You are on page 1of 29

7 Liquid Fuel Synthesis

Matthew Realff

7.1 Introduction

Liquid hydrocarbon fuels are an essential component of our energy system. They have
unprecedented volumetric density of energy, roughly 32 MJ/liter for gasoline, com-
pared to a lithium-ion battery at 2.4 MJ/liter. This means that for activities requiring
high or sustained power delivery such as flying and shipping, there are no current
alternatives. Compared to electric motors, internal combustion engines have signif-
icantly lower efficiency, and the recent improvements in the electric vehicle sector
suggest that at least for light vehicle duty their use is not essential. However, despite
significant government action to promote electrification of the light duty fleet, there
will still be a period of transition that could last well over a decade as the developing
world increases its consumption of transportation services and the world builds the
capacity to electrify personal transportation. Therefore, there is an urgent need to
develop pathways to produce renewable liquid hydrocarbon fuels as part of the energy
transition to low carbon energy systems.
Liquid fuels come in three basic types depending mostly on their boiling point
ranges with specifications on other properties such as when they freeze, or stop being
able to be poured, and their content of pollutants such as sulfur. The three blend
types are gasoline, which has a boiling point in the range of [40–200]°C, kerosene
that is used for diesel and jet engines, [150–300]°C and fuel oils [175–600]°C, which
are further classified depending on their other properties. The specifications reflect
the current production processes, the refining and upgrading of crude oils, which are
themselves complex mixtures of thousands of components that vary depending on
the source of the crude. Different fractions coming from the distillation columns and
other upgrading within the refinery are blended to get the final fuel properties, and
this is why fuel oils are sometimes referred to as distillates. The heavier fuel oils,
given numbers #5 and #6, are the residuals left over when the lighter components of
certain distillate fractions, forming fuel oils #1 and #2, are distilled away. The wide
range of boiling points of components in each of the blends means that refineries
manipulate their blending schemes to maximize the value they can derive from their
distillation streams. This provides a challenge and an opportunity for fuels from other
sources. First, compatibility with the existing fuels infrastructure, both distribution
and use, will make any transition easier but also constrain the types of molecules we

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


Liquid Fuel Synthesis 217

can consider. Second, because fuels are already blends, it is possible to add renewably
sources components and slowly increase their fractions, creating smooth transition
pathways.
In addition to the main fuel components that provide the energy for combustion,
there are many smaller additives that are used to improve the fuel performance, par-
ticularly components containing oxygen, oxygenates, that are used to prevent partial
fuel combustion that leads to soot formation in diesel engines and unburnt hydrocar-
bons in gasoline engines. This is why the specification of the fuels is often perfor-
mance driven, rather than a precise specification on the composition, and any blending
of new components must be tested to see if the blend meets the performance-driven
standards.
Liquid fuels derived from renewable sources are a relatively small fraction of light
duty vehicle gasoline (spark ignition) and an even smaller fraction of diesel (compres-
sion ignition) fuel for heavy duty vehicles, and effectively nonexistent for aviation
and shipping, although there have been renewable fuels that have qualified for jet
fuel and test flights (Doliente et al., 2020). Ethanol for gasoline blending is the largest
renewable fuel and has reached the largest percentage of fuel in Brazil, mandated
at 27%, about 87.3 million liters per day (Knoema, 2021). Ethanol is also required
in blends in the US with most gasoline at the 10% level which can be used in any
gasoline vehicle and 15% for cars of model year 2001 or newer (U.S. Department
of Energy, 2021a). There are flexible fuel vehicles that can use even higher blends
between 51% and 83%, and many cars sold in the US are of this type, but the avail-
ability of the blend is limited (U.S. Department of Energy, 2021b). European fuel
standards target the greenhouse gas emissions per km driven, which encourages the
blending of ethanol and biodiesel fuels with traditional hydrocarbons and have similar
fuel blending standards to those of the US given the global nature of the automotive
industry.

7.2 Renewable Fuels

We consider a renewable liquid hydrocarbon fuel component to have two essential


properties. First, it should be a liquid when stored or blended into a fuel at ambi-
ent conditions, as opposed to renewable hydrogen or natural gas which are covered
in Chapters 5 and 6, respectively. Second, the carbon it contains should have been
derived from carbon fixed either by biological photosynthesis or by carbon capture.
There will be certain cases where additional fossil fuels are used in the production of
the renewable hydrocarbon, particularly fuels for biomass supply chain and natural
gas in biomass conversion, and so not all the energy used in making the fuel will
necessarily have been renewable, but where possible this distinction will be noted.
In these cases, it is possible that the inputs could be replaced by renewable fuels
themselves, but the original analyses did not consider this option. In some of the cases
examined, the source of the CO2 is not specified, or a range of options are considered,

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


218 Matthew Realff

such as using CO2 from point source emissions that are not from biomass conversion,
these would not be renewable hydrocarbons if the point source was from fossil carbon.
In this chapter, we will not consider chemical routes to renewable hydrocarbons
through artificial photosynthesis which is an interesting research challenge but not yet
at the point where it could be deployed in the next decade (Davis et al., 2018). There-
fore, production routes to renewable fuels proceed either by conversion of CO2 itself,
if starting with direct air capture (McQueen et al., 2021), or with a compound that has
a general chemical formula of CxHyOz which captures the majority of biomass compo-
sitions that have been considered as sources for fuels (Fivga et al., 2019; Huber et al.,
2006) and will be extended to include algal biomass that also has some small fraction
of nitrogen which will impact fuel bound nitrous oxide emissions as discussed in the
chapter on emissions. We will include compounds that can be blended with conven-
tional fuels, dimethyl ether (DME) and polyoxymethylene dimethyl ethers (OMEs),
but will exclude alcohols such as methanol which would require a new infrastructure
for distribution and handling and ethanol which has been discussed at length else-
where (Dutta et al., 2011; Dutta and Phillips, 2009; Han et al., 2017; Humbird et al.,
2011; Lynd et al., 2017) and has blending limitations with existing hydrocarbon fuels.
The focus is on process pathways that utilize technology that could be deployed in
the medium term, the next 10 years, and therefore excludes pathways through direct
electrochemical reduction of CO2 by itself, which is covered in Chapter 5. We will
include promising new directions on the high temperature co-electrolysis of water and
CO2 which provides some significant benefits in energy efficiency.
There have been several reviews of renewable fuels from CO2 (Artz et al., 2018;
Becattini et al., 2021; Brynolf et al., 2018; Detz et al., 2018; Herron et al., 2015; van
der Giesen et al., 2014), all of which have concluded that although there is promise in
the technology, there is still a long way to go in improving their economics and their
carbon footprint due to the very large consumption of electricity that is inherent to
these processes. Similarly for renewable fuels from biomass, although the electricity
inputs are lower, the land area and water consumption are a significant challenge
because of the relative inefficiency of the conversion of sunlight to biomass, as will
be explained below. Another challenge of evaluating these systems is to balance the
specific details of the technological route with the ability to draw broader conclusions
about the possible trade-offs. This is seen in the way metrics of performance are cal-
culated and that there has not been a clear consensus on this topic (Albrecht et al.,
2017). There is a tension between the desire to have life-cycle assessment quality
results (de Jong et al., 2017; Han et al., 2013; Huo et al., 2009; Wu et al., 2006) and
detailed technoeconomic analyses and yet be able to draw out the common system
properties (Abanades et al., 2017; Gabrielli et al., 2020; Sutter et al., 2019).
This chapter will navigate this tension by providing a high-level process pathway
analysis that will capture many of the ways that renewable sources of carbon can be
converted into fuels, although an exhaustive listing is not possible, given the diver-
sity of routes and the rapidly changing nature of the field. It will consider the broad
questions of carbon sourcing and carbon footprint for direct air capture of CO2 versus
biomass carbon and what is required for such routes to be carbon negative given the

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


Liquid Fuel Synthesis 219

electrical intensive nature of producing hydrogen that is required for either route. It
will then pick several representative pathways that have been described in the liter-
ature and provide a summary of their performance, emphasizing economic, energy,
and carbon footprint metrics for each pathway. These pathways have been selected
based on different levels of technological maturity, feedstock, and the resulting fuel
to illustrate common challenges. It will conclude by examining the scale required to
replace our existing fuels with renewable fuels and general conclusions.

7.2.1 Carbon Source: Photosynthesis or Direct Air Capture?


A major consideration for renewable hydrocarbon fuels is the source of carbon. It
may seem obvious that biomass would be the starting point for renewable fuels, but
it has been recognized that the competition for land between biomass for energy and
for other human needs, particularly food, and the desire to sequester carbon in natural
systems and retain biodiversity means that there is significant competition for land to
grow biomass. At the heart of this competition is the relative lack of efficiency of pho-
tosynthesis itself. The terrestrial efficiency of sunlight to biomass varies significantly
by plant type and location, many crops are in the 1%–2% range (Hall and Rao, 1998),
but overall terrestrial photosynthesis is 0.6% efficient (Smil, 2008).
This relates to the generations of biofuels that have been developed, the first gen-
eration were based on crops for the obvious reasons of their existing cultivation and
higher photosynthetic efficiency, particularly for sugar cane which has high efficiency
and produces sugar directly. The sugar must be separated from the cane by mechanical
crushing, leading to waste bagasse at the processing plant that can be used to power
the rest of the conversion process. Corn, and other starches, requires an additional step
to break the starch into simple sugars, and the corn stover can be used for energy but
has not been collected because the cobs are easily separated in the field from the stover
compared to the sugar from the cane. This waste bagasse and stover, plus mature
forest industry infrastructure, led to the development of second-­generation biofuels,
lignocellulosics, which do not have as high photosynthetic efficiency as crops but do
not compete for arable land with crops, or because the growth cycle is multiple years,
can yield relatively high biomass per unit area when harvested. The third generation
is based on algae or cyanobacteria. Algae and cyanobacteria, either fresh or salt water,
are grown in dilute media either in large ponds or in photobioreactors. They have
both high photosynthetic efficiency, and the infrastructure can be built on marginal or
nonproductive land. Algae are attractive from a photosynthesis efficiency perspective,
possibly reaching 6%–8%, but to achieve these high values, the CO2 concentration
in the growth medium must be enhanced and the medium agitated to achieve good
transfer of the CO2 to the algae. Algae are also grown at very low biomass density in
the medium and so harvesting and recovering the algal mass is complex and occurs on
more rapid cycling than crops or lignocellulosics.
It is useful to compare the relative amounts of solar energy required to synthesize
fuels via biomass or via CO2 capture and conversion. For one MJ of biomass you
need 167 MJ of sunlight at an average terrestrial photosynthetic efficiency of 0.6%

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


220 Matthew Realff

(Smil, 2008). At the high end of photosynthetic efficiency, this might be reduced by a
factor of 10. Photosynthesis both captures the CO2 and performs an initial reduction.
For example, sugars such as glucose, C6H12O6, have carbon in an oxidation state of
zero, and some biomass components are even more reduced, such as lignin or seed
oils, and hence have higher energy content. If we assume a biomass composition of
CH1.66O0.83 a typical heat of combustion for this is 18 MJ/kg biomass, which means per
kg of biomass you need 3,000 MJ of sunlight, and for the carbon in the biomass this
is 6,820 MJ/kg carbon. Now, compare this value to that required using direct or air
capture of CO2 and water electrolysis for H2 production. Capturing CO2 from the air
takes roughly 7 MJ/kg and to reduce this to a zero oxidation state would take 2 moles
of hydrogen per mole of carbon dioxide, 0.091 kg H2 or 3 MJ per kg of CO2. If we
assume a 20% photovoltaic efficiency and 60% efficiency of an electrolyzer to make
hydrogen, then this is 25 MJ. If for convenience we assume the capture energy is all
electric, then the 7 MJ/kg becomes 35 MJ/kg, and therefore including capture a total
of 60 MJ solar energy/kg CO2, or 220 MJ/kg carbon.
Thus, the technological route to carbon in the same state, nearly zero oxidation
state, as that in biomass, is 3–30 times more energy efficient, if photosynthesis ranges
from 6% to 0.6% efficient. This does not account for other aspects of processing
biomass, such as the moisture content, which is often around 50% by weight, and the
inconvenience of transporting and handling solids as compared to gases and water in
the case of direct air capture. It also assumes that all the energy required to capture
the CO2 is electricity, whereas there are several implementations of direct air capture
(DAC) where low temperature heat is required and hence the efficiency from sunlight
can be made significantly higher for this step. It is apparent that the main use of energy
will be to make hydrogen rather than capturing the CO2, and this will be seen in later
discussion of specific fuels. The low photosynthetic efficiency also means that the
area required to capture the CO2 as biomass is also much larger than that required
to capture the CO2 using DAC and provide the energy to make the hydrogen and
the capture energy. These arguments of land and energy efficiency must be balanced
against the potential capital efficiency of biomass which requires no installation of PV
systems, less hydrogen generation, and no direct air capture technology.

7.2.2 Carbon Footprint Measurement


Assessing the carbon footprint of renewable fuels has proved to be a controversial topic,
particularly in the early years of first-generation biofuels, when it was falsely claimed that
the production of ethanol took more fossil energy than ended up in the fuel. The meth-
odology and data have improved substantially, and now the GREET model, developed
at Argonne National Labs (Argonne National Laboratory, 2021), has been established to
provide standardized and comparable results for different fuel pathways and for vehicles
themselves. The GREET model is regularly updated and now includes renewable fuels
from CO2 as well as from biomass, but recent research advances may take time to be
vetted and added to the database.

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


Liquid Fuel Synthesis 221

In understanding and reporting the carbon footprint of fuels, and renewable fuels
in particular, the key issues are the boundary of the life-cycle assessment and how
one treats the carbon that is in the fuel. For DAC systems, the carbon is removed and
returned to the atmosphere relatively quickly. This means that the carbon footprint is
that of the energy used to capture the carbon, make hydrogen and the fuel, and possi-
bly the infrastructure to do this if the lifetime is relatively short (Deutz and Bardow,
2021; Müller et al., 2020). For biomass, the situation is made more complex because
of plant growth cycles and the operations required to recover and process the biomass.
One concern in both systems is any land use conversion. For DAC systems, because
of their relatively small area, this is less important. For biomass, there has been much
discussion of how to estimate the CO2 that is released when land is converted from
one use, or no use, to another and how carbon that accumulates over a long grow-
ing cycle should be handled (Ahlgren & di Lucia, 2014; Broch et al., 2013). In the
discussion of the various fuels in the rest of the chapter, the carbon footprint will
be described where possible and boundary information of the carbon footprint will be
provided when known.

7.2.3 Electricity Carbon Footprint Bound


The production of a theoretical –CH2– moiety takes one CO2 and 3 H2. When this
is combusted, which is the reverse of the production process, it releases 0.612 MJ
of energy. If it were a regular gasoline would have a 96 g CO2 e/MJ carbon
­footprint. This enables us to bound the carbon footprint of the electricity used to
make the hydrogen for the renewable hydrocarbon. Making the same 60% efficiency
assumption on electrolysis or 56 kwhe/kg, the 6 g of hydrogen required to produce
the –CH2– takes 0.34 kwhe or to produce 1 MJ requires 0.55 kwhe. Therefore for
the renewable hydrocarbon to have a footprint of CO2 less than that of gasoline, the
electricity must have a carbon footprint less than 174 g/kwhe. Thus, any renewable
electricity source, PV, Wind or nuclear electricity meets this requirement (NREL,
n.d.), but the US grid had an average carbon footprint of 420 g/kwhe in 2019 (EIA,
n.d.). This represents an upper bound on the CO2e footprint of the electricity. It
does not account for the energy required to capture the CO2, which would reduce
the required footprint to less than 160 g CO2e, or any additional energy required for
the fuel conversion process.
At the current time, in the US, and in most countries in Europe with the possible
exception of France, the production of renewable fuels from CO2 would not have
a lower carbon footprint, using grid electricity, than those from traditional fossil
sources. Dedicated renewable energy sources would meet this requirement, and other
renewable sources, used when the supply exceeds demand, would also be beneficial.
This latter option should include energy storage, most likely as compressed hydrogen
rather than in batteries. This is because operating a chemical plant intermittently is
both complex, from a control perspective, and also capital inefficient if the utilization
factor is low since these processes are capital intensive. In the next major section,

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


222 Matthew Realff

several specific routes to renewable fuels will be discussed in more detail to illustrate
the general points raised in this section.

7.2.4 Fuel Conversion Process Descriptions and Metrics


The fuels that will be examined span from oxygen containing species such as dime-
thyl either and oligomeric methyl ethers to traditional straight chain hydrocarbons, all
of which are diesel substitutes as well as components that can be blended into gasoline
and jet fuel. The focus on diesel replacements is motivated by the relative difficulty
of replacing heavy duty transportation fuels by electric propulsion, particularly for
aviation jet fuel that is similar in composition. Where possible both a direct CO2 and a
biomass route using similar processes to the fuel will be examined to provide a level
of comparison between the two sources of carbon. In most cases, metrics of energy
conversion efficiency, carbon to fuel efficiency, and costs will be discussed. The orig-
inal source data will be presented, except for some unit conversions to bring as much
of the data on to the same basis as possible. The units of currency will not be changed
from the original source nor the specific year of the evaluation. This is because it is
misleading to simply convert from euros to dollars, when the relative costs of capital
and operating inputs are not the same in different locations, and the rates of inflation
are also different. In most cases, the evaluations are recent enough that the specific
currency amounts can be considered to be contemporary to date of publication and
the year of the evaluation is given when known. The conventional way to scale costs
from one year to another is to use an index, and for chemical plant capital cost, the
Chemical Engineering’s Plant Cost Index is often used. This enables capital costs to
be appropriately inflated, but other costs, such as energy, raw materials, and labor,
do not inflate in the same way. When known the CE index will be reported for the
particular study.

7.3 High Level Process Pathway Analysis

Figure 7.1 shows a schematic of the different pathways from renewable carbon
sources to the final fuels, including some key intermediates. The common feature of
all pathways is that oxygen must be removed from the system.

7.3.1 CO2 Pathways


In the case of using CO2 as a feedstock, this can happen either through reaction with
hydrogen or through direct CO2 electrolysis. As mentioned earlier, we will consider
only reaction with hydrogen for the rest of the chapter as this is the route that is clos-
est to deployment in the next decade. Routes for CO2 to fuels are multiplying rapidly
driven by advances in catalysis and bioprocesses. There are two major routes: indirect
and direct. In the indirect route, which is the one most well-developed, the CO2 is

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


Liquid Fuel Synthesis 223

Figure 7.1 High level schematic of renewable fuel routes.

converted into a syngas and then the syngas further converted using existing process
designs, such as to methanol, through Fischer–Tropsch chemistry, or by fermentation
to alcohols or organic acids. In the second case, CO2 is directly converted within
a single reactor to a fuel molecule such as DME or a hydrocarbon. The difference
is that for direct synthesis, the catalyst must have the functionality to both reduce
the CO2 and then perform the subsequent reactions, a so-called tandem or bifunc-
tional catalyst. It must also be the case that these reactions can be carried out at the
same conditions of temperature and pressure and often this requires compromising the
effectiveness of one component of the reaction pathway. For example, the production
of syngas typically requires high temperatures, but for fermentation, it is necessary to
cool the syngas down which would be impossible in a single catalytic reactor. In all
cases, large amounts of hydrogen are consumed by necessity to remove oxygen from
the CO2, with at least two moles of hydrogen required from a stoichiometric reaction
balance. The only way to reduce the hydrogen required is to leave some of the oxygen
in the fuel.

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


224 Matthew Realff

7.3.2 Biomass Pathways


In the case of biomass, there is the possibility of removing some oxygen with carbon as
CO2 but not all of it can be because this would result in no carbon being retained to turn
into a fuel. In most cases, we want to maximize the carbon that is retained in the fuel
product because otherwise we used photosynthesis to capture CO2 only to then reject
it, but there are cases, such as with fermentation or anaerobic digestion, where this is
an inevitable part of the metabolic process of the organisms or the chemical reaction
pathway. In the production of renewable natural gas through anaerobic digestion, the
gaseous carbon is partitioned between CH4 and CO2 depending on the feedstock, con-
ditions, and bacterial colony. A membrane separation can be used to remove CO2 from
the stream before syngas generation, and this allows for a range of syngas composi-
tions (see Chapter 6). In fermentation to alcohols, the carbon number of the alcohol is
the key factor. For ethanol, see Equation 7.1, the stoichiometric balance means 33% of
the carbon is rejected as CO2 and no oxygen is rejected with hydrogen, and in the case
of butanol, see Equation 7.2, the oxygen is rejected both with hydrogen and carbon.

C6 H12 O6 ® 2C2 H 5OH + 2CO2 (7.1)


C6 H12 O6 ® C 4 H 9 OH + 2CO2 + H 2 O. (7.2)

For thermochemical routes, again oxygen can be rejected both with carbon and
hydrogen depending on the chemistry, but the molecules are a much broader spectrum
of compounds rather than typically the few that are produced in fermentation. This
means at some point in the biomass-to-fuels process, we will end up with compounds
that have CxHyOz compositions that are different from the final fuel and we will need
process steps to change them.
There are two required operations, first to remove more oxygen using hydrogen,
hydrodeoxygenation, and second to change the carbon number from lower or higher
values into the fuel range. In the second step, increasing the carbon number is accom-
plished either through a Fischer–Tropsch mechanism from syngas or through oli-
gomerization of alkenes such as ethene or butene that are produced from methanol or
other higher alcohols or organic acids by the previous hydrodeoxygenation. The pro-
duction of alcohols or organic acids can be accomplished through the fermentation of
sugars. The other thermochemical routes from biomass proceed through intermediates
such as syngas by gasification, pyrolysis oils, or the organic phase of hydrothermal
liquefaction (HTL). The choice between pyrolysis and HTL is driven by the initial
moisture state of the biomass. Pyrolysis proceeds at modestly high temperature in oxy-
gen poor environments to use some of the carbon and hydrogen in the feed to remove
oxygen and concentrate the remaining carbon into a volatile phase that is subsequently
condensed, the oil phase separated from water, and further refined. The oil phase is
often unstable, able to polymerize since it still contains reactive oxygen and double
bonds, and cannot be stored for long periods. Pyrolysis is often carried out with short
residence times, fast pyrolysis, to avoid breaking down the biomass into light gases
to retain as much of the carbon in the liquid phase. Since much of the incoming mass
is vaporized, this means that wet biomass must be dried, which is energy intensive,

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


Liquid Fuel Synthesis 225

before the pyrolysis step, particularly if high rates of heating are desired within the
pyrolysis reactor to enable fast pyrolysis. In contrast, HTL uses high pressures and
moderate temperatures, keeping the water in the liquid phase, and using water’s power
as a solvent to help deconstruct the biomass. This makes it especially suitable for wet
biomass processing and in particular for dilute biomass streams such as algae.
The biomass is thermally deconstructed into a range of carbon numbers with a
significant fraction in the C20+ which are subsequently hydrocracked into smaller
molecules using hydrogen over a catalyst. All of these processes lead to mixtures of
hydrocarbons with carbon numbers in the range for fuels as well as light gases, meth-
ane through propane, and if not hydrocracked, waxes that can be used as lubricants
often with significantly higher value than fuels.
Therefore, the final energy and mass efficiency of any pathway is governed by
how effectively the different process steps manage the carbon, hydrogen, and oxygen
balances, and by how much oxygen the final form of the fuel contains, because remov-
ing more oxygen must inevitably require carbon to be lost, hydrogen to be added, or
electrolysis used to produce oxygen directly at the anode of the electrochemical cell.
The first-order design of the conversion pathway is complemented by the energy
efficiency of specific pathway elements and the final efficiency of the fuel in the com-
bustion system. In the case of biomass, there is the choice between gasification, pyrol-
ysis, and HTL based on the amount of water in the biomass feed. The early steps in
these pathways are covered in detail in Chapter 10 on solid biomass fuels, and so only
a limited discussion will be presented in this chapter. For CO2, the current choices are
more limited, essentially a syngas intermediate or some form of direct electrolysis
which reflects that the forms of biomass are more diverse and complex than CO2. For
the final efficiency of the fuel, this is a factor of its liquid density, oxygen content,
and the form of combustion engine, either spark ignition for gasoline-type blends
or compression ignition for diesel blends. Overall this leads to the concept of “well
to wheels” efficiency which can be measured in different ways but most commonly
looks at the CO2e emissions, say in grams, per km of vehicle travel, considering all
the steps along the supply chain from the original sources of carbon and hydrogen
to the final fuel efficiency in MJ/km, more commonly expressed as miles per gallon
in the US (de Jong et al., 2017; Han et al., 2013; Huo et al., 2009; Wu et al., 2006).

7.4 Specific Fuel Pathways

The discussion will be organized by the process routes depicted in Figure 7.1.

7.4.1 Dimethyl Ether (DME) CH3OCH3-Based Pathways


DME Fuel – DME is a potential substitute for both liquified petroleum gas and com-
pression ignition (diesel) engines. It is not a liquid at ambient conditions but can be
stored at moderate pressure, 5 bar, similar to propane. It can also be used as an inter-
mediate in the production of gasoline for spark ignition engines. DME can be made

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


226 Matthew Realff

directly from CO2 and H2 or indirectly by first producing methanol, CH3OH. In both
cases, syngas is an intermediate that is formed and consumed, and this can either have
a H2:CO ratio of one, Equation 7.3, or two, Equation 7.4, depending on the route to
DME.
3CO + 3H 2 ® CH 3OCH 3 + CO2 (7.3)
2CO + 4H 2 ® CH 3OCH 3 + H 2 O. (7.4)
The most important difference in routes A and B is that A rejects the excess oxy-
gen as CO2, rather than water, and hence is constrained to use only two-thirds of the
carbon feed, but saving one mole of hydrogen. Overall, if starting from CO2, the
consumption of hydrogen to make the syngas is equal to the moles of CO required.
Biomass gasification can produce either of the two syngas hydrogen ratios depending
on the reactor conditions and configuration.
Michailos et al., (2019) presents a technoeconomic assessment of producing DME
from CO2 captured from a cement kiln, using a membrane technology, and hydrogen
produced by PEM electrolysis. The main parameters of the process are summarized
in Table 7.1.
The cost of DME production, euros 75.62/GJ, is approximately five times higher
than diesel price in Europe. The cost is dominated by the electricity which is 71% of
the overall production cost, if the electricity were reduced in price to zero, then the
DME price would drop to 23.7 euros/GJ LHV, still more than the price of diesel, but
this demonstrates the vital role that electricity price plays in renewable fuels. Simi-
larly, the energy consumption is dominated by the electrolysis, which is 95% of the
overall energy use. The capture of CO2 from the more concentrated, 20%–28%CO2,
cement kiln flue gas lowers the energy required by an order of magnitude compared
to direct air capture, but this is avoided emissions rather than negative emissions, and
so is not a renewable carbon source. However, as mentioned earlier, the energy to
capture CO2 is a relatively small contributor to the overall energy to produce the fuel
and hence this is a reasonable proxy for CO2 from dilute sources.
A study by Bongartz et al. (2018) presented the well-to-wheels carbon footprint
and cost estimate of DME, along with hydrogen, methanol and methane which allows
for a direct comparison of these fuels. The source of the CO2 for these fuels is dis-
tributed capture from biogas. The CO2 is sent by pipeline to the conversion facility
which draws hydrogen from a cavern storage system to buffer a renewable generation
source. The production of hydrogen is therefore outside the boundary of this analysis,
and instead, a price per kg of hydrogen is used for economic analysis. The process
uses the direct approach for DME synthesis. The main parameters are summarized in
Table 7.2.
The definition of chemical energy efficiency is given in Equation 7.5
mfuel LHVfuel
hCCE = . (7.5)
mH2 LHVH2

The well-to-wheel CO2e includes the CO2 and H2 transport to fuel production, the
plant electricity consumption, and the distribution but not the rerelease of the CO2

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


Liquid Fuel Synthesis 227

Table 7.1 Data from Michailos et al., 2019

Quantity Value Units

DME price 75.62 Euros 2016/GJ LHV


Overall CO2 conversion to DME 82.3 %
CO2 to DME 2.33 tonne CO2/tonne DME
Conversion energy efficiency 44.4 % (LHV)
Overall energy consumption 18.05 MWh/tonne DME
CO2 capture energy 0.522 GJ/tonne captured
H2 electrolysis energy 61.1 MWh/tonne H2
Electricity price 87.7 Euro/MWh
Total capex 533 Euro/tonne DME

Table 7.2 Data from Bongartz et al., 2018. The plant design
basis was 1 million tonnes of CO2/year

Quantity Value Units

DME price 50 Euros 2016/(GJ LHV)


CO2 to DME 1.41 tonne CO2/tonne DME
Chemical energy efficiency 90 % (LHV)
H2 price 4.5 Euro/kg
Total capex 325 Euro/tonne DME
GHG well-to-wheel 18.2 Kg CO2e/GJ (LHV)

on fuel combustion. It is assumed the fuel carbon was recently captured from the
atmosphere.
DME to Hydrocarbon Fuels – A study presented in (Trippe et al., 2013) calculates
the cost of gasoline fuels derived from biomass-based DME in Europe. This uses
the bioliq® concept developed at Karlsruhe Institute of Technology. The idea is to
collect biomass in a radius of 30 km of multiple pyrolysis plants which then ship the
pyrolysis oil and char to the gasification and fuel synthesis plant. This supply chain
organization produces a tenfold increase in volumetric energy density at the pyrolysis
plant, hence reducing the biomass logistics costs and enabling a large synthesis plant
to be built to take advantages of economies of scale. Once the biomass is gasified, it
can be converted into gasoline or diesel via DME or via a Fischer–Tropsch route, the
FT route via syngas will be discussed later. In addition to producing the liquid fuel,
a significant fraction of the energy in the syngas is used to coproduce electricity. The
main parameters of the process to make gasoline compounds through DME are sum-
marized in Table 7.3. The final gasoline cost at 1.15 euros per liter was about 1.76
times the price of gasoline in Europe in 2011 and included a credit for the coproduced
electricity at a value of 90 euros a MWh. The process generates CO2 from the pyrol-
ysis and gasification sections and the required H2:CO ratio of the syngas for gasoline
production means that some carbon is rejected from the system as CO2 during the
synthesis which can be stripped from the syngas and is suitable for sequestration.

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


228 Matthew Realff

Table 7.3 Gasoline via DME synthesis (biomass pyrolysis and direct to DME) data
from Trippe et al., 2013

Quantity Value Units

Syngas input 144 (758) Tonne/hr (MW HHV)


Gasoline 30 (394) Tonne/hr (MW HHV)
Electricity coproduction 126 MW
Energy efficiency from biomass 38 % (HHV)
Fixed capital investment 215 (M euros 2011 total investment)
Production costs 1.5 (euros/kg)
Volumetric production costs 1.15 (euros/ltr)
CO2 from biomass conversion 292.5 Tonne/hr
CO2 available for sequestration 88 Tonne/hr

The US DOE also examined a route from biomass to gasoline through DME in a
report published in 2015 (Tan et al., 2015). This pathway proceeds through biomass
gasification, production of methanol, conversion of methanol to DME, and then sub-
sequent catalytic conversion of the DME to a high octane (>93) gasoline blendstock.
The process was self-sufficient in energy, generating its own steam heat and power,
but did not export any electricity.
The results were reported in Gallons of Gasoline Equivalent (GGE), which
­provides a way to compare different fuels (U.S. Department of Energy, n.d.). By
this measure, diesel is 1.155 GGE and one kwh of electricity is 0.031 GGE. In
units of energy, a GGE is 121.3 MJ. The US DOE uses the concept of mini-
mum fuel selling price, which represents the price at which a fuel must be sold to
cover the operating costs of the plant, the costs of capital and an investors’ rate of
return, in this case 10%. However, the capital cost is not all equity but rather split
between equity and debt. The assumption in this study was for a 40% equity and
60% debt split with an 8% interest rate and a ten-year maturity, but a thirty-year
plant life. The study used a base year for costs of 2011, and the results are reported
in Tables 7.4 and 7.5 below for the efficiency and financial performance and sus-
tainability metrics.
The discussion of routes to DME from biomass or CO2 and from DME to gaso-
line demonstrates that the costs of these processes today are significantly, three to
five times, higher than the costs of conventional fuels. The energy and capital costs
reported in the studies demonstrate that efficiencies are 40%–45% from biomass and
that hydrogen production is a major factor in the cost from CO2. The costs from CO2
are higher, particularly capital costs, due to the additional carbon capture system.
Table 7.6 summarizes the costs using the units of $/GJ and assuming one Euro is
1.2 dollars. The more expensive fuel in Europe reflects both the difference in cap-
ital structure of the studies and the differences in costs between the United States
and Europe.

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


Liquid Fuel Synthesis 229

Table 7.4 Gasoline via DME synthesis (biomass gasification and methanol intermediates).

Quantity Value Units

Feedstock 2,000 Dry biomass tonnes/day


Efficiency to gasoline 45 % LHV
Carbon conversion efficiency 31.2 % (Carbon in fuel/carbon in feedstock)
Gasoline yield 7,870 (64.9) 7,870 MJ (GGE)/dry tonne
Gasoline production 177.7 Million liters/year
Gasoline MFSP 0.86 (3.41) 2011 $/liter ($/GGE)
Total installed equipment cost 251.3 Millions $ (2011 CE = 585.7)
Total capital investment (TCI) 437.5 Millions $ (2011 CE = 585.7)

Table 7.5 Gasoline via DME synthesis (Tan et al., 2015) (biomass
gasification and methanol intermediates) sustainability metrics

Quantity Value Units

Water consumption 1.7 Liter water/liter gasoline


Fossil GHG emissions 0.6 g CO2e/MJ fuel
Fossil energy consumption 0.006 MJ fossil energy/MJ fuel

Table 7.6 DME route costs for final fuel

Carbon source Final fuel Location (year) Fuel price $/GJ Source

Biogas CO2 DME EU 2016 60 (38)


Cement plant CO2 DME EU 2016 90.7 (37)
Biomass Gasoline EU 2011 47 (39)
Biomass Gasoline US 2011 28 (40)

7.4.2 Polyoxymethylene Dimethyl Ethers (OMEs) CH3O(CH2O)nCH3 Pathway


A class of diesel fuels that has recently gained attention are oligomers with the struc-
ture CH3O(CH2O)nCH3 labelled as OMEn (Bokinge et al., 2020; Burre et al., 2019;
Held et al., 2019). When n = 0, this is equivalent to DME, and when n = 1, this
is dimethoxymethane which is commonly referred to as methylal and is already a
chemical of commerce. Methylal is not suitable as a liquid fuel due to its low boiling
point, but OME3–5 boil in the diesel range. The process route starts with methanol, and
formaldehyde (FA) provides the OME monomer units. There are two routes, a direct
route from FA and methanol, Route A, and an indirect route through methylal and
trixoxane which is a trimer of the FA, Route B. (Held et al., 2019) examined three
sources of CO2 and the two different routes to OME3–5. The overall stoichiometry
of the routes is given in the equations below. The source and route were combined

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


230 Matthew Realff

with three different schemes for recovering heat to increase efficiency. Hydrogen was
provided to the process using PEM electrolysis, and a commercial methanol synthesis
route was used along with the BASF silver catalyst process for methanol conversion
to FA. The key process metrics are given in Table 7.7, the mass efficiency of the var-
ious cases is very similar, but the CO2 source, process route, and energy integration
influence the net energy. Specifically, the least energy will be used when the CO2
source is pure, Equation 7.6 is used, and heat integration is allowed across the entire
process and highest when direct air capture, Equation 7.7, and no integration is used.
The electrical energy consumption in the electrolyzer dominates the energy consump-
tion, as with the DME route.
n −1
CH 3OCH 2 OCH 3   CH 2 O 3 → CH3O  CH 2 O n CH3  H 2 O (7.6)
3
2CH 3OH  nCH 2 O ® CH 3O  CH 2 O n CH 3  H 2 O. (7.7)
There are two different ways to produce FA which differ in their consumption of
hydrogen by n moles. The first is partial oxidation of methanol and the second is the
direct dehydration of methanol, the partial oxidation approach has been developed
commercially but no process of direct dehydrogenation has yet been developed, and
Table 7.7 has results for the partial oxidation approach. The efficiency of electricity
to final fuel varies between 37% and 24%, some of this variation is explained by the
energy required to capture the CO2 but overall, the process is not particularly efficient
compared to the DME and other routes that will be discussed later in the chapter.

Table 7.7 Partial oxidation route to OME (Held et al., 2019)

Quantity Value Units

CO2 consumption 1.829 kg/kg OME3–5


H2O consumption 0.24 Kg/kg OME3–5
Energy consumption [51–78] MJ/kg OME3–5
H2 energy consumption 200 MJ/kg H2 (ηPEM = 60%)
OME energy 18.9 MJ/kg OME3–5

7.4.3 Hydrocarbons via Syngas and Fischer–Tropsch (FT) Pathway


The FT process was originally developed to make synthetic liquid fuels from coal. The
coal is gasified to make syngas and then converted to hydrocarbons using a relatively
low-cost iron or cobalt catalyst. The most important variable for the FT process is the
ratio of H2 to CO in the syngas which needs to be maintained at or above 2 to enable
oxygen to be rejected and the formation of the hydrocarbon chain, see Equation 7.8.
CO  2H 2 → −CH 2 −  H 2 O. (7.8)
The ratio of hydrogen to carbon in coal is less than this, and so additional hydrogen
is created through using the water gas shift (WGS) reaction, shown in Equation 7.9.
CO + H 2 O  CO2 + H 2 . (7.9)

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


Liquid Fuel Synthesis 231

This loses carbon as CO2 but enables the overall process to achieve the necessary
carbon to hydrogen balance. Clearly, in the case of starting with CO2 as a feedstock,
it is not the WGS reaction that is needed but its reverse, RWGS. This is feasible
because the overall reaction is an equilibrium between the four components and so
increasing the hydrogen concentration will result in shifting the CO2 and H2 to the left
and create the necessary CO for reaction. Thus, to reach an H2-to-CO ratio of 2 for the
formation of hydrocarbons, a ratio closer to 3 is needed for H2:CO2 if hydrogen is used
to remove oxygen in the reactor.
The adaptation of the FT process to making renewable hydrocarbons from CO2 has
recently been studied in (Marchese et al., 2021; Zang et al., 2021). In Wang, the CO2
and H2 are assumed to be available at the plant boundary; the suggested starting point
is concentrated CO2 from the production of ethanol along with the H2 so no transpor-
tation costs are included. In the base case, the CO2 is therefore assumed to have a posi-
tive cost. The RWGS is fed the H2 and CO2 at 1:1 to make the CO. The remaining CO2
is removed from the syngas before the FT reactor using a low temperature Selexol™
process and recycled to the RWGS reactor. The rest of the hydrogen for hydrocarbons
is fed to the FT reactor itself at a 2.2:1 ratio. The conversion of CO is not complete,
and the remaining hydrogen is recovered using pressure swing adsorption and recy-
cled. The major energy metrics of the process and process economics are presented
in Table 7.8. The fuel production rate of 351 tonnes per day corresponds roughly to
145 million liters of fuel per year.
The definition of the carbon conversion ratio is given below in Equation 7.10
and represents the amount of the carbon in the CO2 that ends up in the fuel, where
Ci is the mass fraction of carbon in the fuel or CO2, respectively. This is less than
50% for this process because a fraction of the process light gases are burnt for
power generation in a conventional boiler, using air, and therefore the CO2 cannot
be recycled.

 
CCR   ∑ m fuel × cfuel  m CO2 × cCO2 . (7.10)
The mixed fuels cost is significantly greater than current diesel prices of $2.3/­gallon
in 2020 and $3.1 in 2050. To break even at $17.3/tonne CO2, the hydrogen price
would be required to drop to $0.8/kg in 2050. To render the costs equal, a CO2 price
of negative $255/tonne is required in 2050. These are the extreme cases, a more likely
outcome would be to have both a reduction in the cost of producing hydrogen and a
negative price on CO2. Given that the feedstock costs comprise 95% of the variable
costs, the overall costs are more sensitive to the feedstock prices of H2 and CO2 com-
pared to other factors.
Marchese et al. (2021) considers integrated direct air CO2 capture and FT con-
version system along with the production of hydrogen via electrolysis. The overall
goal was to produce higher chain length waxes rather than fuels, but a significant
fraction of the product is in the naptha or middle distillate (diesel) range. The DAC
approach used is that of Carbon Engineering where soluble potassium hydrox-
ide is contacted with air to form potassium carbonate and then the potassium ion
exchanged with calcium hydroxide to create a solid precipitate of calcium carbonate

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


232 Matthew Realff

Table 7.8 Fischer–Tropsch route to hydrocarbon fuel (Zang et al., 2021)

Quantity Value Units

CO2 price 17.3 $/tonne CO2


CO2 use 3.1 kg CO2/kg fuel
H2 price 2 $/kg
H2 use 1.7 MJ H2/MJ fuel (LHV basis)
Cobalt FT catalyst 0.52 Conversion to hydrocarbons
Carbon conversion ratio 45.5%
Chemical energy efficiency 57.5% LHV basis
Production rate 351 Tonne/day fuels mixture
Equipment installed capex 257.8 M$
Depreciable capex 379 M$
Annual opex 20.2, 168.8 Fixed, variable M$
Fuel cost 1.56 (48.7) $/liter ($/GJ)

high temperature regeneration of the precipitate in a calciner is used to drive off the
CO2. The calciner requires significant energy input, and two cases are envisioned as
described in the original paper on the Carbon Engineering process (Keith et al., 2018).
The energy can either come from natural gas, as demonstrated in their pilot process,
or in a future embodiment as electrical heating.
A key integration for this high temperature DAC process is the coproduct stream
of pure oxygen from water electrolysis to provide the hydrogen for the fuel. The
natural gas fired–calciner requires pure oxygen to enable the CO2 to be recovered
without further separation from nitrogen. In the original Carbon Engineering process,
this required a separate air separation unit, which can now be eliminated by using the
oxygen from electrolysis. The overall integrated process schema are shown in Figure
7.2, where different possible cases are overlayed for the recirculation or combustion
of the FT off-gases and the use of natural gas for the calciner and other DAC power
needs through a combined heat and power system or imported electricity.
The key performance indicators for this process are the specific plant energy con-
sumption, the total plant energy efficiency, and the carbon efficiency. The definitions
are given in the equations below. The specific plant energy consumption per tonne of
CO2 removed is given in Equation 7.11 The CO2 removed is defined as the net amount
by subtracting both the CO2 that leaves in the air from the contactor and also any
exhaust CO2 that is produced by combustion of NG or FT off-gases, Equation 7.12.
The efficiency of CO2 removal is the ratio of the removal to the air input CO2; how-
ever, the carbon efficiency is how much of the carbon fed either as natural gas or in the
air that is converted to FT liquid products on a molar basis, hC, as show in Equation
7.13. The total plant energy efficiency, hGI , is the ratio of the energy in the FT liquid
products divided by the plant energy consumption, its value is given in Table 7.9.
WEl  QTh  n NG LHVNG
ESP  (7.11)
m CO2 ,rem

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


Liquid Fuel Synthesis 233

Table 7.9 FT with high temperature DAC captured CO2


performance metrics from Marchese et al., 2021

Quantity Scenario

Power in MW NG used FT off-gas


Electric power (electrolyzer) 838.1 784.3
Electric power (net other) 71.6 11.95
Net heat power 168.2 63.7
FT liquids power 419.8 235.8
ESP kWh/kgCO2 13.9 8.4
hCO2 % 64.4 68.3
hC % 73.8 68.3
hGI % 31.2 27.2
Capex (M euros) 1871 1595
Opex (M euros/y) 1996.3 1604.9
Wax cost (Euro/kg) (58.3/GJ) 30 44

Figure 7.2 Direct air capture integrated with Fischer–Tropsch process.


m CO2 ,rem  m CO2 ,AirIn - m CO2 ,AirOut  m CO2 ,Ex  (7.12)
  
 nC,Air  nC,NG - nC,Air  nC,Ex  
C    . (7.13)
In In Out

nC,AirIn  nC,NGIn

The analysis was performed on a 250,000 tonne/h of air basis containing 150 ­tonne/h
CO2. In the configuration in which natural gas is used both for CHP and the calciner,
an additional 53 tonnes/h of CO2 is added to the system. A second configuration in
which the natural gas is eliminated completely, and the FT off-gas energy is used
to provide all the heat and power for the capture system. The main parameters and

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


234 Matthew Realff

results for each system are given in Table 7.9. The fuel production for the case of
recycling FT off-gas is 235.8 MW, which corresponds to 200 million liters of gasoline
per year.
The NG case has significantly greater export of FT liquids for the given air flow-
rate, but this is partly caused by the increased carbon flow from the natural gas. As
expected, the majority of the electricity is used in the production of the hydrogen. The
net heat demand is lower for the FT off-gas because the flows through the WGS reac-
tor are smaller. In the off-gas case, the net heat is provided by burning the off-gases
rather than natural gas. The amount of carbon included in the fuels is 73 and 68%,
respectively, for the natural gas and FT off-gas cases and the CO2 from the air that is
included in the fuels is slightly lower for the case with natural gas. The overall energy
efficiency is 31% when natural gas is used and 27% for the FT off-gas. The cost of
the paraffin waxes is 30 euros per kg for the NG case, and 45 for the FT off-gas, based
on the baseline electricity price and a return on capital of 12.5%. This is an order of
magnitude higher than the current prices of these waxes, which is 2.5 euros a kg. One
reason for the high price is that the baseline electricity cost is 210 euros/MWh which
reflects a mix of hybrid PV and wind power plants. Using the lowest cost electricity
from hydropower at 41 euros/MWh these prices drop to 9 and 12 euro/kg. The overall
system efficiency is dependent on the energy consumption of the DAC either because
of the NG use in the calciner or because of the fraction of FT off-gas that must be
burnt as opposed to returned to the WGS reactor. Thus, improving the DAC effi-
ciency, particularly by reducing the high temperature heat required, will enable more
FT fuels and waxes to be produced.

7.4.4 Biomass Pathway Analysis


Aside from gasifying biomass to produce syngas another starting process is pyrolysis.
This route is preferred for dry biomass. For wet biomass, where the energy required to
dry it would be prohibitive, HTL is another route that will be described in a later sec-
tion. In both cases, an intermediate “bio-crude” is produced that must be further refined
into the final fuel. The US Department of Energy (DOE) has studied the conversion of
biomass, particularly lignocellulosic biomass, to fuels via pyrolysis extensively over
the previous several decades and has an extensive reporting structure that has analyzed
the state of technology (SOT) for this fuel pathway. The production system can be
divided into two parts: first, the biomass harvesting, handling, transportation logistics,
and preprocessing that brings the biomass to the “throat” of the process and second, the
process conversion itself from the throat to the fuel blending tanks.
Biomass Feedstock Logistics – The cost of the biomass feedstock is one of the
most important factors in biomass to fuels. The US DOE has studied this intensively
with Oakridge National Labs producing a comprehensive analysis on the availabil-
ity of biomass in the United States, commonly referred to as the “billion ton” study
(Langholtz et al., 2016). This considers a variety of different biomass feedstocks,
forestry, agricultural, and microalgae. The Idaho National Labs maintains an update
specifically on woody biomass feedstocks (Hartley et al., 2018). The study includes

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


Liquid Fuel Synthesis 235

Table 7.10 Summary of modeled cost estimates for woody feedstock supply systems for
IDL and CFP (Hartley et al., 2018)

Cost ($/dry/ton) (2016 $) GHG emissions

Clean pine Logging residue Total (kg CO2e/dry ton)

Grower payment 15.73 3.75 9.74


Harvest and collection 9.88 0 4.94 6.88
Field preprocessing 4.73 12.09 8.41 13.61
Transportation 7.67 16.77 12.22 27.24
Preprocessing (IDL) 28.14 15.59 21.87 58.70
Preprocessing (CFP) 30.55 26.55 28.55 75.1
Storage 0.67 0.67 0.67 0.92
Handling 2.65 2.65 2.65 2.13
Preprocess capex (IDL) 2.73 2.73 2.73
Preprocess capex (CFP) 2.96 2.96 2.96
Total (IDL) 72.2 54.25 63.23 109.48
Total (CFP) 74.85 65.45 70.15 125.87

woody biomass supply to a gasification-based indirect liquefaction (IDL), catalytic


fast pyrolysis (CFP) and a HTL process with a blend of algal biomass (Algae-blend
hydrothermal liquefaction (AHTL)). The study utilizes the Biomass Logistics Model
(Cafferty et al., 2013) and is based on a 50/50 blend of clean pine and forest residues
for the IDL and CFP routes and a forest residue stream for the AHTL where 90% of
the biomass comes from algae. The location of the IDL and CFP routes is in the South
East US and for AHTL in the gulf coast. The blending reduces the ash content from
soil entrainment and makes it suitable for the CFP system and increases yields from
gasification. The cost summary taken from Table 7.2 of Hartley et al., (2018) is given
in Table 7.10. The harvest and collection cost for the residue is assumed zero because
it is brought to the field side as part of the harvest process.
A sensitivity study on the energy consumption of the supply chain machinery
showed variations in cost of no greater than a dollar per dry ton. However, the capac-
ity of the grinder for logging residue was considered to vary by a factor of four from
0.4, to 0.8, to 1.6 dry tons/hr with a range of impact from reducing costs from the
middle case of $8.84 and increasing by $17.68. The total US availability of the pine
chips and pine residues in 2019 was roughly 12 million dry tons of planted pine and
10 million dry tons of pine residues and that 16.2 million dry tons could be accessed
at the price of $70 dollars or less. The estimate for this feedstock cost has significantly
decreased from 2014 to 2019. It was roughly $100/ton in 2014 and $87.80 in 2018,
but there is no projected decrease from 2019 to 2022 with a slight increase due to
inflation. This suggests that without a radical redesign of the logistics and preprocess-
ing systems the cost has plateaued.
The composition of the woody biomass in Hartley et al. (2018) was 50% c­ arbon,
which puts the cost per tonne of carbon at $140/tonne, with an energy content
of 18.3 MJ/kg dry biomass, that is, this much energy will be released on combustion.

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


236 Matthew Realff

Again, to compare with DAC CO2, this has an energy content of zero and the cost
range is similar.
Biomass Pyrolysis Route Analysis – The Pacific Northwest National Laboratory
(PNNL) has been the lead institution within the US DOE lab system for thermochem-
ical routes from biomass to fuels and has produced reports over the last decade on this
topic. In Zhu et al. (2011), technoeconomic analyses of five thermochemical routes
from biomass to fuels are presented, including pyrolysis and HTL. The pyrolysis and
hydrotreating routes were subsequently updated in reports (Jones et al., 2013, 2016).
In Jones et al. (2014), the HTL of algae is presented, and an update on algae wood
blends given in Zhu et al. (2020), which is congruent with the feedstock update by
Hartley et al. (2018).
A block diagram of the Fast Pyrolysis process based on Jones et al. (2016) is given
in Figure 7.3. The study is based on two parallel trains of 1,000 dry tonnes/day capac-
ity with a pyrolysis oil yield of 62% by mass. The main focus of process improvement
has been on the hydrotreating component of the process as this is a major cost and
different from the hydrotreating of conventional hydrocarbons. Tables 7.11 and 7.12
summarize the SOT projections both technical and sustainability metrics for 2014
and those projected at the time for 2017. Table 7.11 shows that natural gas is burnt to
provide energy or hydrogen to the process, and so technically this would not qualify
according to the chapter scope. If we assume that the natural gas is mostly used in the
steam reformer to provide hydrogen for the hydroprocessing units, then we can recal-
culate the amount of external hydrogen represented by this natural gas. We assume
all the natural gas combustion energy is used to make hydrogen with 100% efficiency
with regard to hydrogen LHV, which will provide an upper bound on the amount of
hydrogen. This hydrogen has been added as the last row of Table 7.11. Note that more
hydrogen may be used in the treating sections, but this is produced by converting
some of the energy or hydrogen in the biomass itself. The improvements projected
between 2014 and 2017 were significant and based mostly on improving the hydro-
processing economics by optimizing the catalyst performance, cost and lifetime, and
reactor operation. The overall carbon efficiency of the process leads to about half the
carbon in the biomass included in the fuel. These processes are also currently con-
figured to use fossil energy in the form of natural gas to provide hydrogen, and this
contributes to a positive carbon footprint of 20 g CO2e/MJ which is significantly less,
about 20% of that of regular gasoline, as shown in Table 7.12.
Hydrothermal Liquefaction Pathway: As mentioned in the introduction, and in
Chapter 10, wet biomass, and particularly dilute algal biomass cannot be efficiently
dried and mechanical means, for example, filtration or centrifugation, is used to
remove as much water as possible as an initial step to concentrate the biomass up
to approximately 20wt% solids. The slurry is then pressurized, to avoid water evap-
oration, and then heated to moderately high temperatures, 300–350°C to produce
reactor products, a gas phase, organic liquid phase, and aqueous liquid phase (Elliott
et al., 2015). The gas phase often contains significant methane and can be burnt to
partially provide the energy to heat the feed. The aqueous phase has a mixture of
water-soluble organics, particularly organic acids, and ammonia and must be treated
before being recycled in the case of using algal feedstocks. The organic liquid phase

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


Liquid Fuel Synthesis 237

Table 7.11 Pyrolysis process performance 2015 state of technology and projected 2017 from
Jones et al. (2016), note that converting from GGE to liters gasoline equivalent is just the volumetric
conversion of 3.78 liters/gallon

Processing area cost


contributions and technical
parameters State of technology 2014 Projected 2017 Units

Conversion contribution 4.11 2.47 $/GGE


total (gasoline/diesel)
Feedstock contribution 1.17 0.92 $/GGE
Feedstock cost 101.45 80.00 $/dry ton
Fast Pyrolysis total (capex/ 0.78 (0.66/0.12) 0.75 (0.64/0.11) $/GGE
opex)
Upgrading hydroprocessing 2.40 (0.62/1.78) 0.95 (0.42/0.52) $/GGE
Fuel finishing hydrocrack 0.25 (0.16/0.09) 0.14 (0.07/0.07) $/GGE
and distill
Balance of plant 0.68 (0.29/0.39) 0.63 (0.29/0.34) $/GGE
Yield blendstock 87 87 GGE/dry ton
Pyrolysis yield (dry wt) 0.62 0.62 Organics/Wood
Upgrading carbon efficiency 0.68 0.68 Wt fraction
Natural gas use 2.76 2.66 GJ/dry ton
Implied external H2 (upper 22.9 22 Kg H2/dry ton
bound)

Table 7.12 Summary of progress on sustainability metrics

Sustainability metric State of technology 2014 Projected 2017 Units

Fossil emission GHG 19.4 18.9 g-CO2e/MJ fuel


Fossil energy 0.31 0.3 MJ Fossil/MJ fuel
consumption
Carbon-to-fuel 0.47 0.47 C in fuel/C in biomass
Water consumption 5.67 5.29 Liter/GGE

Figure 7.3 Hydrotreating in biomass routes.

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


238 Matthew Realff

is then treated in similar ways to the biocrude from a pyrolysis process, hydrotreating,
and cracking, to make gasoline and diesel fuel blend components. For more details on
the initial hydrothermal conversion see Chapter 10.

7.5 Future Process Concepts: Co-electrolysis of CO2 and H2O

An alternative approach, that has significant potential, is to use co-electrolysis of CO2


and H2O at high temperature to produce syngas directly. In this case, the efficiency of
electricity to syngas can be very high because the heat dissipated due to electrochemi-
cal inefficiencies can drive the reaction equilibrium, achieving theoretical efficiencies
of 100% minus the heat losses from the unit. For example, in Detz et al. (2018), this
pathway is estimated to have an overall efficiency of 77% based on an efficiency of
96% in the co-electrolysis and 80% in the FT conversion.
Kulkarni et al. (2021) describe a process based on co-electrolysis, FT, and a third
reactor that oligomerizes the light olefins that are produced through FT. The ratio of H2
to CO in this system was adjusted to 0.7, much lower than that in other processes and not
all the oxygen can be removed. This means that the final products contain some liquid
oxygenates that can be used as fuel additives. The process has two modes, one where
renewable electricity provides the chemical energy for the syngas and another where
humid methane is substituted for the air at the anode. The humid methane increases the
efficiency and allows for an operating strategy where methane is fed during times when
renewable energy is expensive. A summary of the process performance is provided in
Table 7.13. The scale of the plant is relatively small at 79,000 liters of fuel per day and
increasing this to 790,000 liters per day reduces the cost. However, a plant at the scale
of 790,000 liters per day requires ten times the renewable energy, close to half a GW,
which is larger than most renewable energy sources being developed today. However,
this example demonstrates that the very high costs of renewable fuels that have been
found in other studies do not reflect what is possible with highly efficient electrochem-
ical technologies that use electricity directly, rather than H2, to reduce the CO2 to CO.
However, this analysis does point to a problem with renewable fuels if they are to scale
to meet significant fractions of demand for liquid fuels.
The efficiencies of the different routes are hard to compare since they start with
biomass or CO2 as the carbon source and water or hydrogen as the hydrogen source.
None of the pathways have efficiencies exceeding 70% and pathways from CO2 cap-
tured from air have lower efficiencies due to the immaturity of technology for this
operation. Regardless of the source of carbon, the range of costs for the different fuels
spans $30–60/GJ compared to conventional gasoline at roughly $10/GJ indicating
that renewable drop-in fuels are still some distance from being cost competitive with
conventional ones. If conventional gasoline emits 96 kg CO2 per GJ, then this would
require a carbon tax of roughly $200–500/tonne of CO2 which well above the prices
for CO2 that have been proposed. The difference in cost is explained partly because of
the scale of modern oil refining systems compared to the renewable fuel production
examined here and also because refining systems are technologically mature.

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


Liquid Fuel Synthesis 239

Table 7.13 Future pathway based on solid oxide co-electrolysis of


water and CO2 from (Kulkarni et al., 2021)

Quantity Value Units

Throughput 79,000 Liters fuel/day


Electricity demand 48 MW
Electricity price 50 $/MWh
Plant efficiency 67 %
Energy demand 52.5 MJ/liter
TCI 60.3 (67.3) $M (NG configuration)
OPEX 30.4 $M
Levelized cost product 1.29 (0.91) $/liter (790,000 liter/day)
Levelized cost product 40.3 (28.4) $/GJ (790,000 liter/day)

7.6 Size and Scale of a Renewable Hydrocarbon Economy

Each of the fuel production processes examined in this chapter has been at the scale of
hundreds of millions of liters of liquid fuel per year. In 2019, the US domestic motor
gasoline demand was 146 billion gallons (550 billion liters), 1,000 times higher than
the annual capacities of the plant designs. The current fuel demand in the US is met
by roughly 100 US refineries. For the case of using CO2 as a feedstock and using the
approach of Kulkarni et al. (2021), which is the most energy efficient of those examined
in this chapter, this would require 8,000 TWh of electricity to produce the fuel. The
total generation of electricity in the US in 2019 was 4,000 TWh. Therefore, it would be
necessary to double the existing generation of the US to produce this fuel. Furthermore,
this electricity would have to have a low carbon footprint. It seems unlikely that we
will replace the US light duty vehicle demand with renewable hydrocarbons from CO2.
However, the rapid direct electrification of the light duty vehicle transportation sec-
tor means that this is projected to be unnecessary. The average fuel efficiency of new
gasoline vehicles was 10.6 km per liter (25 miles per gallon) in 2019. Each liter of gas-
oline contains 36 kwh of energy, leading to an energy consumption of 0.9 kwhfuel/km.
Electric vehicles, such as the Tesla 3, use 0.15 kwhe/km of electricity, which is over
six times as efficient as the gasoline power vehicle. Therefore, with much lower elec-
tricity generation, the same miles can be driven, provided the necessary distribution
and charging infrastructure can be constructed, the electric car manufacturing industry
grows at a sufficient rate, and sufficient numbers of consumers accept the transition or
are forced to by government policy.
There are other sectors, such as heavy-duty diesel vehicles, trains, ships, and air-
planes for which using current lithium-ion battery technology is currently infeasible
due to its relatively low energy density. If we assume that air travel in the next decade
returns to similar levels to 2019, then US jet fuel consumption would be about 100
billion liters per year or roughly one-fifth of the motor gasoline demand, which if met
by renewable hydrocarbons still places a significant demand on electrical generation,
40% of the current US level, if all of this were to be provided by CO2 conversion.

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


240 Matthew Realff

Biobased hydrocarbons could also provide significant fractions of the required jet
and diesel fuels, without taxing electrical energy generation but with the caveat that
the land area required is substantial due to the low efficiency of photosynthesis. For
example, at 329 liters of fuel per dry ton of wood, as reported in (Jones et al., 2016),
this would require 300 million dry tons of wood/year to meet the aviation fuel demand.
The yield of dry woody biomass per acre has been estimated to be, 4 tons/acre/year so
this would require 75 million acres, or roughly 10% of all forested lands in the United
States (USDA, n.d.).

7.7 Conclusion

Renewable hydrocarbons manufactured either from CO2 or from biomass are both a
unique opportunity and challenge. They present a low disruption pathway to decar-
bonizing transportation, requiring very little or no modification of consumer behavior.
They require enormous capital expenditure to implement at scale. In the case of CO2
conversion, the requirements for renewable energy are truly enormous, equivalent
to half the US electricity production today to meet US demands for aviation fuels
and doubling the production to meet gasoline demands. For biomass-based renewa-
ble fuels, the land areas required are very large, 50% or more of US Forests to meet
gasoline and aviation fuel demands. This would suggest that electrification of light
vehicles would be a significantly less energy intensive approach that requires differ-
ent investments and changing consumer behavior. Therefore, a hybrid approach that
utilizes biomass or CO2 for the parts of fuel use that are very difficult to electrify, such
as aviation or heavy duty trucking, and electrifying light duty vehicles seems more
viable. Whatever energy transition pathway we embark upon, we should not under-
estimate the costs. These will be manifested both in terms of the capital required for
new energy system components, such as renewable electricity, charging and vehicles,
hydrogen production and new processing trains for biomass or CO2, and in societal
adaptations required for mass adoption of electrified transportation.

References

Abanades, J. C., Rubin, E. S., Mazzotti, M., & Herzog, H. J. (2017). On the climate change
mitigation potential of CO2 conversion to fuels. Energy & Environmental Science, 10(12),
2491–99.
Ahlgren, S., & di Lucia, L. (2014). Indirect land use changes of biofuel production – a review of
modelling efforts and policy developments in the European Union. Biotechnology for Biofu-
els, 7(1), 1–10.
Albrecht, F. G., König, D. H., Baucks, N., & Dietrich, R.-U. (2017). A standardized methodol-
ogy for the techno-economic evaluation of alternative fuels – a case study. Fuel, 194, 511–26.
Argonne National Laboratory, The Greenhouse Gases, Regulated Emissions, and Energy Use in
Technologies Model. (2021). https://greet.es.anl.gov/

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


Liquid Fuel Synthesis 241

Artz, J., Müller, T. E., Thenert, K., Kleinekorte, J., Meys, R., Sternberg, A., Bardow, A., & Leit-
ner, W. (2018). Sustainable conversion of carbon dioxide: An integrated review of catalysis
and life cycle assessment. Chemical Reviews, 118(2), 434–504.
Becattini, V., Gabrielli, P., & Mazzotti, M. (2021). Role of carbon capture, storage, and utiliza-
tion to enable a net-zero-CO2-emissions aviation sector. Industrial & Engineering Chemis-
try Research, 60(18), 6848–62.
Bokinge, P., Heyne, S., & Harvey, S. (2020). Renewable OME from biomass and electricity –
evaluating carbon footprint and energy performance. Energy Science & Engineering, 8(7),
2587–98.
Bongartz, D., Doré, L., Eichler, K., Grube, T., Heuser, B., Hombach, L. E., Robinius, M.,
Pischinger, S., Stolten, D., & Walther, G. (2018). Comparison of light-duty transportation fuels
produced from renewable hydrogen and green carbon dioxide. Applied Energy, 231, 757–67.
Broch, A., Hoekman, S. K., & Unnasch, S. (2013). A review of variability in indirect land use
change assessment and modeling in biofuel policy. Environmental Science & Policy, 29,
147–57.
Brynolf, S., Taljegard, M., Grahn, M., & Hansson, J. (2018). Electrofuels for the transport
sector: A review of production costs. Renewable and Sustainable Energy Reviews, 81,
1887–905.
Burre, J., Bongartz, D., & Mitsos, A. (2019). Production of oxymethylene dimethyl ethers from
hydrogen and carbon dioxide – part II: Modeling and analysis for OME3–5. Industrial &
Engineering Chemistry Research, 58(14), 5567–78.
Cafferty, K. G., Muth Jr, D. J., Jacobson, J. J., & Bryden, K. M. (2013). Model based biomass
system design of feedstock supply systems for bioenergy production. International Design
Engineering Technical Conferences and Computers and Information in Engineering Confer-
ence, 55867, V02BT02A023.
Davis, S. J., Lewis, N. S., Shaner, M., Aggarwal, S., Arent, D., Azevedo, I. L., Benson, S. M.,
Bradley, T., Brouwer, J., & Chiang, Y.-M. et al. (2018). Net-zero emissions energy systems.
Science, 360(6396), eaas9793.
de Jong, S., Antonissen, K., Hoefnagels, R., Lonza, L., Wang, M., Faaij, A., & Junginger, M.
(2017). Life-cycle analysis of greenhouse gas emissions from renewable jet fuel production.
Biotechnology for Biofuels, 10(1), 1–18.
Detz, R. J., Reek, J. N. H., & van der Zwaan, B. C. C. (2018). The future of solar fuels: When
could they become competitive? Energy & Environmental Science, 11(7), 1653–69.
Deutz, S., & Bardow, A. (2021). Life-cycle assessment of an industrial direct air capture process
based on temperature – vacuum swing adsorption. Nature Energy, 6(2), 203–13.
Doliente, S. S., Narayan, A., Tapia, J. F. D., Samsatli, N. J., Zhao, Y., & Samsatli, S. (2020).
Bio-aviation fuel: A comprehensive review and analysis of the supply chain components.
Frontiers in Energy Research, 8, 110.
Dutta, A., & Phillips, S. D. (2009). Thermochemical ethanol via direct gasification and mixed
alcohol synthesis of lignocellulosic biomass. National Renewable Energy Lab.
Dutta, A., Talmadge, M., Hensley, J., Worley, M., Dudgeon, D., Barton, D., Groendijk, P., Ferrari,
D., Stears, B., & Searcy, E. M. (2011). Process design and economics for conversion of ligno-
cellulosic biomass to ethanol: Thermochemical pathway by indirect gasification and mixed
alcohol synthesis. National Renewable Energy Lab.
EIA, How much carbon dioxide is produced per kilowatthour of U.S. electricity generation?
(n.d.). www.eia.gov/tools/faqs/faq.php?id=77&t=11

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


242 Matthew Realff

Elliott, D. C., Biller, P., Ross, A. B., Schmidt, A. J., & Jones, S. B. (2015). Hydrothermal lique-
faction of biomass: Developments from batch to continuous process. Bioresource Technol-
ogy, 178, 147–56.
Fivga, A., Speranza, L. G., Branco, C. M., Ouadi, M., & Hornung, A. (2019). A review on the cur-
rent state of the art for the production of advanced liquid biofuels. Aims Energy, 7(1), 46–76.
Gabrielli, P., Gazzani, M., & Mazzotti, M. (2020). The role of carbon capture and utilization,
carbon capture and storage, and biomass to enable a net-zero-CO2 emissions chemical indus-
try. Industrial & Engineering Chemistry Research, 59(15), 7033–45.
Hall, D., & Rao, K. (1998). Photosynthesis (6th ed.). Cambridge University Press.
Han, J., Elgowainy, A., Dunn, J. B., & Wang, M. Q. (2013). Life cycle analysis of fuel produc-
tion from fast pyrolysis of biomass. Bioresource Technology, 133, 421–28.
Han, J., Tao, L., & Wang, M. (2017). Well-to-wake analysis of ethanol-to-jet and sugar-to-jet
pathways. Biotechnology for Biofuels, 10(1), 1–15.
Hartley, D. S., Thompson, D. N., Hu, H., & Cai, H. (2018). Woody feedstock 2018 state of tech-
nology report. Idaho National Lab. (INL).
Held, M., Tönges, Y., Pélerin, D., Härtl, M., Wachtmeister, G., & Burger, J. (2019). On the ener-
getic efficiency of producing polyoxymethylene dimethyl ethers from CO2 using electrical
energy. Energy & Environmental Science, 12(3), 1019–34.
Herron, J. A., Kim, J., Upadhye, A. A., Huber, G. W., & Maravelias, C. T. (2015). A general
framework for the assessment of solar fuel technologies. Energy & Environmental Science,
8(1), 126–57.
Huber, G. W., Iborra, S., & Corma, A. (2006). Synthesis of transportation fuels from biomass:
Chemistry, catalysts, and engineering. Chemical Reviews, 106(9), 4044–98.
Humbird, D., Davis, R., Tao, L., Kinchin, C., Hsu, D., Aden, A., Schoen, P., Lukas, J., Olthof,
B., & Worley, M. (2011). Process design and economics for biochemical conversion of lig-
nocellulosic biomass to ethanol: dilute-acid pretreatment and enzymatic hydrolysis of corn
stover. National Renewable Energy Lab.
Huo, H., Wang, M., Bloyd, C., & Putsche, V. (2009). Life-cycle assessment of energy use and
greenhouse gas emissions of soybean-derived biodiesel and renewable fuels. Environmental
Science & Technology, 43(3), 750–56.
Jones, S. B., Meyer, P. A., Snowden-Swan, L. J., Padmaperuma, A. B., Tan, E., Dutta, A., Jacob-
son, J., & Cafferty, K. (2013). Process design and economics for the conversion of ligno-
cellulosic biomass to hydrocarbon fuels: Fast pyrolysis and hydrotreating bio-oil pathway.
Pacific Northwest National Lab.
Jones, S. B., Snowden-Swan, L. J., Meyer, P. A., Zacher, A. H., Olarte, M. V., Wang, H., &
Drennan, C. (2016). Fast pyrolysis and hydrotreating: 2015 state of technology R&D and
projections to 2017. Pacific Northwest National Lab.
Jones, S. B., Zhu, Y., Anderson, D. B., Hallen, R. T., Elliott, D. C., Schmidt, A. J., Albrecht, K.
O., Hart, T. R., Butcher, M. G., & Drennan, C. (2014). Process design and economics for the
conversion of algal biomass to hydrocarbons: Whole algae hydrothermal liquefaction and
upgrading. Pacific Northwest National Lab.
Keith, D. W., Holmes, G., Angelo, D. S., & Heidel, K. (2018). A process for capturing CO2 from
the atmosphere. Joule, 2(8), 1573–94.
Knoema, Brazil – Fuel ethanol consumption. (2021). https://knoema.com/atlas/Brazil/topics/
Energy/Renewables/Fuel-ethanol-consumption
Kulkarni, A. P., Hos, T., Landau, M. V., Fini, D., Giddey, S., & Herskowitz, M. (2021).
­Techno-economic analysis of a sustainable process for converting CO2 and H2O to feed-
stock for fuels and chemicals. Sustainable Energy & Fuels, 5(2), 486–500.

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


Liquid Fuel Synthesis 243

Langholtz, M. H., Stokes, B. J., & Eaton, L. M. (2016). 2016 billion-ton report: Advancing domes-
tic resources for a thriving bioeconomy, volume 1: Economic availability of feedstock. Oak
Ridge National Laboratory, Oak Ridge, Tennessee, Managed by UT-Battelle, LLC for the US
Department of Energy, 2016, 448.
Lynd, L. R., Liang, X., Biddy, M. J., Allee, A., Cai, H., Foust, T., Himmel, M. E., Laser, M. S.,
Wang, M., & Wyman, C. E. (2017). Cellulosic ethanol: Status and innovation. Current Opin-
ion in Biotechnology, 45, 202–11.
Marchese, M., Buffo, G., Santarelli, M., & Lanzini, A. (2021). CO2 from direct air capture as
carbon feedstock for Fischer-Tropsch chemicals and fuels: Energy and economic analysis.
Journal of CO2 Utilization, 46, 101487.
McQueen, N., Gomes, K. V., McCormick, C., Blumanthal, K., Pisciotta, M., & Wilcox, J.
(2021). A review of direct air capture (DAC): Scaling up commercial technologies and inno-
vating for the future. Progress in Energy, 3(3), 032001.
Michailos, S., McCord, S., Sick, V., Stokes, G., & Styring, P. (2019). Dimethyl ether synthesis
via captured CO2 hydrogenation within the power to liquids concept: A techno-economic
assessment. Energy Conversion and Management, 184, 262–76.
Müller, L. J., Kätelhön, A., Bringezu, S., McCoy, S., Suh, S., Edwards, R., Sick, V., Kaiser, S.,
Cuéllar-Franca, R., & el Khamlichi, A. (2020). The carbon footprint of the carbon feedstock
CO2. Energy & Environmental Science, 13(9), 2979–92.
NREL Life Cycle Assessment Harmonization. (n.d.). www.nrel.gov/analysis/life-cycle-­
assessment.html
Smil, V. (2008). Energy in nature and society: General energetics of complex systems. MIT Press.
Sutter, D., van der Spek, M., & Mazzotti, M. (2019). 110th anniversary: Evaluation of CO2-
based and CO2-free synthetic fuel systems using a net-zero-CO2-emission framework.
Industrial & Engineering Chemistry Research, 58(43), 19958–72.
Tan, E. C. D., Talmadge, M., Dutta, A., Hensley, J., Schaidle, J., Biddy, M., Humbird, D.,
Snowden-Swan, L. J., Ross, J., & Sexton, D. (2015). Process design and economics for the
conversion of lignocellulosic biomass to hydrocarbons via indirect liquefaction. Thermo-
chemical research pathway to high-octane gasoline blendstock through methanol/dimethyl
ether intermediates. National Renewable Energy Lab. Golden, CO (United States).
Trippe, F., Fröhling, M., Schultmann, F., Stahl, R., Henrich, E., & Dalai, A. (2013). Comprehen-
sive techno-economic assessment of dimethyl ether (DME) synthesis and Fischer–Tropsch
synthesis as alternative process steps within biomass-to-liquid production. Fuel Processing
Technology, 106, 577–86.
U.S. Department of Energy, Ethanol Fuel Basics. (2021a). https://afdc.energy.gov/fuels/etha-
nol_fuel_basics.html
U.S. Department of Energy, Flexible Fuel Vehicles. (2021b). https://afdc.energy.gov/vehicles/
flexible_fuel.html
U.S. Department of Energy Fuel Conversion Factors to Gasoline Gallon Equivalents. (n.d.).
https://afdc.energy.gov/fuels/equivalency_methodology.html
USDA, Land Use and Land Cover Estimates for the United States. (n.d.). www.ers.usda.gov/
about-ers/partnerships/strengthening-statistics-through-the-icars/land-use-and-land-cover-
estimates-for-the-united-states/
van der Giesen, C., Kleijn, R., & Kramer, G. J. (2014). Energy and climate impacts of producing
synthetic hydrocarbon fuels from CO2. Environmental Science & Technology, 48(12), 7111–21.
Wu, M., Wu, Y., & Wang, M. (2006). Energy and emission benefits of alternative transportation
liquid fuels derived from switchgrass: A fuel life cycle assessment. Biotechnology Progress,
22(4), 1012–24.

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press


244 Matthew Realff

Zang, G., Sun, P., Elgowainy, A. A., Bafana, A., & Wang, M. (2021). Performance and cost
analysis of liquid fuel production from H2 and CO2 based on the Fischer-Tropsch process.
Journal of CO2 Utilization, 46, 101459.
Zhu, Y., Jones, S. B., Schmidt, A. J., Billing, J. M., Thorson, M. R., Santosa, D. M., Hallen, R.
T., & Anderson, D. B. (2020). Algae/wood blends hydrothermal liquefaction and upgrading:
2019 state of technology. Pacific Northwest National Lab.
Zhu, Y., Tjokro Rahardjo, S. A., Valkenburt, C., Snowden-Swan, L. J., Jones, S. B., & Machi-
nal, M. A. (2011). Techno-economic analysis for the thermochemical conversion of biomass
to liquid fuels. Pacific Northwest National Lab.

https://doi.org/10.1017/9781009072366.010 Published online by Cambridge University Press

You might also like