbons fuels using the APR process can be achieved
through conventional catalytic condensation and hydrotreating techniques (Tompsett et al., 2011). In this process, different types
of biomass are initially processed to extract carbohydrate fractions and/or polysaccharides for further hydrolysis to generate C6 and
C5 sugars. These carbohydrates undergo hydrogenolysis and/or hydrogenation to produce short-chain oxygenated compounds
(e.g., glycerol, ethylene glycol) or sugar alcohols, respectively, which serve as chemical intermediates (platform molecules) for the
APR process. In the APR process, the oxygenated intermediates obtained in the previous hydrotreating step react with water at
moderate temperatures (170–300 1C) and pressures (10–90 bar) over a heterogeneous catalyst to form H2, CO2, alkanes, alde-
hydes and other oxygenates (ketones, furans, phenolics and organic acids) suitable for condensation to longer-chain hydro-
carbons. The products from APR can be finally converted to either gasoline-range hydrocarbons (e.g., isoalkanes and aromatics) by
acid condensation or jet fuel-range hydrocarbons by base condensation or by dehydration/oligomerization (Blommel and
Cortright, 2008).

5.16.1.5 Lignin-to-jet (LtJ)


Jet fuels range hydrocarbons can be produced from biomass-derived lignin through a series of steps, including pretreatment of the
feedstock for lignin extraction and purification, depolymerization for decomposition of lignin polymers into bio-oil compounds
with lower molecular weight, upgrading of bio-oils by HDO or zeolite cracking, alkylation and fractionation (Cheng and Brewer,
2017). Since lignin is the only biomass component rich in aromatic benzene ring structures, the LtJ route leads to the production
of low molecular weight alkyl-benzenes. These aromatics were identified as desirable molecules for the production of fully
synthetic jet fuels as they play an important role in the elastomeric swelling, material compatibility, and lubricity characteristics of
jet fuels (Hileman and Stratton, 2014).

5.16.2 Integration of thermochemical routes in biorefineries to produce aviation biofuels

5.16.2.1 Why and how to integrate thermochemical and biochemical plants


Thermochemical conversion processes have been studied since the late 17th century with wood and coal, but the development of
such routes using lignocellulosic biomass sources only grew in the 1970s with the outbreak of the first and second oil crises. Such
processes consist of the conversion of carbonaceous feedstocks into liquid or gaseous products for further production of electricity,
heat, chemicals or fuels, in the processes of combustion, gasification, pyrolysis, torrefaction, and HTL (Demirbas, 2009). Ther-
mochemical processes are characterized by high temperatures (300–1300 1C), catalysts that may be cheap and recyclable, high
reaction rates, short reaction times, no need for sterilization procedures as in biological processes, and high flexibility to feedstock
composition and morphology as thermochemical conversion is not sensitive to biomass recalcitrance (Zhang, 2016; Motta et al.,
2018). Despite such advantages, thermochemical plants are not cost-competitive with fossil-derived processes and most have not
achieved commercial technology readiness levels (TRLs). The integration of thermochemical plants with biorefineries may
attenuate this scenario, bringing advantages to both sites. In the integration, the existing biorefineries may act as host plants that
supply the raw materials, utilities, and infrastructure to the thermochemical plants. Also, liquid and gaseous streams may be
integrated (Stichnothe et al., 2016; Choudhary et al., 2020), and the thermochemical plants may contribute for the heat and water
integration grids of biorefineries (Martinez Hernandez and Ng, 2018).
As both biochemical and thermochemical plants use biomasses as raw materials, the integration of both pathways may
maximize the use of biomass components. The first strategy consists of both plants sharing the biomass supply. Examples include
Sustainable Aviation Fuels: Production, Use and Impact on Decarbonization 351

the supply of sugarcane bagasse for biochemical and thermochemical conversion plants to produce ethanol (Seabra et al., 2010),
corn stover to produce alcohols in an acetone-butanol-ethanol (ABE) system and thermochemical plant with gasification and
mixed alcohol synthesis plants (Martinez Hernandez and Ng, 2018), and sugarcane bagasse for an ABE system integrated with a
BtL plant (Furtado Júnior et al., 2020). Alternatively, the biorefineries may supply biomass residues to the thermochemical plants.
Sugarcane bagasse from 1G ethanol biorefineries may be sent to thermochemical plants for use as a feedstock (Bressanin et al.,
2020). As biorefineries mostly process the cellulose and hemicellulose of lignocellulosic biomass sources, the lignin part can be fed
to the thermochemical plants, which possibly is the only process applied at industrial scale for lignin conversion. Thermochemical
plants are interesting lignin processing options as they may convert any lignin source without pretreatment (Ahmad and Pant,
2018). Co-feeding is also possible, reducing the dependence upon seasonal biomass (such as agricultural crops and their residues)
and enabling the operation of the integrated plant throughout the whole year, as in the potential use of eucalyptus during
sugarcane off-season (Bressanin et al., 2020). As thermochemical processes are more flexible in terms of feedstock composition
than biochemical processes, the former can operate with difficult biomasses, further increasing biomass utilization. Challenges
concerning raw materials include feedstock and product purification, which account for 70% of the production costs, and the need
of large-scale production systems to take advantage of economies of scale (Martinez Hernandez and Ng, 2018).
The sharing of infrastructure may reduce the capital expenses, mitigate risks related to the implementation of new technologies,
and increase the utilization of the biorefinery installed capacity and the efficiency of the whole value chain (Seabra et al., 2010).
The shared infrastructure may include utilities (heat, steam, electricity), combined heat and power plants, water and wastewater
treatment processes, catalytic reactors, and distillation columns for hydrocarbon fractionation (Klein et al., 2018; Bressanin et al.,
2020). As the products of the thermochemical plants exit the reactors at high temperatures (300–1300 1C), there is an increased
potential for heat integration using pinch analysis for high-level tasks focusing on steam generation (e.g., using hot flue gases) and
low-level tasks focusing on hot water production. Water pinch analysis may be conducted to identify water recycling possibilities
in the integrated plant to save fresh water (Martinez Hernandez and Ng, 2018).
Liquid and gaseous streams may also be integrated in the thermochemical and biochemical plants. CO2 produced via anae-
robic digestion can be captured by thermochemical plants, either by feeding this stream to gasifiers and pyrolyzers or by sharing a
carbon capture and storage unit (Stichnothe et al., 2016). The thermochemical conversion areas may produce streams that are
useful for biorefineries (e.g., H2 and CH4 from biomass gasification for hydrogenation and steam-methane reforming processes,
respectively; intermediate products such as alcohols, etc.) (Parvez et al., 2020). Wastewater from processes such as pyrolysis and
HTL can also be treated in the biorefinery (Choudhary et al., 2020).
Finally, the integration of thermochemical and biochemical plants may lead to multiproduct generation (Seabra et al., 2010;
Haro et al., 2013), which increases market opportunities with a larger portfolio of products, including aviation fuels (Bressanin
et al., 2020).

5.16.2.2 Thermochemical pathways to produce aviation biofuel with potential for integration in biorefineries
Fig. 1 shows the main thermochemical routes that can be used to produce bio jet fuel and be integrated into existing biorefineries.
The following subsections give a more detailed description on such routes, and their potential synergies for integration in
biorefineries.

5.16.2.2.1 Gasification-based routes


The production of syngas and subsequent conversion into bio jet fuel is the core of gasification-based pathways and can be
achieved via two routes. The first process consists of directly converting a clean and conditioned syngas stream into a mixture of
hydrocarbons in a FT synthesis reactor in the so-called biomass-to-liquids (BtL) route. In the second approach, syngas is first
converted into alcohols catalytically or through fermentation, and then transformed into hydrocarbons. Both routes are described
hereafter as well as their possible integration with biorefineries.

Fig. 1 Examples of thermochemical routes that can be used to produce bio jet fuel and be integrated into biorefineries.
352 Sustainable Aviation Fuels: Production, Use and Impact on Decarbonization

Fig. 2 Main stages of BtL processes.

5.16.2.2.1.1 BtL
The BtL routes were developed in analogy to coal-to-liquids and gas-to-liquids processes, and their main steps (Neuling and
Kaltschmitt, 2018) are shown in Fig. 2. BtL fuels have several environmental advantages over fossil-derived fuels, such as lower
NOx emissions due to their high cetane number, lower production of particulates and aromatics due to their lower sulfur content,
and lower emissions of CO and hydrocarbons (Li et al., 2010b; Zhang, 2016). Also depending upon the feedstock, they may have
neutral or even negative carbon emission.
As shown in Fig. 2, dry biomass is fed to the gasifier and converted into syngas, tar, char, and impurities. Biomass can be
either in the form of pellets to ensure a uniform particle size and feeding (Motta et al., 2018) or pretreated via pyrolysis or
torrefaction to guarantee a higher energy density and improve feedstock logistics (Neuling and Kaltschmitt, 2018). As syngas
contains several impurities that may affect its conversion, it undergoes cleaning steps (tar, char, ash, N, Cl, and S species
removal) and conditioning steps (H2/CO ratio adjustment, CH4 and CO2 removal) (Dayton et al., 2011) and then goes to the
FT synthesis reactor. In FT synthesis, syngas is converted into a wide range of hydrocarbons (mainly n-paraffins and olefins) via
polymerization (reactions 1 and 2) following the Anderson-Schulz-Flory (ASF) statistical distribution, as well as the water-gas
shift (WGS) reaction 3, providing additional H2 for the FT reactions 1 and 2 (Bukur et al., 2016). Finally, the FT products are
upgraded and hydrocracked to convert gaseous products, naphtha, and wax into more hydrocarbons as per the bio jet fuel
specifications (Neves et al., 2020).
ð2n þ 1ÞH2 þ nCO-Cn H2nþ2 þ nH2 O ð1Þ

2H2 þ nCO-Cn H2n þ nH2 O ð2Þ

CO þ H2 O⇌CO2 þ H2 ð3Þ
The product yields and composition depend on several parameters such as temperature, pressure, residence time, reactor and
catalyst type, and H2/CO ratio, which should be carefully selected depending on the desired hydrocarbon fraction. To produce
light-weight hydrocarbons such as gasoline (C5-C12) and bio jet fuel (C12-C14), the high-temperature FT (HTFT) process is
usually selected, occurring at temperatures of 300–350 1C, pressures of 10–40 bar to avoid the formation of short-chain hydro-
carbons (Zhang, 2016), H2/CO ratio of 1.7:1, and using Fe catalysts supported on silica, alumina, or zeolites (Neves et al., 2020)
placed in multitubular fixed-bed or slurry-phase reactors (Dry, 2002).
Although the BtL process is the thermochemical route closest to a commercial TRL, it faces several challenges. First, biomass
availability and variability in composition and morphology affect a steady operation in terms of outputs and product com-
position. Second, the BtL route has a high complexity from a process engineering point of view. Finally, the syngas cleaning
stage has the challenge of removing several impurities and achieving strict syngas compositions with high energy efficiencies,
which is generally very expensive (Motta et al., 2018; Neuling and Kaltschmitt, 2018). Specifically, tar is one of the most
undesired impurities, which must be removed to avoid catalyst deactivation and clogging and corrosion of downstream
equipment (Neves et al., 2020). Although FT synthesis is a well-established process that runs commercially in coal-to-liquids
plants such as SASOL in South Africa (Chiaramonti et al., 2014), the need for high pressures increases costs and catalyst design
plays a strategic role to ensure selectivity and avoid catalyst deactivation in the presence of sulfur, nitrogen, and chlorine
compounds. The upgrading and hydrocracking stages have high H2 demands, which can be supplied by water electrolysis,
cryogenic separation, and pressure swing adsorption of part of the H2 in the syngas stream, tar thermal cracking, and tar steam
reforming (Zhang, 2016; Neves et al., 2020). In terms of quality, bio jet fuel produced via BtL routes have low lubricity and must
be blended with conventional fuels. The BtL plant has lower efficiencies than coal-to-liquids and gas-to-liquids plants, and the
bio jet fuel price is higher (Zhang, 2016).
Regarding the integration of BtL and biochemical plants, syngas produced in the BtL plant may provide H2 and CH4 for the
biorefineries, which can be used in hydrotreating and steam reforming processes, respectively. The syngas cleaning area has several
cooling and heating stages, which can be incorporated into the heat integration grid of the biorefinery and/or produce excess
steam. The FT synthesis reactor also produces steam as it is highly exothermic and requires cooling. FT synthesis also produces an
off-gas stream, which can be fed to the biorefinery combined heat and power plant. The BtL plant may also supply extra energy for
the heat-demanding processes of the biorefinery.
Sustainable Aviation Fuels: Production, Use and Impact on Decarbonization 353

5.16.2.2.1.2 Alcohol synthesis and AtJ


Alternatively, the syngas produced via biomass gasification can be converted into alcohols (methanol, ethanol, butanol, iso-
butanol, or a mixture) either catalytically or via fermentation, and transformed into bio jet fuel through the different stages of the
AtJ route.
The catalytic synthesis of methanol can be achieved by the reaction of CO and H2 (H2/CO ratio of 2:1 required, shown in
reaction 4), or H2 and CO2 (H2/CO2 ratio of 3:1 required, shown in reaction 5). In both cases, temperatures of 220–280 1C,
pressures of 50–100 bar, and the use of Cu and Zn-based catalysts in alumina supports lead to good selectivity and minimum
formation of side products. After the methanol synthesis, unconverted syngas and hydrocarbons formed by parallel reactions are
separated, followed by the cooling and liquefaction of the methanol stream, and distillation to increase methanol purity (Rauch
et al., 2018).
CO þ 2H2 ⇌CH3 OH ð4Þ

CO2 þ 3H2 ⇌CH3 OH þ 2H2 O ð5Þ


The catalytic production of higher linear alcohols (ethanol and butanol) is shown in reaction 6. Higher alcohol synthesis has a
complex mechanism, in which methanol is first produced and CO molecules are gradually added to form the higher alcohols
(C2-C4) (Rauch et al., 2018), as well as hydrocarbons and CO2 (Atsonios et al., 2015). This process operates above 40 bar and at a
250–320 1C temperature range using modified methanol synthesis catalysts (K, Zn, Cr, Cu-based), modified FT synthesis catalysts
(Fe, Ni, Co, Mo, Rh-based) (Atsonios et al., 2015), and alkali-doped sulfides (Rauch et al., 2018). This alcohol synthesis still
presents poor yield and selectivity of the catalysts, and difficulties in the recycling of the unconverted syngas (Rauch et al., 2018).
nCO þ 2nH2 -Cn H2nþ1 OH þ ðn  1ÞH2 O ð6Þ
Otherwise, alcohols can be produced via syngas fermentation, an anaerobic process in which acetogenic bacteria ferment CO
and H2 via the reductive acetyl-coenzyme A (acetyl-CoA) pathway, also known as the Wood-Ljung-dahl pathway. In this process,
acetyl-CoA is an intermediate that is converted into higher alcohols such as ethanol, propanol, and n-butanol. The syngas
composition and bacteria type affect the process efficiency and composition of end-products, and temperatures of 30–66 1C and
pH of 5.4–8.5 are reported as ideal for the metabolism of the bacteria. Syngas fermentation is still under development by
companies such as Swedish Biofuels and LanzaTech, which make efforts to overcome mass transfer limitations and improve yields,
efficiency, and product recovery in their pilot plants (Pechstein et al., 2018).
The produced alcohols go the AtJ section of the plant, which consists of the steps of dehydration, oligomerization, and
hydrogenation (Pechstein et al., 2018), as shown in Fig. 3.
In the dehydration step (see reaction 7), an oxygen atom is removed from the hydroxyl group in the alcohol molecule, forming
an alkene and a water molecule. Dehydration occurs in the presence of catalysts (metal oxides such as alumina, zeolites such as
ZSM-5, and Ru-based catalysts) at high pressures and temperatures of 170–540 1C (Atsonios et al., 2015; Pechstein et al., 2018).
Cn H2nþ1 OH-Cn H2n þ H2 O ð7Þ
The short-chain alkenes are converted into long-chain alkenes via oligomerization, which occurs in a stirred-tank reactor with
the use of a catalyst that favors oligomerization instead of cracking, dehydrogenation, and polymerization. The reaction conditions
depend on the catalyst and the raw material, and process conditions for either ethylene, n-butene, or isobutene oligomerization
are reported elsewhere (Atsonios et al., 2015; Pechstein et al., 2018). Reaction 8 shows an example of the oligomerization of
butene.
nC4 H8 -C4n H8n ; n ¼ 2; 3; 4 ð8Þ
As the double bonds in the long-chain alkenes are unstable and undesired for bio jet fuel, the alkenes go through hydro-
genation (reaction 9) to be converted into alkanes in the presence of metal catalysts based on Ni, Pt, or Pd at 200–350 1C and
pressures above 20 bar. Finally, the products are fractionated in a distillation column similar to those used in crude oil refineries,
recovering the light fractions on the top and the heavy fractions at the bottom (Atsonios et al., 2015; Pechstein et al., 2018).
Cn H2n þ H2 -C2n H2nþ2 ; n ¼ 8; 12; 16 ð9Þ
The commercial success of the above-described routes mainly depends on the technical development of the alcohol synthesis
pathways, as the AtJ process is a well-known technology. As both alcohol synthesis routes run with syngas as a feedstock, the
disadvantages related to biomass availability, gasifier operation complexity, and syngas cleaning exist. Also, H2 is needed in the
hydrogenation of alkenes and this usually impacts significantly the production costs.
In terms of integration in biorefineries, similarly to the BtL plant, the alcohol synthesis and AtJ route may provide H2, CH4,
excess steam, heat integration possibilities from syngas cooling, and extra energy to the biochemical plant. Specifically, the CO2

Fig. 3 Simplified diagram of the AtJ process.


354 Sustainable Aviation Fuels: Production, Use and Impact on Decarbonization

stream from the biorefinery can be used as a feed for alcohol synthesis in the case of catalytic methanol production and syngas
fermentation using H2 and CO2. Also, the biorefinery may provide ethanol for the AtJ plant.

5.16.2.2.2 Pyrolysis-based routes


Depending on the feedstock residence time and heating rate, pyrolysis results in different product yields and may be classified as
(a) slow pyrolysis (0.005 1C min1, residence times from minutes to days, higher char yields); (b) intermediate pyrolysis
(1–10 1C min1, residence times in minutes, intermediate char and bio-oil yields); (c) fast pyrolysis (5–100 1C min1, residence
times of 1–5 s, higher bio-oil yields); and (d) flash pyrolysis (up to 104 1C min1, residence times o1 s) (Zhang, 2016). Bio jet
fuel may be produced using bio-oil as an intermediate and, therefore, fast pyrolysis is selected for this purpose. However, fast
pyrolysis operation has limitations concerning scale-up, which increases production costs (Neuling and Kaltschmitt, 2018), and
stable bio-oil composition, which varies with feedstock, inorganic content, and operating conditions (Zhang, 2016). Also, bio-oil
has some properties that require upgrading steps (high oxygen and water content, low heating value, high viscosity, thermal and
chemical instability), and immiscibility with conventional liquid fuels, which usually have high H2 demands (Darda et al., 2019).
To overcome such challenges, R&D focuses on strategies to reduce bio-oil oxygen content, increase hydrocarbon production, and
decrease H2 requirements in bio-oil upgrading. The following sections describe the routes of catalytic fast pyrolysis, bio-oil
upgrading, and bio-oil coprocessing aiming at the production of bio jet fuel.

5.16.2.2.2.1 Catalytic fast pyrolysis


Catalytic fast pyrolysis (CFP) occurs at 500–600 1C and atmospheric pressure over catalysts, resulting in a bio-oil fraction of lower
oxygen levels, increased higher heating value, and higher hydrocarbon contents (mostly aromatics and olefins). The CFP catalyst
must be robust and the reactor design must allow catalyst regeneration. In addition, catalysts may be deactivated via coking and
condensation of polyaromatics. CFP is usually conducted over ZSM-5, an inexpensive catalyst known for deoxygenation reactions
and an increase of aromatic compounds, resulting in high organic liquid yields (B35 wt%) with minimal coke formation. Other
zeolites (such as SAPO-34, Y zeolite, beta zeolite, MCM-22, and MCM-44) and mesoporous silica (MCM-41) can also be
employed. Impregnation of Ni, Co, Fe, and Ga are known to increase the hydrocarbon yields. Bio-oil mixtures produced via CFP
may still require physical and chemical upgrading, which are described in the following subsection (Zhang, 2016; Darda et al.,
2019; He et al., 2019).

5.16.2.2.2.2 Bio-oil physical and chemical upgrading


Fig. 4 shows bio-oil stabilization and upgrading routes, which comprise physical and chemical upgrading steps. Crude bio-oil is
first fractionated into separate phases, which usually occurs in a scrubber. The fractionated bio-oil then goes through cyclones and
filters to remove its solids (char and sand). To avoid bio-oil polymerization and condensation (aging) during storage, antioxidants
(hydroxyanisole, butylated hydroxytoluene, tertiary butyl hydroquinone, propyl gallate or tert-butyl) are added. Finally, polar
solvents (methanol, ethanol, acetone, etc.) are added to reduce and stabilize the bio-oil viscosity for long-term storage (Thormann
and de Oro, 2018).
The chemical upgrading is based on bio-oil deoxygenation. Although HDO can significantly reduce the oxygen content of bio-
oil from 40–50 wt% to 1–28 wt% (Zhang, 2016), the high H2 consumption in HDO results in high operating costs and, therefore,
an intermediate deoxygenation step (either ketonization, aldol condensation, acid-catalyzed esterification, or thermal catalytic
reforming) is advised (Thormann and de Oro, 2018).

Fig. 4 Bio-oil stabilization and upgrading routes, based on (Thormann and de Oro, 2018).
Sustainable Aviation Fuels: Production, Use and Impact on Decarbonization 355

Intermediate deoxygenation processes consist of condensation reactions between the oxygen-polar groups of the bio-oil
compounds, generating a water molecule. In ketonization, two carboxylic acids react forming a C–C bond, which produces a
symmetric ketone of higher carbon length and eliminates a water molecule and CO2. Ketonization can be performed in either
vapor (around 250–300 1C) or liquid phases (lower temperatures in the presence of water) using weak acid catalysts such as TiO2
and ZrO2 possibly containing Ru, Au, and Fe. In aldol condensation, C–C bonds are formed by the reaction between two carbonyl
groups (either aldehydes or ketones), forming unsaturated ketones and water. Aldol condensation happens in vapor or liquid
phases at mild temperatures (50–120 1C), 10 bar, and using zeolitic catalysts with either acid, base, or acid-base bifunctional
properties. Acid-catalyzed esterification is an equilibrium reaction in which organic esters and alcohols react in mild conditions to
produce an ester and water using low-cost catalysts. In thermocatalytic reforming (TCR), biomass first undergoes intermediate
pyrolysis (350–450 1C, residence times of 5–10 min), and the pyrolysis vapors go to a catalytic reforming section (maximum
temperature of 750 1C) to be converted into hydrogen-rich syngas and condensable organic vapors/bio-oil of remarkable quality.
Among the intermediate deoxygenation routes, esterification is the only process at a pilot scale, while the others remain at the R&D
phase (Thormann and de Oro, 2018).
In the HDO process, H2 has two main roles: it reacts with the bio-oil oxygen-containing groups generating a water molecule
and oxygen-free polymerized compounds, and it saturates double bonds and aromatic rings. HDO happens at 300–600 1C and
70–200 bar in the presence of catalysts (Co–MoS2 and Ni–MoS2, transition metals supported on alumina or silica) (Thormann
and de Oro, 2018; He et al., 2019), which may be deactivated depending on the type of coke formation, catalyst acidity, and
process operating conditions. HDO is still between pilot and demonstration phases, requiring efforts to reach a commercial scale.
The hydrotreated bio-oil has excellent cold flow properties, a composition rich in aromatics and paraffins, and high heating
value. Thus, it can be blended to synthetic paraffines from FT synthesis or go through fractionation to obtain different hydrocarbon
fractions including bio jet fuel (Thormann and de Oro, 2018).
When integrated to a biorefinery, a pyrolysis-based bio jet fuel production plant may supply excess bio-oil for steam reforming
or direct use as a fuel in conventional equipment. Also, the CO2 streams from the biorefineries may be fed to pyrolyzers, as recent
works have demonstrated that CO2 fed to pyrolysis reactors increases the hydrocarbon contents in bio-oil due to different reactions
between biomass and CO2 rather than biomass and N2—only note that CO2 addition may increase the pyrolyzer energy demands
and add CO2 compression costs (Parvez et al., 2020).

5.16.2.2.3 HTL-based routes


HTL happens under supercritical conditions (pressures of 50–400 bar and temperatures of 200–400 1C), and in the absence of
catalysts or active organisms (Zhang, 2016; He et al., 2019) to convert wet feedstocks (lignocellulosic biomass, sewage sludge,
press cakes, animal manure, microalgae, slurries, etc.) mostly into a liquid mixture known as biocrude. Biocrude is a stable,
viscous, hydrophobic, heavy-oil-like material of high carbon content (50–75% of the original feedstock) containing hundreds of
organic compounds whose composition depends on the type and properties of the feedstock. Biocrude has oxygen and water
contents in the ranges of 7–20 wt% and 1–5 wt%, respectively, which are much lower than those of bio-oil (25–45 wt% and
20–40 wt%, respectively) (Kaltschmitt and Neuling, 2017). Fig. 5 shows the conceptual design of an HTL plant that produces bio
jet fuel.
Biomass or residues are first pretreated, which includes particle comminution and solvent addition to produce a pumpable
slurry that is pressurized and fed to the HTL reactor. After HTL, the product phases separate spontaneously, resulting in biocrude,
an aqueous phase, char, and a gas phase rich in CO2. The aqueous phase is recycled to the HTL reactor, the biochar may be used as
a fertilizer or as a load in polymeric materials, and the biocrude can be co-processed with crude oil or upgraded via hydrotreating
to produce a bio jet fuel stream under the desired specifications of acidity, heating value, and viscosity (Kaltschmitt and Neuling,
2017).
Although promising, the development of a continuous HTL operation is still in the R&D phase, and the different processing
steps of Fig. 5 have only been studied individually. Several technical challenges prevent HTL from achieving commercialization,

Fig. 5 Conceptual design of an HTL plant to produce bio jet fuel, based on (Kaltschmitt and Neuling, 2017).
356 Sustainable Aviation Fuels: Production, Use and Impact on Decarbonization

which are related to the high operating pressures, corrosive nature of biocrude, feedstock impurities (He et al., 2019), and biocrude
upgrading (Kaltschmitt and Neuling, 2017). HTL plants have high capital costs due to the expensive equipment required to ensure
safe operation under high pressures and endure the corrosive nature of biocrude (He et al., 2019). High pressures pose scaling up
challenges, especially considering the difficulties related to the high-pressure feeding step (Zhang, 2016). Although HTL accepts a
wide variety of feedstocks, impurities may lead to fouling, plugging, coke formation, and catalyst deactivation. Regarding biocrude
upgrading, it is still not clear whether the complete refining of biocrude towards bio jet fuel is a better alternative than its use as a
drop-in fuel in a refinery or biorefinery. Regardless, as biocrude is stabler than bio-oil, it could be transported to refineries or
biorefineries without stabilization. HTL has low yields of middle distillates over the gasoline carbon range, which may require the
blending of different types of feedstock to increase the share of middle distillates (Kaltschmitt and Neuling, 2017).
Ideally, part of the biocrude produced in an HTL-based process can be diverted to be steam-reformed or used as a fuel in
engines in the biorefinery. The HTL plant may supply extra energy for the heat-demanding equipment in the biorefinery.

5.16.2.3 On-going initiatives to integrate thermochemical and biochemical plants to produce aviation biofuels
Studies on the integration of thermochemical and biochemical plants to produce aviation biofuels currently focus on biomass
management and on the development of the concepts of stand-alone and centralized plants.
Biomass has several harvesting, handling, storage, and transportation problems due to its composition, density, and mor-
phology. The fact that biomass availability is usually seasonal imposes a difficult supply, which can be overcome with the co-
processing of different renewable feedstocks in thermochemical plants. Biomass has low mass and energy densities, which require
pretreatment techniques. Biomass of low mass densities may require pelletization and densification to avoid feeding issues,
especially in the case of high-pressure operations in gasifiers and HTL reactors. Drying, pyrolysis, and torrefaction are feasible
pretreatment options for low-energy-density biomass, improving the efficiency of thermochemical conversion reactors and
facilitating transportation and storage (Zhang, 2016; Motta et al., 2018).
The integration of thermochemical and biochemical plants to produce aviation biofuels must also have some flexibility in
terms of product portfolio as well as heat-power supply. An integrated thermo-biochemical plant can aim at producing not only
bio jet fuel, but electricity and other biofuels (ethanol, green diesel, green gasoline, etc.) (Zhang, 2016). Such an integrated
biorefinery may have centralized or stand-alone plants. In the centralized concept, the thermochemical area provides the heat,
steam, and power for the integrated units while, in the stand-alone concept, thermochemical and biochemical plants are inde-
pendently self-sufficient in terms of utilities (Neves et al., 2020).
Currently, two twin projects under the European Union Horizon 2020 initiative investigate different aspects concerning
biomass utilization and integration of thermochemical and biochemical technologies to obtain aviation fuels from renewable
feedstocks: the BECOOL (2017) and the BioValue (LNBR, 2020) projects.
The BECOOL project (launched in June 2017, ending in May 2022) is a consortium composed of 14 partners from universities,
research institutions, and large industries from seven European countries (European Commission, 2020a). The thermochemical
conversion investigations within the BECOOL project are held by the VTT Technical Research Centre of Finland, ECN-TNO, and
RE-CORD at the University of Florence. The studies focus on logistics, process operation, economics, and sustainability of
gasification and FT synthesis to produce Jet A-1 kerosene either using biomass or intermediate energy carriers obtained through
fast, intermediate, or oxidative pyrolysis. Intermediate energy carriers have several advantages: (i) less complex composition and
increased energy density than raw biomass; (ii) higher efficiency in gasification and improved syngas quality; (iii) possibility of
transport for longer distances, enabling distinct logistics scenarios. Among the preliminary routes, the thermochemical division of
BECOOL focuses on (i) decentralized fast-pyrolysis and bio-oil gasification in an entrained-flow gasifier; (ii) centralized biomass
gasification in an indirect Milena gasifier; and (iii) decentralized pyrolysis to bio-oil slurries and centralized gasification (BECOOL,
2017; Boymans et al., 2019).
The BioValue project (launched in June 2019, ending in June 2023) is an initiative that gathers the know-how and work of four
companies and 11 Brazilian research institutions (BECOOL, 2019), focusing on the production of aviation fuels from the
perspective of Brazilian biorefineries. The project concentrates on strategies to reuse abundant Brazilian resources (mostly
sugarcane residues, energy cane, and forest residues) as fuels for either stand-alone or centralized thermochemical and biochemical
plants, analyzing the impacts of the different aviation fuel production chain scenarios on the economics, sustainability, and
logistics. The studies are conducted from lab to semi-industrial scale, using pilot plants and laboratories of the several BioValue
partners, as well as the Virtual Sugarcane Biorefinery, a powerful simulation platform (LNBR, 2020).
BECOOL and BioValue benefit from the cooperation, as both partners identify the synergies of the European and Brazilian
experience in the conversion of biomass into advanced biofuels.

5.16.3 Process integration, scale-up, and techno-economic aspects related to the production of aviation
biofuels

Today's world does not have aviation fuel, but only manufacturing plant. Kerosene is produced within a range of oil refinery
products, utilizing the full crude oil intake. Product portfolios in a modern refinery context range from C2 (ethane, ethylene and
Sustainable Aviation Fuels: Production, Use and Impact on Decarbonization 357

ethyne) and C3 (propane and propenes) platforms and higher linear and branched alkanes/alkenes, the various (substituted)
aromatic platforms to larger molecules such as waxes and heavies that are purposed for asphalt and other heavy uses. This
portfolio of products and the extremely high efficiencies and integration of products and auxiliary resources as power, heat, steam
and process water, have led to industrial complexes such as in Singapore, Rotterdam and Houston with extremely high yields on
crude oil and very minimal wastes and emissions. Moreover, crude oil-based products have generally high energy contents
(expressed as heat of combustion), making them ideal as energy carriers, but due to their abundance, high quality and low costs
also as a material of choice for a wide range of daily life's products at affordable prices. The combination of products in such a
portfolio are the key to the tremendous economic success of the petrochemical industry.
A general and critical techno-economic evaluation of various production technology/feedstock combinations was presented by
de Jong et al. (2015). An important outcome was an overview of the estimated cost contributions of the Minimum Fuel Selling
Price (MFSP) of SAFs. Their work indicated that at the short term, none of the pathways assessed (HEFA, FT, HTL, pyrolysis, AtJ,
and DSHC) are able to reach price parity with petroleum-derived jet fuel. The pioneer plant analysis suggests that the HEFA
pathway is currently the best option, as the technology achieves the lowest MFSP of 29.3 €/GJ (1289 €/t) and is deployed on
commercial scale already. In the short term, HTL and pyrolysis emerge as promising alternatives, yielding MFSPs of 21.4 €/GJ (939
€/t) and 30.2 €/GJ (1326 €/t), respectively. The pioneer plant analysis shows considerable MFSP increases for producing drop-in
fuels using HTL and pyrolysis as both technologies are relatively immature. Hence, further R&D efforts into these pathways are
recommended. Co-production strategies for lower emissions and improving process economics were mentioned but not really
explored.
A roadmap exercise for the development of sustainable production of SAF and other energy carriers for long haul and heavy
transport was undertaken in the framework of AgroPolo, a FAPESP and BE-Basic joint initiative in the Brazilian State of São Paulo.
Here, the coproduction option was also mentioned as a potential successful strategy to lower emissions and improve economics
(Fig. 6). This is an important lesson for any renewable or SAF plant whether based on biomass (such as agricultural or forestry
residues) or specific energy crops, the newer schemes of synfuels based on C1 gases from steel and cement industry, and the myriad
of proposed approaches using renewable electricity and/or H2.
Taking biobased feedstocks with a gross carbon-hydrogen-oxygen composition –CH2O– as an example for the production of SAFs
and other energy carriers with a gross atomistic composition –CH2–, it is clear for stoichiometric considerations that half or more of
the feedstock mass will not end as a marketable product, but will be emitted as water or CO2 (depending on technology used). No
revenue is produced and in case of carbon capture and storage, it will lead to significant storage costs. Even without detailed
calculations, from a sustainability point of view, this is dramatic—transporting massive amounts of biomass to biorefinery processes
and only utilizing half or less. This is valid even in case of sustainably produced and certified biomass. This can be visualized in a
C–H–O compositional diagram as shown in Fig. 7. In this triangular diagram, the corner points represent respectively carbon,
hydrogen and oxygen, and the side molecules that are composed of the corresponding corner atoms only. Typical fossil resources as
crude oil with average composition CH2 and natural gas (CH4) are on the base side of the triangle. As indicated, their hydrogen
content is high and oxygen is almost absent, and their average energy content or heat of combustion is high. This is obviously

Fig. 6 Roadmap for techno-economic-societal innovations in industry sectors: aviation/marine biofuels.


358 Sustainable Aviation Fuels: Production, Use and Impact on Decarbonization

Fig. 7 C–H–O compositional diagram correlating mass composition for biobased and fossil feedstocks and products.

beneficial for fuels, but undesired for many of the chemical products such as the commodity polymers indicated polyethylene (PE),
polypropylene (PP), polystyrene (PS) and polyvinylchloride (PVC) which are collectively over 90% of the plastics market.
Biobased feedstocks such as biomass (as chips, pellets or syngas), sugars or derivative products such as ethanol, are placed centrally
in this compositional diagram (Fig. 7). Producing energy carriers such as SAFs, implies that part of the feedstock leads to products on the
base side of the diagram. For de-oxygenation of the biobased feedstock by (renewable) hydrogen or power, water is formed, which does
not contribute itself negatively to the footprint of the process, but does not contribute economically in a positive manner.
Coproduction of oxygen containing chemicals such as (di)carboxylic acids, alkylcarbonates, diols, hydroxyacids and derivatives
thereof or inorganic mono or bi carbonate materials (such as in biocement) are an attractive and potentially much more
economically sound strategy. Derivatives can be polymers such as PLA and other polyesters, polyethers, polyurethanes, as well as a
range of green solvents. Hence, large scale production of SAF as a long-term green strategy cannot be developed in isolation, but
needs disruptive changes at the materials and chemicals markets.

5.16.3.1 The role of integral biorefineries


From a techno-economic perspective, most advanced biofuels production technologies produce a range of fuel(s) and other
products. At the fuels-side, most technologies produce a portfolio for marine (“heavy” range), road/diesel (“mid” range), and
kerosene (“mid and top” range) applications. At the feedstock side, the oxygen-rich biomass and MSW-streams are roughly split in
energy dense fuels and oxygen rich (energy poor) material flows (due to “energy conservation law”). This sort of biorefinery
systems do exist (e.g., sugar and pulp and paper mills) but are often designed and optimized for one product, treating the rest as
waste. Therefore, the development of efficient biofuels production systems can only happen when full and integral redefinition of
this biorefinery-concept is taken into account. Several of these refinery-redefinition programs are being developed nowadays such
as the REDEFINERY program around the Port of Rotterdam (Netherlands), and the BIOFOREVER project (a consortium of
chemical and energy industries in Europe, also centered around the Port of Rotterdam).
A relevant analysis of the co-production of SAF and chemical coproducts was done by Alves et al. (2017). In this study, the authors
evaluated the use of different feedstocks (sugar crops, oil crops, and lignocellulosic biomass) for co-production of SAF and higher value-
added products in a biorefinery platform. The coproduction of renewable fuels and biochemicals in a biorefinery context was
demonstrated to be economically feasible. The main cost drivers of such a platform were the feedstock biomass cost, selling price of the
biochemicals, investment costs, and key process-inherent parameters such as conversion yields, obtained fermentable sugars, or oil
content in the feedstock. Sugarcane was the most promising feedstock for a biorefinery in the region of Minas Gerais, while soybean was
the most promising feedstock for a biorefinery in the region of Rio Grande do Sul, despite its higher uncertainty. In this study, succinic
acid was found to be the most promising chemical intermediate for biopolymers industry due to its relatively high market value, unique
market opportunity as a substitute of conventional petrochemicals, and forecast for future growth. But this result may be coincidental
due to specific assumptions, and a broader analysis should be made. AtJ and HEFA were the most attractive SAF production routes from
an economic point of view. Nevertheless, the risk analysis via Monte Carlo simulations indicated that scenarios for co-production of SAF
and biochemicals in Brazil present large uncertainties and high financial risk, mainly due to the use of second-generation feedstocks
(both in availability/logistics and technology (costs)), and the introduction of new technologies and products.

5.16.4 Aviation biofuels and sustainability: The decarbonization challenge

5.16.4.1 SAF and the decarbonization of the aviation sector


The aviation sector is responsible for around 3% of the global energy demand which represents 11% of the energy consumed by
the transportation sector (IEA, 2019), while emitting approximately 2.5% of global anthropogenic CO2 (IEA, 2020a). Despite
these modest contributions, the aviation industry features specific aspects:
Sustainable Aviation Fuels: Production, Use and Impact on Decarbonization 359

• It depends almost exclusively on fossil fuels, mostly fossil kerosene.


• The energy intensity of commercial aircrafts—on average values (1.8 MJ/passenger km)—is 3 times higher than buses and
railways, and it is similar to passenger cars, which already have consolidated initiatives for using biofuels (IEA, 2020a).
• The relevant growth rate of the global aviation sector (3.8% per year, 1973–2017) in terms of energy use, is near to road
transportation (4.2%) (IEA, 2019). Furthermore, following the increase of the commercial flights activity, even with the
improvements from operational and technical measures and new aircraft projects, aviation emissions have risen on average 2%
yearly since 2000 (Teter, 2020). The contribution for the total CO2 emissions could reach 3% of the total emissions in 2030
(Cortez et al., 2014), or even 6% by 2050 (Maniatis et al., 2011).

International flights lead to the greatest issues, since they have corresponded to around 60% of the fuel demanded in the
aviation sector (IEA, 2020b), 63% of the global operations in terms of Revenue Passenger Kilometer (RPK), and 70% of global
operations in terms of Revenue Tonne Kilometer (RTK), including passengers (ICAO, 2019a). However, different from domestic
aviation, international operations were not addressed by the intended nationally-determined climate (INDC’s) actions from the
Paris Agreement, which has driven the International Civil Aviation Organization (ICAO) to take the lead regarding to this issue
(Gonçalves, 2017).
ICAO has set forth some ambitious goals for decarbonizing international flights: (i) improve CO2 efficiency by an average of
1.5% per year from 2009 until 2020; (ii) achieve carbon-neutral growth by 2020; and (iii) reduce the carbon emissions by 50% in
2050 compared to 2005 levels (ICAO, 2010). To achieve these targets, several actions could be implemented such as operation/
infrastructure improvements, economic-based measures, and technology development. In general, operation actions comprise
more efficient flight procedures, baggage loading strategies, and weight reduction measures. Infrastructure improvements mean
implementing efficient measures for air traffic management and improving airport infrastructure. In turn, market-based measures
are related to carbon offsetting by emission units purchased in the carbon market. Finally, technological actions comprise changes
on aircraft designs, composite lightweight materials, advances in engine technology, and by partially replacing fossil fuels for
alternative fuels. The use of aviation biofuels plays an important role in this context due to the technical limitations of the other
decarbonization measures as well as the concerns related to the effective carbon offsetting from carbon credits trading.
Initiatives for expanding the use of aviation biofuels have already popped-up in several places. Since 2011, more than 250
thousand commercial flights have already operated with aviation biofuels, six airports worldwide have already regularly supplied
them (Aviation Benefits Beyond Borders, 2020), and a relevant scientific background has been built to support related themes.
Furthermore, the production and use of aviation biofuels can lead the aviation sector to a historic energy transition, which could
also comply with strategic sustainability issues.
According to the ICAO guidelines (ICAO, 2019f) and the current decarbonization goals, the sustainability of aviation biofuels
is currently evaluated by their performance on providing Greenhouse Gases (GHG) reduction, including those carbon emissions
from land use changes. Then, a Sustainable Aviation Fuel shall: (i) achieve a minimum 10% GHG reduction in comparison with
fossil kerosene on a life cycle basis; and (ii) not be made from biomass obtained from high carbon stocks after 1 January 2008,
such as primary forest, wetlands, or peatlands.
Despite the relevance of decarbonization actions, the production and use of aviation biofuels can lead the aviation sector to a
historic and sustainable energy transition. Then, this great challenge should be tackled in a broader perspective, which combines
environmental and socio-economic issues beyond GHG reductions, under different assessment methods as well. These issues are
briefly discussed in the following sections.

5.16.4.2 Performance of aviation biofuels on GHG reduction emissions


The potential GHG reduction of biofuels has been commonly estimated using the Life Cycle Assessment (LCA), where GHG
emissions along the whole biofuel life cycle (which can comprise feedstock procurement, biomass conversion, transportation, and
biofuels use) are inventoried and compiled. Different results are expected from different feedstocks, conversion technologies, and
supply-chains. However, the high sensitivity of the outcomes with respect to methodological choices such as system boundaries,
inventories assumptions, characterization factors, and co-products handling are well known, and it can lead to a wide range of
results for the same pathway.

5.16.4.2.1 GHG reduction accounting under different approaches


Some authors suggest that the way the systems are modeled in LCA should be strictly dependent on the questions that are
addressed (Guinée et al., 2011; Sonnemann and Vigon, 2011). In this sense, if the interest lays on attributing impacts to a specific
product based exclusively on its supply-chain flows, or on estimating the consequences by changing demand, the LCA can be
carried out through two different approaches: attributional LCA (ALCA) or consequential LCA (CLCA), respectively (US EPA,
2010; Miller and Keoleian, 2015; Weidema et al., 2018).
ALCA describes the production system using average data assuming a status-quo configuration. Furthermore, the environmental
burdens of the processes are allocated to the multiple products considering physical or economic relation among them, e.g., the
GHG emissions along the life cycle are allocated between products A and B with respect to mass, energy content, or the revenues
obtained from these products. On the other hand, CLCA focuses on the direct (and indirect) effects of a demand of a product or
service. The inventory should be comprised of marginal data, including the possible effects related to co-production, which is
360 Sustainable Aviation Fuels: Production, Use and Impact on Decarbonization

Fig. 8 Range of LCA results for aviation biofuels in comparison to fossil kerosene (Jet A, 89 gCO2e/MJ). Black dots present results from
attributional analysis with different allocation procedures. Green dots include consequential aspects. Red dots include land use change (LUC)
emissions. 2G alcohol: second generation ethanol or butanol. AtJ: Alcohol-to-Jet. CG/1G alcohol: first generation ethanol or butanol from corn
grains. DSCH: Direct Sugar-to-hydrocarbon Jet Fuel. FT: Fischer-Tropsch. HEFA: Hydroprocessed Esters and Fatty Acids. SC/1G alcohol: first
generation ethanol or butanol from sugarcane. Results are adapted from Capaz and Seabra (2016), including values from de Jong et al. (2017)
and Klein et al. (2018).

suitably handled by system expansion (Finnveden et al., 2009; Brander, 2017). In the latter case, displacement effects—e.g.,
avoided or induced GHG emissions—are also accounted for.
Fig. 8 complies results on LCA GHG emissions for aviation biofuel production from several studies reported in literature which
analyzed different technologies, various types of feedstocks and both LCA approaches, i.e., attributional and consequential. Those
results are directly compared to the LCA GHG emissions of fossil kerosene (Jet A), and all data presented consider the combustion
phase. According to this figure, most pathways lead to GHG emissions reduction compared to fossil kerosene (Jet A). In general,
results achieved through ALCA (black dots) range lower than when consequential aspects are included (green dots), or when
emissions from land use changes are accounted for (red dots).
Variations in results based on ALCA are due to process description (e.g., system boundaries and cut off criteria), background
data, and allocation procedures. The GHG emissions of aviation biofuel obtained from camelina through HEFA technology, which
has been the most studied conversion technology, could vary between 24.0 and 31.4 gCO2e/MJ if GHG emissions are allocated by
mass or by energy, respectively (Lokesh et al., 2015). On the other hand, the same pathway presented 46 gCO2e/MJ under a
different process description (Han et al., 2013), even when considering energy allocation. In turn, aviation biofuel from micro-
algae has been estimated within a wide range (29–476 gCO2e/MJ (Lokesh et al., 2015)), influenced by the growth conditions and
allocation methods. The GHG emissions for palm as feedstock, which ranged from 22.0 to 38.5 gCO2e/MJ under different
processing conditions (Stratton et al., 2010), could reach 58.5 gCO2e/MJ if the palm industrial effluents are not correctly treated.
For soybean-based pathway, in similar studies (Wong, 2008; Stratton et al., 2010), the GHG emissions values varied from 29.9 to
50.8 gCO2e/MJ assuming different agricultural and industrial yields, chemical inputs, and energy demand. On the other hand, the
values ranged from 29.7 to 41.5 gCO2e/MJ assuming mass and economic allocation for soybean system in Brazilian conditions,
respectively. Finally, for the AtJ-based pathways, the sugarcane-based production presented lower values (6.8–22 gCO2e/MJ) than
those obtained from corn grain (55–117.4 gCO2e/MJ) or from lignocellulosic materials (17.3–35.0 gCO2e/MJ) (Staples et al.,
2013; de Jong et al., 2017).
Sustainable Aviation Fuels: Production, Use and Impact on Decarbonization 361

Table 1 GHG emissions on life cycle basis for different aviation biofuels, under attributional approach and energy allocation.

Pathway Other references (without LUC) in (gCO2e/MJ) CORSIA default values

Core LCA in (gCO2e/MJ) iLUC in (gCO2e/MJ) Total in (gCO2e/MJ)

HEFA—Palm 22.5–59.8 37.4 39.1 76.5 (closed pond)


60.0 99.1 (open pond)
HEFA—Rapeseed 39.8–75.9 47.4 24.1 71.5
HEFA—Soybean 27.3–59.2 40.4 27.0 67.4
HEFA—Tallow 13.5–31.6 22.5 n.a. 22.5
HEFA—UCO 16.8–27.0 13.9 n.a. 13.9
FT—Residues 5.4–13.0 7.7 n.a. 7.7 (agricultural)
8.3 8.3 (forestry)
FT—Lignocellulose 4.4–26.0 10.4 (  32.9) (  22.5) (miscanthus)
10.4 (  3.8) 6.6 (switchgrass)
AtJ—SC/1G alcohol 22.2–35.0 24.0 7.3 31.3 (via butanol)
24.1 8.7 32.8 (via ethanol)
AtJ—CG/1G alcohol 55.0 55.8 22.1 77.9 (via butanol)
65.7 25.1 90.8 (via ethanol)
AtJ—2G alcohol 22.4–35.0 43.4 (  54.1) (  10.7) (via butanol, miscanthus)
29.3 n.a. 29.3 (via butanol, residues)
DSCH 44.0 32.8 11.3 44.1 (sugarcane)
32.4 20.2 52.6 (sugarbeet)

As a general trend, feedstocks production is usually the main contributor to GHG emissions (Souza et al., 2015; Capaz and
Seabra, 2016). However, as assumed by several regulatory schemes (European Commission, 2010; ANP, 2018; ICAO, 2019d),
residual feedstocks—such as agricultural trash, forestry residues, greases, or off-gases—are not burdened with any environmental
impact related to the upstream stages, except for harvesting operations and transportation. Due to this typical assumption,
residues-based pathways present, in general, lower values than food-based ones. The GHG emissions results for aviation biofuel
production from beef tallow and used cooking oil ranged in 13–31.6 gCO2e/MJ and 16.8–27.0 gCO2e/MJ, respectively. In turn,
thermochemical conversion of agro-forestry residues presented values between 6.0 and 13.6 gCO2e/MJ.
Regardless the type of feedstock (food-based or residues-based), it is worth mentioning that hydrogen demand for aviation
biofuels production corresponds to an important contribution to overall GHG emissions. It resulted in 25–80% of the GHG
emissions for fuels obtained from HEFA technology, and 15–23% for that produced from AtJ processes (de Jong et al., 2017; Capaz
et al., 2020), assuming hydrogen production from steam methane reform. On the other hand, ethanol production from lig-
nocellulosic material can correspond to around a half of the total emissions in systems based on AtJ technology (Capaz et al., 2020).
Relevant variations on results are observed when consequential aspects (see Fig. 8, green dots) are accounted for, such as the
avoided emissions related to co-products usage (through the application system expansion approach), which can even lead to
negative values. For instance, if the camelina meal obtained along with the extracted vegetable oil were used as a fuel or as a
substitute of soybean meal, the GHG emissions of aviation biofuel produced from camelina oil could be 60 gCO2e/MJ (Agus-
dinata et al., 2011) or  18 gCO2e/MJ (Fan et al., 2013), respectively. Likewise, if seedcake and husk were used as fertilizers, the
performance of aviation fuel from jatropha would be 40 gCO2e/MJ. On the other hand, if they were used for heat or power
generation, displacing the use of fuel oil or US grid electricity, the overall GHG emissions would be  134 gCO2e/MJ (Bailis and
Baka, 2010) and  45 gCO2e/MJ (Stratton et al., 2010), respectively.
Furthermore, induced effects, if captured, can change the overall performance of the system. The diversion of beef tallow from
its current use to produce aviation biofuel could induce the demand for soybean oil in a Brazilian context, leading to total GHG
emissions of 44 gCO2e/MJ, which differs from the value of 13 gCO2e/MJ initially estimated (i.e., when no effects are accounted
for). In turn, aviation biofuels from sugarcane molasses through the DSCH pathway would lead to 80 gCO2e/MJ, mostly due to
the effects of the incremental use of sorghum to replace the current use of molasses (Cox et al., 2014). This same pathway showed
a 8.5 gCO2e/MJ, without effects related to feedstock displacement, but including displacement of natural gas by power surplus
generation (Moreira et al., 2014).

5.16.4.2.2 Land use change (LUC)


Emissions related to LUC can lead to high variations on the LCA results. This has been the most contentious issue in evaluating
GHG effects of biofuels, especially the accounting of indirect effects (Souza et al., 2015). In a general definition (ISO, 2018), the
direct LUC effects (dLUC) comprise changes in the use or management of land within the system being assessed, while the indirect
LUC (iLUC) refers to a change in the use or management of land which is a consequence of direct land use change, but that occurs
outside the system being assessed. Differently from dLUC, iLUC cannot be directly measured or observed; instead, it is projected
with economic models, which are only able to capture both effects together.
362 Sustainable Aviation Fuels: Production, Use and Impact on Decarbonization

According to the values presented in Fig. 8 (see red dots), the aviation biofuel production from jatropha would present 13–141
gCO2e/MJ if pasture lands or shrublands were converted into jatropha crops (Bailis and Baka, 2010). Similarly, aviation fuels
obtained from soybean through the direct conversion of Cerrado fields (Brazilian savanna) and tropical rainforest to soybean crop
could lead to overall emissions of 80 and 775 gCO2e/MJ, respectively (Stratton et al., 2010). In turn, assuming the direct
conversion of peatland rainforests, palm-based biofuel could result in terrible performance of 648 gCO2e/MJ (Wong, 2008).
Finally, the performance of DSHC of sugarcane would be increased to 21 gCO2e/MJ (Moreira et al., 2014), if direct and indirect
effects of sugarcane expansion in Brazil were considered.

5.16.4.2.3 GHG reduction according to the reduction scheme for international aviation (CORSIA)
The ICAO goals have been managed by the Carbon Offsetting and CORSIA (ICAO, 2020). Similarly to other Low-Carbon Policies
—such as the Renewable Energy Directive (RED II) (European Commission, 2018) in Europe, the Renewable Fuel Standard (RFS)
(US EPA, 2010) in the United States, and Renovabio (Brazilian Government, 2017) in Brazil—CORSIA has specific guidelines
(ICAO, 2019d) for carbon accounting. In this case, the GHG emissions on life cycle basis are estimated using attributional
approach, covering the feedstock procurement up to biofuel use, with energy allocation for multiple products. Furthermore,
default iLUC values (ICAO, 2019b) must be considered to compose the total values (gCO2e/MJ), when necessary. Table 1
summarizes the values presented in Fig. 8 and some default values suggested by CORSIA.
For food-based pathways, the default values suggested by CORSIA for life cycle stages are within the values estimated in
literature. For these pathways, LUC emissions, which have been estimated from economic models like GTAP-BIO and GLOBIOM
(ICAO, 2019e), are more relevant to the total results. In case of aviation biofuels obtained from low-LUC risks areas (ICAO,
2019d), LUC emissions can be assumed to be zero, which would substantially increase the GHG reduction potential for food-
based pathways. Low-risk areas for land use changes are possible when feedstocks are produced with management practices that
provide an increase in the agricultural yield, without land expansion, or from unused lands with little risk for displacement of
other services, such as food, feed, and bioenergy. Furthermore, negative GHG emissions indicate carbon sequestration, and they
are mostly related to lignocellulosic feedstocks.
According to CORSIA, residues-based pathways are more beneficial, since the emissions related to upstream stage are dis-
regarded and they are not related to any land use effect. The default values are close to what has been reported in literature.

5.16.4.3 Other sustainability aspects of aviation biofuels beyond carbon footprint


As mentioned before, the potential GHG reduction from using aviation biofuels has been largely investigated. On the other hand,
possible trade-offs between climate change and other environmental impact categories remain rather unexplored. In this context,
an LCA study comparing different strategic production pathways in Brazil and analyzing eight environmental impact categories,
revealed that, while aviation biofuels provide lower global-scale impact than fossil kerosene, such as climate change and fossil
depletion, relevant trade-offs are observed in categories related to local impacts, such as eutrophication, toxicity and air quality-
related categories (Fig. 9).
Aviation biofuels from sugarcane (via AtJ technology) and soybean oil (via HEFA technology) have potential to provide GHG
emissions reduction of over 50% with respect to fossil kerosene, without considering land use effects. However, the sugarcane-
based pathway is related to greater impacts in terms of eutrophication (two-fold higher than fossil kerosene) and air quality (30%

Fig. 9 Environmental trade-offs of aviation biofuels normalized by the highest values in each impact category. CC, climate change; FD, fossil
depletion; TAC, terrestrial acidification; EUT, eutrophication; HTX, human toxicity; ETX, environmental toxicity; PMF, particulate material formation;
POF, photochemical oxidant formation. Adapted from Capaz RS, et al. (2020) Environmental trade-offs of renewable jet fuels in Brazil: Beyond the
carbon footprint. Science of the Total Environment 714: 136696, https://doi.org/10.1016/j.scitotenv.2020.136696.
Sustainable Aviation Fuels: Production, Use and Impact on Decarbonization 363

Table 2 Minimum selling price (MSP), life cycle carbon emissions, and related abatement costs for aviation biofuels via multiple technologies.

Pathway MSPa (USD/GJ) References GHG emissions (kgCO2e/GJ)b Abatement costsc (USD/tCO2e)

Fossil kerosene 10.8–14.0 US EIA (2020) 89.0 n.a.


HEFA—Soybean 23.1 Klein et al. (2018) 67.9–67.4 442–1313
37.2 Bann et al. (2017)
HEFA—Palm 18.4 Klein et al. (2018) 76.5–99.1 383–661d
HEFA—UCO 28.4 Bann et al. (2017) 13.9 194–304
33.3 de Jong et al. (2015)
HEFA—Tallow 33.1 Bann et al. (2017) 22.5 292–340
AtJ—SC/1G alcohol 27.2 Klein et al. (2018) 32.8 239–743
44.9 Santos et al. (2018)
51.8 Diederichs et al. (2016)
AtJ—2G alcohol (lignocellulosic residues) 36.0 Klein et al. (2018) 23.8–29.3 343–1158
55.5 de Jong et al. (2015)
64.0–67.7 Santos et al. (2018)
78.8 Diederichs et al. (2016)
FT—residues  6.9–11.2 Klein et al. (2018) 7.7–8.3 (  35)–567e
46.6 de Jong et al. (2015)
56.0 Diederichs et al. (2016)
a
Minimum Selling Price (MSP). When necessary, the MSP were converted in USD/GJ taking the exchange rate, density and heating value assumed in the original reference. It was
assumed 32.0 GJ/m3 and 0.735 t/m3, as LHV and density of aviation biofuel (ANP, 2020), respectively, when these data are not available in the original reference. For fossil kerosene,
it was assumed the FOB price in 2017–19 for US Gulf Coast.
b
Default values according to CORSIA (ICAO, 2019d).
c
It was estimated diving the difference between the fuels price, i.e., “ “ by the GHG emissions reduction, i.e., “GHGfossil_kerosene  GHGaviation_biofuel“.
For fossil kerosene, it was considered the FOB price range in 2017–19 for US Gulf Coast. The abatement costs present negative values, when the MSP of aviation biofuels is lower than
fossil kerosene price.
d
The default value for palm oil production with open ponds (99.1 gCO2e/MJ) as not considered here.
e
The negative MSP of aviation biofuel (  6.9 USD/GJ) was not considered for abatement costs calculation.

higher), mostly because of fertilizers use and bagasse burning at the ethanol mill. Furthermore, the soybean-based pathway could
cause larger impacts on human and environmental toxicity (five-fold higher than those for fossil kerosene), because of agro-
chemical applications. On the other hand, the GHG emissions reductions are estimated to be around 70% in the residues-based
pathways since the environmental burden of the upstream stages are not considered. In the latter case, no relevant trade-offs are
observed, except on air quality impacts for hydrolysis-based pathways with wood and sugarcane residues, due to biomass burning
at the ethanol mill.
A few other studies have carried out similar analysis. According to Klein et al. (2018), who analyzed different strategies to produce
aviation biofuels integrated with sugarcane mills in Brazil, the pathway based on sugarcane ethanol presented similar trends to the
previous study, i.e., higher values for terrestrial acidification (two-fold higher) and for human toxicity (27% higher) than fossil
kerosene, while it provides lower fossil depletion (85%). In turn, Cavalett and Cherubini (2018) investigated the production of
aviation biofuels from forest residues in Norway. They recommended FT as the most interesting option in terms of environmental
performance. According to them, this pathway presented lower terrestrial acidification (24%), particulate matter formation (11%),
and photochemical oxidant formation (6%) than fossil kerosene. For these same categories, those authors reported that aviation
biofuels from lignocellulosic ethanol provided greater impact relative to fossil kerosene, such as terrestrial acidification (13% higher),
particulate material (34%), and photochemical oxidant formation (30%). The description of the whole supply chain in Norway,
which included field and industrial operations, and transport, can explain the differences between the studies.

5.16.4.4 GHG abatement costs: The link between economic and environmental performance
While several studies have confirmed the potential GHG reduction from the use of aviation biofuels, the vast majority of them are
not yet economically competitive in comparison to fossil kerosene (Table 2), which has presented an average free on board (FOB)
price of 12.6 USD/GJ in 2017–19 (US EIA, 2020). Some trends are observed between the studies, such as the low minimum selling
price (MSP) related to aviation biofuels obtained from used cooking oil, and the high values for biofuels produced from
lignocellulosic residues via 2G alcohols. The low values reported by Klein et al. (2018) highlight the benefits of considering
integrated aviation biofuels production along with an ethanol plant (for instance, the MSP of aviation biofuel obtained via FT
technology could even reach negative values due to the great profit from the ethanol revenue). On the other hand, the authors
pointed out the complexity of the integration of this latter technology, due to the high requirement for integrating mass and
energy flows.
Even so, the cheapest route is not always the one that provides the most significant carbon reduction and vice versa. Thus,
economic efforts to develop this or that route may or may not be worth it from the perspective of ICAO goals. In this context, the
“abatement costs” (i.e., how much is the carbon reduced by a specific technology) can raise some intriguing issues.
364 Sustainable Aviation Fuels: Production, Use and Impact on Decarbonization

Although it is appropriate that the economic feasibility analysis and LCA are carried out based on the same supply-chain
description, it was possible to indicate some trends, according to the values estimated in Table 2. Residues-based pathways can
present low abatement costs, except for those based on ethanol produced from lignocellulosic material. In turn, the costs related to
oil-bearing plants are mostly influenced by LUC emissions, and can decrease by more than 50% if the feedstock are produced on
low-iLUC risks lands. Thermochemical conversion of agro-forestry residues can even present negative values. Integrated produc-
tion designs, as proposed by Klein et al. (2018) can lead to substantial decrease in the total costs.
Currently, CORSIA allows for GHG reductions by purchasing emission units in the carbon market. Furthermore, the current
prices of the emissions units (1.02–3.13 USD/tCO2e) (Hamrick and Gallant, 2018; Donofrio et al., 2019), or even future ones
(5.90–55.2 USD/tCO2e) (Piris-Cabezas et al., 2018), are much lower than the abatement costs estimated through aviation biofuels
production (Pavlenko et al., 2019). On the other hand, several concerns about the credibility of the emissions units and their
effective mitigation effects indicate that aviation biofuels could play a fundamental role in aviation sector goals. In fact, only
aviation biofuels, in the current circumstances and considering specific conditions of production, could lead the aviation sector to
an effective energy transition associated with direct GHG reductions. However, a new sector of fuels with all technology obstacles
and current financial limits does not appear overnight and should be supported by robust policies. In the case of the current
CORSIA guidelines, which treat carbon offset and carbon reduction equally, any effort to support initiatives for the production of
aviation biofuels may be discouraged or postponed since the carbon reduction targets can be achieved by airlines in the short-term
by purchasing emission units that are much cheaper than abatement costs of aviation biofuels.

5.16.4.5 Social aspects of aviation biofuels


Finally, studies about social aspects of biofuels production could be assumed to be valid for aviation biofuels. Generally, in
bioenergy systems, the social impacts are more relevant in the agricultural stage, involving labor conditions, labor rights, food
competition and others. Therefore, employment generation and salaries along the supply chain must be considered too (Brinkman
et al., 2018; Cardoso et al., 2018).
Specifically for the supply-chain of aviation biofuels, a very few studies have been developed. Recently, a study examined the
major macro socio-economic aspects (such as employment, gross domestic product (GDP) generation, and trade balance) to
supply the Brazilian kerosene demand in 2050 under different scenarios of biofuels promotion (Wang et al., 2019). Besides the
new supply-chains have showed large positive net effects on employment and GDP, the aviation biofuel obtained from macauba
oil presented the best results. The authors estimated that the production of aviation biofuel (assuming a coverage of 3%–15% of
the expected Brazilian demand of aviation fuel by 2050), could create around 12,000–65,000 jobs, while contributing US
$200–1100 million to Brazil's GDP, also with an increase in total imports in the range of US$80–250 million. These potential
benefits should also be considered when discussing broader aspects of sustainability for aviation biofuel production.

5.16.5 Prospects for a wider production and use of aviation biofuels: An industrial point of view

5.16.5.1 The commitment of international aviation industry with the climate change
ICAO, a United Nations’ body, has highlighted that the international aviation emissions is projected to increase due to the
expected continuous growth of the civil aviation (commercial and business aviation) (ICAO, 2016). ICAO set a strong com-
mitment to attend the aspirational international goals of the Paris Agreement. Aware of its responsibility with climate change and
aligned to support ICAO's position in this theme, a coalition was created by the aviation industry to support the GHG reduction
setting an important goal of reducing CO2 emissions by 50% by 2050 having 2005 levels as baseline. In order to achieve this goal,
this coalition defined a four-pillar strategy based in technology, operations, infrastructure and a global market-based measure.
Some actions over these pillars can be understood as:

• Improvements in technology related to the aircraft design (better aerodynamic and lightweight materials, for example) and
engine technologies like, high bypass ratio engines, which can collaborate to improve the aircraft fuel efficiency.
• The operational pillar focus on solutions that can save weights in aircraft interior such as lighter-weight seats and carts, but in
the adoption of operational procedures that save jet fuel including, but not limited to single engine taxing, fuel planning,
among others.
• Infrastructure's strategy is related to air traffic management (ATM) investments and trajectory optimization, which reduces the
flight time and, consequently, reducing the fuel consumption (e.g., continuous descent into airports, procedures to avoid
airborne holding).
• The milestone for the global market-based measures was the adoption of the CORSIA by ICAO, in 2016. CORSIA will be
implemented in three phases, preceded by a baseline period, 2019, for the CO2 emissions. The pilot phase from 2021 to 2023
and first phase from 2024 to 2026 are voluntaries. At the second phase from 2027 to 2035, it will be applied by Member States
with an individual share of international aviation activity in year 2018 above 0.5% of total activity or States which cumulative
share has reached 90% of total activity (ICAO, 2018).
Sustainable Aviation Fuels: Production, Use and Impact on Decarbonization 365

The expected growth of CO2 emission by the international aviation and how the contributions of all measures previously
mentioned can participate to mitigate the emission through the years, have been recently reported by ICAO (2019c). According to
them, the overall benefits of the actions based in the four pillars for the climate change is not only related to the CO2 emission, but
also for the non-CO2 emissions, such as contrails, soot, nitrogen oxides, water vapor and induced cloudiness.

5.16.5.1.1 Dependence on liquid fuels


The aviation industry has been using aviation jet fuel in turbine engines for decades. It is an energy-intensive and liquid-dependent
industry. Unlike other transportation modes, most of the original equipment manufacturers (OEM) for aircraft and aeronautical
systems continue to believe in the use of liquid fuel for some decades (Hileman et al., 2013). Full electrification is still far from
being commercially feasible, mainly to be available for medium to long haul flights. Batteries need to reach very high levels of
energy density while need to reduce their weight. For a comparison, jet fuel has a minimum energy density of 42.8 MJ/kg, while
some batteries used in electric vehicles provide an energy density of around 0.72 MJ/kg (approximately 200 Wh/kg). Certifying a
fully-electric commercial or business aircraft will be a severe process, mainly considering battery characteristics as thermal runway,
charging and discharging (NAE, 2020).
The use of hydrogen, for fuel cell propulsion, e.g., can reduce the climate impacts up to 90% (McKinsey and Company, 2020).
However, the use of hydrogen, in many ways, will demand the development of new turbines, lighter tanks and needs to create a
new distribution infrastructure in the airport.
Another point that supports the liquid fuel priority is that an aircraft, after leaving the production line, from its entry-in-service
date, will be operational for the next two or three decades (respecting all maintenance tasks required).
The search for an alternative fuel has motivated not only the aviation industry, but also energy governmental sectors, research
institutions and oil companies around the world. A new step on alternative fuels evolution happened in 2008 with a flight of a
Boeing B747 operating one of the engines with a blend of jet fuel and a biofuel made from babassu and coconut oil (Greenair,
2008). Initiatives like this improved the learning curve on alternative fuels towards to the development of new low carbon fuel
chain to attain the goals of the industry on climate change. An alternative fuel must be drop-in, it means, be indistinguishable
from the fossil fuel, causing no impact or changes in the aircraft, engines or at fuel distribution infrastructure.

5.16.5.2 Harmonization on jet fuel requirements


Ensuring safety and performance has driven the updates on jet fuel standards around the globe. Any change in the specification
needs to have its impact evaluated either in the aircraft or the supply system (which includes refining process, control quality and
delivery). Worldwide, there are two standards that have served as reference for many countries and their regulatory Agencies: one
issued by the American ASTM International designated ASTM D1655 “Standard Specification for Aviation Turbine Fuels” (ASTM,
2020a), and the Defense Standard (DefStan) 91-091 “Turbine Fuel, Kerosene Type, Jet A-1; NATO Code: F-35; Joint Service
Designation: AVTUR” issued by United kingdom's Ministry of Defense (UK Ministry of Defence, 2019).
Aviation turbine fuels under the requirements of ASTM D1655 and DefStan 91-091 are made from feedstock like petroleum,
oil sand or shale derived feedstocks in refineries. Synthesized hydrocarbons produced from these sources have been used and
historically presented similar properties. There is a provision in both mentioned standards for certain synthetic components from
unconventional sources. Some specific feedstocks used for co-processing are allowed being limited to a blend of 5% by volume in
the final product (with the fossil jet fuel).

5.16.5.3 Approval process for new alternative fuels


In the event of a new technology based on feedstocks not listed in the aforementioned standards be considered to produce an
alternative jet fuel, this fuel needs to attend the requirements of ASTM D7566 “Standard Specification for Aviation Turbine Fuel
Containing Synthesized Hydrocarbons”, standard where this new technology will be included. The edition of ASTM D7566,
approved on May 1, 2020 lists seven approved technologies, as described in Table 3 (ASTM, 2020b).

Table 3 ASTM D7566 approved technologies.

ASTM D7566—Approved Technologies Listed by Annex Number Acronym % maximum Approval


used blending year

A1. Fischer-Tropsch hydroprocessed synthesized paraffinic kerosene FT-SPK 50 2009


A2. Synthesized paraffinic kerosene from hydroprocessed esters and fatty acids HEFA-SPK 50 2011
A3. Synthesized isoparaffins from hydroprocessed fermented sugars SIP-SPK 10 2014
A4. Synthesized kerosene with aromatics derived by alkylation of light aromatics from nonpetroleum FT-SPK/A 50 2015
sources
A5. Alcohol-to-jet synthetic paraffinic kerosene AtJ-SPK 50 2016
A6. Synthesized kerosene from hydrothermal conversion of fatty acid esters and fatty acids CHJ-SPK 50 2020
A7. Synthesized paraffinic kerosene from hydroprocessed hydrocarbons, esters and fatty acids HC-HEFA 10 2020
366 Sustainable Aviation Fuels: Production, Use and Impact on Decarbonization

Tier 1 Tier 2 Tier 3 Tier 4

Fit-For- Phase 1 OEM Review Component/


Specification ASTM Engine
Purpose & Tier 3 & 4 Rig/APU
Process Research Testing
Properties Requirements Testing
Report

Accepted ASTM
Review &
Ballot Phase 2
OEM
Rejected Re-Eval FAA ASTM
Review &
as required Review Research
Approval
ASTM Report
Specification
Included at
ASTM ASTM Balloting
Specification Process

Fig. 10 ASTM D4054 Test Program (based on ASTM, 2019).

Fig. 11 ASTM D4054 Fast Track Program (based on ASTM, 2019).

In parallel of the latest ASTM D7566’s approved technologies, an equivalent number of new technology candidates are at ASTM
pipeline seeking for approval. The forum for proposal, discussions and evaluation of these new technologies happens at the ASTM
International subcommittee D02.J0 “Aviation Fuels”. It is required that sponsors of a new candidate fuel follow the test program
indicated at ASTM D4054 “Standard Practice for Evaluation of New Aviation Turbine Fuels and Fuel Additives”. This Practice
defines the requisites for testing and information about the volume required for evaluation. The test program consists of four tiers:
1—Fuel Specification Properties; 2—Fit-for-Purpose Properties; 3—Component and Rig Tests and 4—Engine Tests. Fig. 10
demonstrates the workflow of the Test Program, including the OEM and FAA reviews and the consequent balloting to approval, or
not, the candidate technology as a new ASTM specification (ASTM, 2019).
Currently, the whole ASTM D4054 test program is highly expensive, long and requests large number of interactions. The level
of investment required for the sponsor who seeks the approval is high, from 5–15 millions of dollars, being the sponsor
responsible to supply almost 1 million liters of the neat candidate alternative fuel (US DOE, 2020). There is no limit of time to
obtain the approval; it can take years. More recently, motivated to turn this process more expedite, a fast-track OEM qualification
and approval process was included in D4054, but under some limitations. The most important limitation is related to the
maximum blend ratio for neat alternative fuel of 10% (in volume) in the final blending. The sponsor needs to provide com-
positional data of the neat alternative jet fuel components and information on the maturity of conversion process used. Due to
expected similarity in terms of composition and/or process with the fossil jet fuel, the Tiers 3 and 4 are not required (a very
expensive part of the program) and only a Tier 1 “Plus” is performed (Fig. 11).
Seeking ways to make the approval process fast and reliable, FAA, the US Federal Aviation Administration, through its Aviation
Sustainability Center (named ASCENT) created the Clearing House. Under the leadership of University of Dayton Research
Institute (UDRI), the Clearing House is an initiative to manage all demands for alternative fuel test program data that a candidate
needs to provide to seek for approval at ASTM. Candidates can be part of the Clearing House to perform the complete ASTM
D4054 program or through the fast track one.
Sustainable Aviation Fuels: Production, Use and Impact on Decarbonization 367

5.16.5.4 New technologies, more feedstocks


The current approved technologies at ASTM D7566 can process a large range of feedstocks, such as agricultural residues, forest
residues, woody crops, municipal solid waste, tallow, used cooking oil, vegetable oils, sugars and waste gases. For the aviation
industry it is relevant that the feedstocks used to produce alternative jet fuel meet the requirements of CORSIA, to be elected as
CORSIA Eligible Fuel (ICAO, 2018) complying with all sustainability criteria required.
Feedstock has been a significant part of the final cost for alternative jet fuel. In some cases, like the HEFA technologies, some
feedstock contribution reaches up to 76% (Seber et al., 2014). In order to provide an attractive cost and price-parity with fossil jet
fuel, technology providers are seeking for process able to use residues, solid wastes and waste gases due to their reduced cost and
potential environmental benefits.

5.16.5.5 SAF availability


A relevant information highlighted by the European initiative ReFuelEU Aviation is that the production of alternative jet fuel
represents only 0.05% of the total jet fuel consumption (European Commission, 2020c). As of 5 July 2020, 88 States Members of
ICAO, which represent 77% of international aviation activity, have decided to participate in the first phase of CORSIA, but only a
few countries have SAF industrial production; most of them in North America and Europe (ATAG, 2020) as SAF is still more
expensive than jet fuel.
Today, there is a very small production of SAF in the world and, even the predicted expansions, do not represent a considerable
volume of production compared to world consumption (Table 4). Except for some companies that develop technologies that
directly produce 100% SAF (or do not declare their co-products), it is observed that SAF production is always associated with the
production of green diesel, normally at a proportion of around 20% of the volume of this diesel, associated with the technological
options of each company.
The question that crosswise affects airlines is “how to make the adoption of SAF with final costs substantially higher than fossil
fuels without disrupting financial balance and competitiveness?”
Air transport is a business very sensible to the jet fuel (and oil) prices. According to IATA, in 2019 the expenses with jet fuel
reached 23.7% of operational expenses for their associate operators, based on an average oil barrel price of US$ 65.0 (IATA, 2019).
There is not a single solution to this situation; different fronts need to be attacked in an orchestrated way. Country policies plays an
important role to foster SAF chain, creating policy measures for R&D or low blending mandates, for example. A joint action
including airlines, oil companies and technology providers reaching a commitment to buy SAF can motivate financial institutions
to provide enough capital building SAF facilities.

Table 4 Declared or intended SAF production around the world.

Company Country Green diesel SAF

Actual Planned Actual Planned

Americas
Diamond Green Diesel United States 1.000 1.000 400
ECB Group Paraguay 1000 200
Fulcrum United States 55
GEVO United States 1
LanzaTech United States 1 30
Mexico Mexico 500 100
RedRock United States 55
Renewable Energy Group United States 250 50
World Energy United States 120 900 10 200
Total Americas 1.370 3.400 12 1.090
European Community
ENI Italy 750 150
Neste Porvoo Finland 260 50
Neste Rotterdam The Netherlands 1000 200
Preem Sweden 800 100
Sky NRG The Netherlands 100
ST1 Sweden 185
Total France 500 100
Total European Community 2510 800 0 885
Asia
Neste Singapore Singapore 1000 1000 400
Sinopec China 20
Total Asia 1000 1000 0 420
World Totals 4.880 5.200 12 2.395
368 Sustainable Aviation Fuels: Production, Use and Impact on Decarbonization

Usage mandates are frequent in land-based fuels and have emerged in discussions as alternatives to start using SAF, as well as
policies to encourage use in some countries. However, until today, aviation presents a blending mandate in Norway, with a
mandatory blending of 0.5% starting in 2020 and considering a target of 30% in 2030 (Norwegian Government, 2018).
Countries that are analyzing the possibility of mandates or objectives for SAF supplies (Velarde, 2020) include: (1) Germany
using 2% of SAF in 2030; (2) France planning to supply 2% in 2025 and 5% in 2030; (3) Spain focusing on 2% supply in 2025;
(4) Sweden with a blending mandate of approx. 1% in 2021, increasing to around 30% by 2030 and targeting a carbon neutral
country in 2045 (Swedish Government, 2019); (5) Finland mandatory blending to reach 30% in 2030 and targeting a carbon
neutral country in 2035; (6) The Netherlands leading to a blend of 14% and an ambition to replace all fossil fuels with SAF by
2050 (van Wamelen-Sibbes, 2020). The European Community (EC) has an initial discussion within the European Green Deal,
launched in December 2019, which considers legislative options to boost the production and adoption of sustainable products
and alternative fuels for the different modes of transport. Specifically, the discussion of the SAF is in the subset of actions called
ReFuelEU, among a set of possible actions to encourage use in study, including mandates.
Additional public policies can accelerate the introduction of the use of SAF, such as the Renewable Fuel Standard (RFS) (US
EPA, 2019) in the United States, the California Low Carbon Fuel Standard (LCFS) (CARB, 2020), Renovabio (Ministerio de Minas
e Energia, 2020), in Brazil, as the new national policy for biofuels, however these public policies have still not proved to be
sufficient to promote a parity of costs with fossil aviation fuel, but they can help to reduce cost gap. Carbon pricing policies have
been edited in several countries and, specifically for aviation, the EU ETS (European Commission, 2020b) and the recent CORSIA
(ICAO, 2020) by ICAO, but until now, they are also insufficient to close the cost difference renewable in relation to fossil fuel.

5.16.6 Conclusion and vision ahead

Biobased production of aviation fuels offers a great potential for GHG emissions reduction and fossil depletion reduction in
comparison to fossil kerosene, while it also provides socio-economic benefits in terms on job creation and gross domestic product
generation. Furthermore, the residues-based pathways could avoid potential negative environmental impacts occurring at a local
scale when 1G feedstocks are used (e.g., eutrophication, toxicity and air quality-related categories). However, despite the
encouraging environmental and social benefits, all biobased technologies for aviation fuel production still report significantly
higher production costs than those of fossil kerosene. Hence, the major challenges for large-scale commercial implementation of
aviation biofuel production are related to both technological improvement (for costs reduction) and institutional support (to
account for the environmental and social benefits from these novel supply chains).
In fact, the entire SAF chain, including the aviation industry, needs to find ways to foster the SAF production process and
approval. Moreover, customers are becoming selective and demanding aviation to show strong actions on climate change.
The possible composition of several different policies can make possible to reduce the costs of SAF to acceptable levels. These
impacts must be transparent and predictable. Such policies should not generate market distortions between airlines within the
same market, with the responsibility for reducing emissions being proportionally equal for each of the final consumers and they
must be built on a rationale of shared responsibility for use throughout all the product's value chain, from its origin to final
consumers, directly linked to the greenhouse gas emissions cycle. Also, public policies should reflect the goals of countries in
reducing emissions in the same way that this industry contributes (proportionally) to the country's emissions inventory, avoiding
unbalanced responsibilities for different players.

Acknowledgments

Authors acknowledge the Novo Nordisk Foundation (NNF), Denmark (grant number NNF20SA0066233) and the São Paulo
Research Foundation (FAPESP), Brazil (process #2015/20630-4).

References

Agusdinata DB et al. (2011) Life cycle assessment of potential biojet fuel production in the United States. Environmental Science & Technology 45(21): 9133–9143.
doi:10.1021/es202148g.
Ahmad E and Pant KK (2018) Lignin conversion: A key to the concept of lignocellulosic biomass-based integrated biorefinery. In: Bhaskar T, et al. (eds.) Waste Biorefinery:
Potential and Perspectives. pp. 409–444.1st edn. Elsevier.doi:10.1016/C2016-0-02259-3.
Alves CM et al. (2017) Techno-economic assessment of biorefinery technologies for aviation biofuels supply chains in Brazil. Biofuels, Bioproducts and Biorefining 11(1):
67–91. doi:10.1002/bbb.1711.
ANP (2018) Renovabio, National Agency of Petroleum Natural Gas and Biofuels. Available at: http://www.anp.gov.br/producao-de-biocombustiveis/renovabio (Accessed: 10
September 2020).
ANP (2020) Renovacalc - v.6.1, National Agency of Petroleum Natural Gas and Biofuels. Available at: http://www.anp.gov.br/producao-de-biocombustiveis/renovabio/renovacalc
(Accessed: 10 September 2020).
ASTM (2019) ASTM D4054-17 Standard practice for evaluation of new aviation turbine fuels and fuel additives (Vol. 05.02). West Conshohocken, PA: ASTM International.
ASTM (2020a) ASTM D1655 Standard specification for aviation turbine fuels (Vol. 05.01). West Conshohocken, PA: ASTM International.
ASTM (2020b) ASTM D7566 - Standard specification for aviation turbine fuel containing synthesized hydrocarbons (Vol. 05.04). West Conshohocken, PA: ASTM International.
Sustainable Aviation Fuels: Production, Use and Impact on Decarbonization 369

ATAG (2020) Who volunteers for CORSIA, Air Transport Action Group. Available at: https://aviationbenefits.org/environmental-efficiency/climate-action/offsetting-emissions-
corsia/corsia/who-volunteers-for-corsia/ (Accessed: 19 September 2020).
Atsonios K et al. (2015) Alternative thermochemical routes for aviation biofuels via alcohols synthesis: Process modeling, techno-economic assessment and comparison.
Applied Energy 138: 346–366. doi:10.1016/j.apenergy.2014.10.056.
Aviation Benefits Beyond Borders (2020) Sustainable aviation fuel. Available at: https://aviationbenefits.org/environmental-efficiency/climate-action/sustainable-aviation-fuel/
(Accessed: 10 September 2020).
Bailis RE and Baka JE (2010) Greenhouse gas emissions and land use change from jatropha curcas-based jet fuel in Brazil. Environmental Science & Technology 44(22):
8684–8691. doi:10.1021/es1019178.
Bann SJ et al. (2017) The costs of production of alternative jet fuel: A harmonized stochastic assessment. Bioresource Technology 227: 179–187. doi:10.1016/j.
biortech.2016.12.032.
Bauen A and Nattrass L (2018) Sustainable aviation biofuels: Scenarios for deployment. In: Kaltschmitt M and Neuling U (eds.), Biokerosene. pp. 703–721.Berlin, Heidelberg:
Springer.doi:10.1007/978-3-662-53065-8_27.
BECOOL (2017) Activities, Brazil-EU cooperation for development of advanced lignocellulosic biofuels. Available at: https://www.becoolproject.eu/activities/ (Accessed: 10
September 2020).
BECOOL (2019) Agreement signed for the kick off of the Brazilian BioValue project, Brazil-EU cooperation for development of advanced lignocellulosic biofuels. Available at:
https://www.becoolproject.eu/2019/05/07/agreement-signed-for-the-kick-off-of-the-brazilian-biovalue-project/ (Accessed: 10 September 2020).
Benjamin KR et al. (2016) Developing commercial production of semi-synthetic artemisinin, and of b-farnesene, an isoprenoid produced by fermentation of Brazilian sugar.
Journal of the Brazilian Chemical Society 27: 1339–1345. doi:10.5935/0103-5053.20160119.
Blommel PG and Cortright RD (2008) Production of conventional liquid fuels from sugars, pp. 1–14. Available at: https://www.etipbioenergy.eu/images/
Virent_Technology_Whitepaper.pdf.
Boymans E et al. (2019) A value chain for large scale FT production: The case of pyrolysis oil-char slurry gasification. 27th European Biomass Conference and Exhibition,
pp. 544–550. doi:10.5071/27thEUBCE2019-2BO.10.5.
Brander M (2017) Comparative analysis of attributional corporate greenhouse gas accounting, consequential life cycle assessment, and project/policy level accounting: A
bioenergy case study. Journal of Cleaner Production 167: 1401–1414. doi:10.1016/j.jclepro.2017.02.097.
Brazilian Government (2017) Law 13.576 - 26 December 2017, National Policy of Biofuels (Renovabio), Brazil. Available at: http://www.planalto.gov.br/ccivil_03/_ato2015-2018/
2017/lei/l13576.htm (Accessed: 10 September 2020).
Bressanin JM et al. (2020) Techno-economic and environmental assessment of biomass gasification and Fischer–Tropsch synthesis integrated to sugarcane biorefineries.
Energies 13(17): 4576. doi:10.3390/en13174576.
Brinkman MLJ et al. (2018) Interregional assessment of socio-economic effects of sugarcane ethanol production in Brazil. Renewable and Sustainable Energy Reviews 88(Feb):
347–362. doi:10.1016/j.rser.2018.02.014.
Bukur DB, Todic B, and Elbashir N (2016) Role of water-gas-shift reaction in Fischer–Tropsch synthesis on iron catalysts: A review. Catalysis Today 275: 66–75. doi:10.1016/j.
cattod.2015.11.005.
Capaz RS and Seabra JEA (2016) Life cycle assessment of biojet fuels. In: Chuck CJ (ed.), Biofuels for Aviation: Feedstocks, Technology and Implementation. pp. 279–294.
Amsterdam: Academic Press.
Capaz RS et al. (2020) Environmental trade-offs of renewable jet fuels in Brazil: Beyond the carbon footprint. Science of the Total Environment 714: 136696. doi:10.1016/j.
scitotenv.2020.136696.
CARB (2020) Low carbon fuel standard, California Air Resources Board. Available at: https://ww2.arb.ca.gov/our-work/programs/low-carbon-fuel-standard (Accessed: 21
September 2020).
Cardoso TF et al. (2018) Economic, environmental, and social impacts of different sugarcane production systems. Biofuels, Bioproducts and Biorefining 12(1): 68–82.
doi:10.1002/bbb.1829.
Cavalett O and Cherubini F (2018) Contribution of jet fuel from forest residues to multiple sustainable development goals. Nature Sustainability 1(12): 799–807. doi:10.1038/
s41893-018-0181-2.
Cheng F and Brewer CE (2017) Producing jet fuel from biomass lignin: Potential pathways to alkyl-benzenes and cycloalkanes. Renewable and Sustainable Energy Reviews 72:
673–722. doi:10.1016/j.rser.2017.01.030.
Chiaramonti D et al. (2014) Sustainable bio kerosene: Process routes and industrial demonstration activities in aviation biofuels. Applied Energy 136: 767–774. doi:10.1016/j.
apenergy.2014.08.065.
Choudhary P et al. (2020) A review of biochemical and thermochemical energy conversion routes of wastewater grown algal biomass. Science of the Total Environment 726:
137961. doi:10.1016/j.scitotenv.2020.137961.
Cortez, Luís Augusto Barbosa et al. (2014) Roadmap for Sustainable Aviation Biofuels for Brazil—A Flightpath to Aviation Biofuels in Brazil. Edited by Luis Augusto Barbosa
Cortez. São Paulo: Edgard Blücher. doi: https://doi.org/10.5151/BlucherOA-Roadmap.
Cortez LAB et al. (2015) Perspectives for sustainable aviation biofuels in Brazil. International Journal of Aerospace Engineering 2015: 1–12. doi:10.1155/2015/264898.
Cox K et al. (2014) Environmental life cycle assessment (LCA) of aviation biofuel from microalgae, Pongamia pinnata, and sugarcane molasses. Biofuels, Bioproducts and
Biorefining 8(4): 579–593. doi:10.1002/bbb.1488.
Darda S, Papalas T, and Zabaniotou A (2019) Biofuels journey in Europe: Currently the way to low carbon economy sustainability is still a challenge. Journal of Cleaner
Production 208: 575–588. doi:10.1016/j.jclepro.2018.10.147.
Dayton DC, Turk B, and Gupta R (2011) Syngas cleanup, conditioning, and utilization. In: Brown RC (ed.), Thermochemical Processing of Biomass. pp. 78–123.1st edn.
Chichester, UK: John Wiley & Sons, Ltd. doi:10.1002/9781119990840.ch4.
de Jong S et al. (2015) The feasibility of short-term production strategies for renewable jet fuels—A comprehensive techno-economic comparison. Biofuels, Bioproducts and
Biorefining 9(6): 778–800. doi:10.1002/bbb.1613.
de Jong S et al. (2017) Life-cycle analysis of greenhouse gas emissions from renewable jet fuel production. Biotechnology for Biofuels 10(1): 64. doi:10.1186/s13068-017-
0739-7.
Demirbas A (2009) Thermochemical conversion processes. In: Demirbas A (ed.), Biofuels: Green Energy and Technology. pp. 261–304.London: Springer. doi:10.1007/978-1-
84882-011-1_6.
Diederichs GW et al. (2016) Techno-economic comparison of biojet fuel production from lignocellulose, vegetable oil and sugar cane juice. Bioresource Technology 216:
331–339. doi:10.1016/j.biortech.2016.05.090.
Donofrio S, et al. (2019) Financing emissions reductions for the future. Washington. Available at: https://www.forest-trends.org/wp-content/uploads/2019/12/SOVCM2019.pdf.
Dry ME (2002) High quality diesel via the Fischer-Tropsch process—A review. Journal of Chemical Technology & Biotechnology 77(1): 43–50. doi:10.1002/jctb.527.
European Commission (2010) Renewable Energy Directive—Directive 2010/335/UE—Calculation of land carbon stocks.
European Commission (2018) Directive (EU) 2018/2001—Promotion of the use of energy from renewable sources, p. 128.
European Commission (2020a) Brazil-EU cooperation for development of advanced lignocellulosic biofuels, CORDIS—Horizon 2020. Available at: https://cordis.europa.eu/
project/id/744821 (Accessed: 10 September 2020).
European Commission (2020b) EU emissions trading system (EU ETS). Available at: https://ec.europa.eu/clima/policies/ets_en (Accessed: 12 September 2020).
370 Sustainable Aviation Fuels: Production, Use and Impact on Decarbonization

European Commission (2020c) Sustainable aviation fuels—ReFuelEU aviation. Available at: https://ec.europa.eu/info/law/better-regulation/have-your-say/initiatives/12303-
Sustainable-aviation-fuels-ReFuelEU-Aviation (Accessed: 19 September 2020).
Fan J et al. (2013) A life cycle assessment of pennycress (Thlaspi arvense L.)-derived jet fuel and diesel. Biomass and Bioenergy 55: 87–100. doi:10.1016/j.
biombioe.2012.12.040.
Finnveden G et al. (2009) Recent developments in life cycle assessment. Journal of Environmental Management 91(1): 1–21. doi:10.1016/j.jenvman.2009.06.018.
Furtado Júnior JC et al. (2020) Biorefineries productive alternatives optimization in the brazilian sugar and alcohol industry. Applied Energy 259: 113092. doi:10.1016/j.
apenergy.2019.04.088.
Geleynse S et al. (2018) The alcohol-to-jet conversion pathway for drop-in biofuels: Techno-economic evaluation. ChemSusChem 11(21): 3728–3741. doi:10.1002/
cssc.201801690.
Gonçalves VK (2017) Climate change and international civil aviation negotiations. Contexto Internacional 39(2): 443–458. doi:10.1590/s0102-8529.2017390200012.
Greenair (2008) Biofuelled Virgin Boeing 747 takes to the skies on pioneering first flight. Available at: https://www.greenaironline.com/news.php?viewStory=116 (Accessed: 9
April 2020).
Griffin DW and Schultz MA (2012) Fuel and chemical products from biomass syngas: A comparison of gas fermentation to thermochemical conversion routes. Environmental
Progress & Sustainable Energy 31(2): 219–224. doi:10.1002/ep.11613.
Guinée JB et al. (2011) Life cycle assessment: Past, present, and future. Environmental Science & Technology 45(1): 90–96. doi:10.1021/es101316v.
Hamrick, K. and Gallant, M. (2018) Voluntary carbon markets insights: 2018 Outlook and first-quarter trends, Washington. Available at: https://www.forest-trends.org/wp-content/
uploads/2018/09/VCM-Q1-Report_Full-Version-2.pdf.
Han J et al. (2013) Life-cycle analysis of bio-based aviation fuels. Bioresource Technology 150: 447–456. doi:10.1016/j.biortech.2013.07.153.
Haro P et al. (2013) Potential routes for thermochemical biorefineries. Biofuels, Bioproducts and Biorefining 7(5): 551–572. doi:10.1002/bbb.1409.
He J et al. (2019) Pyrolysis of heavy metal contaminated Avicennia marina biomass from phytoremediation: Characterisation of biomass and pyrolysis products. Journal of
Cleaner Production 234: 1235–1245. doi:10.1016/j.jclepro.2019.06.285.
Hileman JI and Stratton RW (2014) Alternative jet fuel feasibility. Transport Policy 34: 52–62. doi:10.1016/j.tranpol.2014.02.018.
Hileman JI et al. (2013) The carbon dioxide challenge facing aviation. Progress in Aerospace Sciences 63: 84–95. doi:10.1016/j.paerosci.2013.07.003.
IATA (2019) Fuel fact sheet. International Air Transport Association. Available at: https://www.iata.org/contentassets/ebdba50e57194019930d72722413edd4/fact-sheet-fuel.pdf.
ICAO (2010) Resolutions adopted at the 37th session of the assembly. International Civil Aviation Organization, p. 106. Available at: https://www.icao.int/Meetings/AMC/
Assembly37/Documents/ProvisionalEdition/a37_res_prov_en.pdf.
ICAO (2016) Resolution A39–2: Consolidated statement of continuing ICAO policies and practices related to environmental protection – Climate change. International Civil
Aviation Organization. Available at: https://www.icao.int/environmental-protection/Documents/Resolution_A39_2.pdf.
ICAO (2018) Annex 16. Environmental protection. Volume IV – Carbon offsetting and reduction scheme for international aviation (CORSIA). International Civil Aviation
Organization. Available at: https://www.unitingaviation.com/publications/Annex-16-Vol-04/#page ¼ 1.
ICAO (2019a) Annual report of the council—2018, Annual Reports of the Council, Montreal. Available at: https://www.icao.int/annual-report-2018/Pages/default.aspx (Accessed:
10 September 2020).
ICAO (2019b) CORSIA Default life cycle emissions values for CORSIA eligible fuels. Montreal. Available at: https://www.icao.int/environmental-protection/CORSIA/Documents/
ICAO document 06 - Default Life Cycle Emissions.pdf.
ICAO (2019c) CORSIA implementation plan brochure. International Civil Aviation Organization. Available at: https://www.icao.int/environmental-protection/CORSIA/Pages/
CORSIA-communication.aspx.
ICAO (2019d) CORSIA Methodology for calculating actual life cycle emissions values, Montreal. Available at: https://www.icao.int/environmental-protection/CORSIA/Documents/
ICAO document 07 - Methodology for Actual Life Cycle Emissions.pdf.
ICAO (2019e) CORSIA Supporting document: Eligible fuels—Life cycle assessment methodology. Montreal. Available at: https://www.icao.int/environmental-protection/CORSIA/
Documents/CORSIA Supporting Document_CORSIA Eligible Fuels_LCA Methodology.pdf.
ICAO (2019f) CORSIA Sustainability criteria for CORSIA eligible fuels. Montreal. doi: 10.18356/fc81178c-en.
ICAO (2020) Carbon offsetting and reduction scheme for international aviation (CORSIA). International Civil Aviation Organization. Available at: https://www.icao.int/
environmental-protection/CORSIA/Pages/default.aspx (Accessed: 10 September 2020).
IEA (2019) Key world energy statistics 2019. Paris.
IEA (2020a) Data and statistics. International Energy Agency. Available at: https://www.iea.org/data-and-statistics?country ¼ WORLD&fuel¼ Energy
supply&indicator ¼ TPESbySource (Accessed: 10 September 2020).
IEA (2020b) IEA Sankey diagram. International Energy Agency. Available at: https://www.iea.org/sankey/ (Accessed: 10 September 2020).
ISO (2018) ISO 14067:2018—Greenhouse gases—Carbon footprint of products—Requirements and guidelines for quantification’. Geneva: International Organization for
Standardization, p. 46. Available at: https://www.iso.org/standard/71206.html.
Kaltschmitt M and Neuling U (2017) In: Kaltschmitt M and Neuling U (eds.), Biokerosene: Status and Prospects. 1st edn. Berlin: Springer. doi:10.1007/978-3-662-53065-8.
Klein BC et al. (2018) Techno-economic and environmental assessment of renewable jet fuel production in integrated Brazilian sugarcane biorefineries. Applied Energy 209:
290–305. doi:10.1016/j.apenergy.2017.10.079.
Li L et al. (2010a) ‘Catalytic hydrothermal conversion of triglycerides to non-ester biofuels. Energy & Fuels 24(2): 1305–1315. doi:10.1021/ef901163a.
Li Y-W, Xu J, and Yang Y (2010b) Diesel from syngas. In: Vertés A.A., Qureshi N., Blaschek H.P. and Yukawa H. (eds.), Biomass to Biofuels. pp. 123–139.Oxford, UK:
Blackwell Publishing Ltd. doi:10.1002/9780470750025.ch6.
LNBR (2020) BioValue, Brazilian Biorenewables National Laboratory. Available at: https://lnbr.cnpem.br/biovalue/ (Accessed: 10 September 2020).
Lokesh K et al. (2015) Life cycle greenhouse gas analysis of biojet fuels with a technical investigation into their impact on jet engine performance. Biomass and Bioenergy 77:
26–44. doi:10.1016/j.biombioe.2015.03.005.
Maniatis, K., Weitz, M. and Zschocke, A. (2011) 2 million tons per year: a performing biofuels supply chain for EU aviation. Available at: https://ec.europa.eu/energy/sites/ener/
files/20130911_a_performing_biofuels_supply_chain.pdf.
Martinez Hernandez E and Ng KS (2018) Design of biorefinery systems for conversion of corn stover into biofuels using a biorefinery engineering framework. Clean
Technologies and Environmental Policy 20(7): 1501–1514. doi:10.1007/s10098-017-1477-z.
McKinsey & Company (2020) Hydrogen-powered aviation. A fact-based study of hydrogen technology, economics, and climate impact by 2050. https://doi.org/10.2843/471510.
Miller SA and Keoleian GA (2015) Framework for analyzing transformative technologies in Life Cycle Assessment. Environmental Science & Technology 49(5): 3067–3075.
doi:10.1021/es505217a.
Ministerio de Minas e Energia (2020) RenovaBio. Available at: http://www.mme.gov.br/web/guest/secretarias/petroleo-gas-natural-e-biocombustiveis/acoes-e-programas/
programas/renovabio (Accessed: 17 September 2020).
Moreira M, Gurgel AC, and Seabra JEA (2014) Life cycle greenhouse gas emissions of sugar cane renewable jet fuel. Environmental Science & Technology 48(24):
14756–14763. doi:10.1021/es503217g.
Motta IL et al. (2018) Biomass gasification in fluidized beds: A review of biomass moisture content and operating pressure effects. Renewable and Sustainable Energy Reviews
94(May): 998–1023. doi:10.1016/j.rser.2018.06.042.
NAE (2020) Summer bridge issue on aeronautics. National Academy of Engineering. Available at: https://www.nae.edu/234398/Summer-Bridge-Issue-on-Aeronautics.
Sustainable Aviation Fuels: Production, Use and Impact on Decarbonization 371

Neuling U and Kaltschmitt M (2018) Conversion routes from biomass to biokerosene. In: Kaltschmitt M and Neuling U (eds.), Biokerosene: Status and Prospects.
pp. 435–474.1st edn. Berlin: Springer. doi:10.1007/978-3-662-53065-8.
Neves RC et al. (2020) A vision on biomass-to-liquids (BTL) thermochemical routes in integrated sugarcane biorefineries for biojet fuel production. Renewable and Sustainable
Energy Reviews 119: 109607. doi:10.1016/j.rser.2019.109607.
Norwegian Government (2018) Luftfarten skal bruke 0,5 prosent avansert biodrivstoff fra 2020, Regjeringen.no. Available at: https://www.regjeringen.no/no/aktuelt/biodrivstoff-i-
luftfarten/id2613122/ (Accessed: 15 September 2020).
Parvez AM et al. (2020) Utilization of CO2 in thermochemical conversion of biomass for enhanced product properties: A review. Journal of CO2 Utilization 40(June): 101217.
doi:10.1016/j.jcou.2020.101217.
Pavlenko AN, Searle S, and Christensen A (2019) The cost of supporting a alternative jet fuels in the European Union. Available at: https://theicct.org/sites/default/files/
publications/Alternative_jet_fuels_cost_EU_2020_06_v3.pdf.
Pechstein J et al. (2018) Alcohol-to-Jet (AtJ). In: Kaltschmitt M and Neuling U (eds.) Biokerosene: Status and Prospects. pp. 543–574.Springer Nature: Berlin, Heidelberg.
doi:10.1007/978-3-662-53065-8_21.
Piris-Cabezas P, Lubowski R, and Leslie G (2018) Carbon prices under carbon market scenarios consistent with the Paris Agreement: Implications for the Carbon Offsetting and
Reduction Scheme for International Aviation (CORSIA). New York. Available at: https://www.edf.org/sites/default/files/documents/CORSIA Carbon Markets Scenarios_0.pdf.
Rauch R et al. (2018) Biokerosene production from bio-chemical and thermo-chemical biomass conversion and subsequent Fischer-Tropsch synthesis. In: Kaltschmitt M. and
Neuling U. (eds.), Biokerosene: Status and Prospects. pp. 497–542.1st edn Berlin: Springer. doi:10.1007/978-3-662-53065-8.
Rumizen M (2018) Aviation biofuel standards and airworthiness approval. In: Kaltschmitt M and Neuling U (eds.) Biokerosene: Status and Prospects. pp. 639–663. Springer
Nature: Berlin, Heidelberg. doi:10.1007/978-3-662-53065-8_24.
Rye L, Blakey S, and Wilson CW (2010) Sustainability of supply or the planet: A review of potential drop-in alternative aviation fuels. Energy & Environmental Science 3(1):
17–27. doi:10.1039/B918197K.
Santos CI et al. (2018) Integrated 1st and 2nd generation sugarcane bio-refinery for jet fuel production in Brazil: Techno-economic and greenhouse gas emissions assessment.
Renewable Energy 129: 733–747. doi:10.1016/j.renene.2017.05.011.
Seabra JEA et al. (2010) A techno-economic evaluation of the effects of centralized cellulosic ethanol and co-products refinery options with sugarcane mill clustering. Biomass
and Bioenergy 34(8): 1065–1078. doi:10.1016/j.biombioe.2010.01.042.
Seber G et al. (2014) Environmental and economic assessment of producing hydroprocessed jet and diesel fuel from waste oils and tallow. Biomass and Bioenergy 67:
108–118. doi:10.1016/j.biombioe.2014.04.024.
Sonnemann G and Vigon B (2011) Global guidance principles for life cycle assessment databases. A basis for greener processes and products. “Shonan guidance principles”.
United Nations Environment Programme, p. 160. Available at: https://www.lifecycleinitiative.org/wp-content/uploads/2012/12/2011—Global Guidance Principles.pdf.
Souza GM et al. (2015) Bioenergy & Sustainability: Bridging the Gaps. Paris: Scientific Committee on Problems of the Environment.
Staples MD et al. (2013) Water consumption footprint and land requirements of large-scale alternative diesel and jet fuel production. Environmental Science & Technology 47
(21): 12557–12565. doi:10.1021/es4030782.
Stichnothe H et al. (2016) Development of second-generation biorefineries. Technical, market, and environmental lessons from bioenergy. In: Lamers P, et al. (eds.), Developing
the Global Bioeconomy. pp. 11–40.1st edn. Elsevier Inc. doi:10.1016/C2015-0-02444-3.
Stratton RW, Wong HM, and Hileman JI (2010) Life cycle greenhouse gas emissions from alternative jet fuels, partnership for air transportation noise and emission reduction.
Cambridge. Available at: https://ascent.aero/documents/2020/02/life-cycle-greenhouse-gas-emissions-from-alternative-jet-fuels-version-1-2.pdf/.
Swedish Government (2019) Biojet för flyget. Stockholm. Available at: https://www.regeringen.se/493238/contentassets/6d591e58fd9b4cad8171af2cd7e59f6f/biojet-for-flyget-sou-
201911.
Teter J (2020) Aviation—tracking report, International Energy Agency. Available at: https://www.iea.org/reports/aviation (Accessed: 10 September 2020).
Thormann L and de Oro PP (2018) Fuels from pyrolysis. In: Kaltschmitt M and Neuling U (eds.) Biokerosene: Status and Prospects. pp. 575–606.1st edn. Berlin: Springer.
doi:10.1007/978-3-662-53065-8.
Tompsett GA, Li N, and Huber GW (2011) Catalytic conversion of sugars to fuels. In: Brown RC (ed.) Thermochemical Processing of Biomass. pp. 232–279.1st edn.
Chichester, UK: John Wiley & Sons, Ltd. doi:10.1002/9781119990840.ch8. Wiley Online Books
UK Ministry of Defence (2019) Defence Standard 91-091. Turbine Fuel, Kerosene Type, Jet A-1; NATO Code: F-35; Joint Service Designation: AVTUR. Bristol: UK Defence
Standardization.
US DOE (2020) Sustainable aviation fuel: Review of technical pathways. Washington. Available at: https://www.energy.gov/sites/prod/files/2020/09/f78/beto-sust-aviation-fuel-
sep-2020.pdf.
US EIA (2020) U.S. Energy information and administration. Available at: https://www.eia.gov/ (Accessed: 10 September 2020).
US EPA (2010) Renewable fuel standard program (RFS2)—Regulatory impact analysis, EPA-420-R-10-006. Available at: https://nepis.epa.gov/Exe/ZyPDF.cgi/P1006DXP.PDF?
Dockey=P1006DXP.PDF.
US EPA (2019) Renewable fuel standard program, Environmental Protection Agency. Available at: https://www.epa.gov/renewable-fuel-standard-program (Accessed: 28 March
2019).
van Wamelen-Sibbes L (2020) Sustainable aviation fuels. Paris: Organisation for Economic Co-operation and Development. Available at: https://www.itf-oecd.org/sites/default/
files/docs/agenda-decarbonising-air-transport-expert-workshop-24-25-february-2020.pdf.
Velarde C (2020) Availability and supply: How much SAF can aviation get? European Commission Workshop on Sustainable Aviation Fuels.
Wang W-C and Tao L (2016) Bio-jet fuel conversion technologies. Renewable and Sustainable Energy Reviews 53: 801–822. doi:10.1016/j.rser.2015.09.016.
Wang Z et al. (2019) Socioeconomic effects of aviation biofuel production in Brazil: A scenarios-based input-output analysis. Journal of Cleaner Production 230: 1036–1050.
doi:10.1016/j.jclepro.2019.05.145.
Wang H et al. (2020) Catalytic routes for the conversion of lignocellulosic biomass to aviation fuel range hydrocarbons. Renewable and Sustainable Energy Reviews 120:
109612. doi:10.1016/j.rser.2019.109612.
Wei H et al. (2019) Renewable bio-jet fuel production for aviation: A review. Fuel 254: 115599. doi:10.1016/j.fuel.2019.06.007.
Weidema BP et al. (2018) Attributional or consequential life cycle assessment: A matter of social responsibility. Journal of Cleaner Production 174: 305–314. doi:10.1016/j.
jclepro.2017.10.340.
Why ESK et al. (2019) Renewable aviation fuel by advanced hydroprocessing of biomass: Challenges and perspective. Energy Conversion and Management 199: 112015.
doi:10.1016/j.enconman.2019.112015.
Wong HM (2008) Life-cycle assessment of greenhouse gas emissions from alternative jet fuels. Massachusetts Institute of Technology. Available at: https://citeseerx.ist.psu.edu/
viewdoc/download?doi ¼ 10.1.1.717.8365&rep¼ rep1&type ¼ pdf.
Xu X et al. (2013) Two-step catalytic hydrodeoxygenation of fast pyrolysis oil to hydrocarbon liquid fuels. Chemosphere 93(4): 652–660. doi:10.1016/j.
chemosphere.2013.06.060.
Zhang X (2016) Essential scientific mapping of the value chain of thermochemically converted second-generation bio-fuels. Green Chemistry 18(19): 5086–5117. doi:10.1039/
C6GC02335E.

You might also like