Download as pdf or txt
Download as pdf or txt
You are on page 1of 20

sustainability

Review
Energy Return on Investment of Major Energy Carriers: Review
and Harmonization
David J. Murphy 1, *, Marco Raugei 2,3 , Michael Carbajales-Dale 4 and Brenda Rubio Estrada 1

1 Environmental Studies Department, St. Lawrence University, Canton, NY 13617, USA; [email protected]
2 School of Engineering, Computing and Mathematics, Oxford Brookes University, Wheatley,
Oxford OX33 1HX, UK; [email protected]
3 Center for Life Cycle Assessment, Columbia University, New York, NY 10027, USA
4 Environmental Engineering & Earth Sciences, Clemson University, Clemson, SC 29634, USA;
[email protected]
* Correspondence: [email protected]

Abstract: Net energy, that is, the energy remaining after accounting for the energy “cost” of extraction
and processing, is the “profit” energy used to support modern society. Energy Return on Investment
(EROI) is a popular metric to assess the profitability of energy extraction processes, with EROI > 1 indi-
cating that more energy is delivered to society than is used in the extraction process. Over the past
decade, EROI analysis in particular has grown in popularity, resulting in an increase in publications
in recent years. The lack of methodological consistency, however, among these papers has led to a
situation where inappropriate comparisons are being made across technologies. In this paper we
provide both a literature review and harmonization of EROI values to provide accurate comparisons
of EROIs across both thermal fuels and electricity producing technologies. Most importantly, the
authors advocate for the use of point-of-use EROIs rather than point-of-extraction EROIs as the
energy “cost” of the processes to get most thermal fuels from extraction to point of use drastically
Citation: Murphy, D.J.; Raugei, M.; lowers their EROI. The main results indicate that PV, wind and hydropower have EROIs at or above
Carbajales-Dale, M.; Rubio Estrada, B. ten while the EROIs for thermal fuels vary significantly, with that for petroleum oil notably below ten.
Energy Return on Investment of
Major Energy Carriers: Review and Keywords: energy return on investment; EROI; net energy; fossil fuels; electricity; renewable energy;
Harmonization. Sustainability 2022, harmonization
14, 7098. https://doi.org/10.3390/
su14127098

Academic Editor: Muyiwa


1. Introduction
S. Adaramola
Energy is used to power natural ecosystems as well as human societies, and it is an
Received: 6 May 2022
integral part of all economic productivity. Over the course of the industrial revolution,
Accepted: 8 June 2022
the proliferation of the use of fossil fuels led to massive increases in physical productivity
Published: 9 June 2022
within economies, leading to wealth accumulation, technological development, population
Publisher’s Note: MDPI stays neutral growth, and, in short, to the rise of modernity. Society is now faced with the aftermath of
with regard to jurisdictional claims in the past two hundred years of growth in the form of climate change, land-cover change,
published maps and institutional affil- biodiversity loss, and other environmental woes.
iations. The “energy transition”, that is, the transition from fossil fuels to low carbon and
renewable energy sources, is viewed by many as the means by which we can maintain
our technology-intensive way of life without the concomitant environmental damage
imposed by the utilization of fossil fuels. There is debate, however, about whether or
Copyright: © 2022 by the authors.
not renewable energy can replace the manifold functions of fossil fuels in current society.
Licensee MDPI, Basel, Switzerland.
One way in which academics frame this debate is around the concept of Energy Return
This article is an open access article
on Investment [1–4]. EROI is a measure of energy profitability of energy sources and
distributed under the terms and
technologies [5]. In general, society benefits from energy resources and technologies that
conditions of the Creative Commons
Attribution (CC BY) license (https://
are highly profitable, which provide energy to society at very little energy cost (i.e., a large
creativecommons.org/licenses/by/
proportion of net energy). On the other hand, societies that have sources of energy that
4.0/).

Sustainability 2022, 14, 7098. https://doi.org/10.3390/su14127098 https://www.mdpi.com/journal/sustainability


Sustainability 2022, 14, 7098 2 of 20

have low profitability (i.e., low EROI resources) tend to be constrained in their growth
potential, among other things.
The debate about whether renewable energy can provide enough net energy to society
to replace fossil fuels is beset with controversy. For instance, many papers indicate that
solar photovoltaic (PV) electricity has a similar or higher EROI than natural gas electric-
ity [6,7]; yet, there are others saying the opposite, i.e., that the EROI of PV electricity is
much lower [8,9]. The same occurs for other renewable energy technologies, too. The
contributions that EROI can make to the framing of the energy transition are thus limited
by the seeming inability to form consensus in the literature.
This controversy is not new, and a number of authors and even the International
Energy Agency have called for stricter guidelines on how EROI is calculated [10,11]. More
specifically, the argument has been made for EROI practitioners to adopt the more formal
methodological framework established by the life cycle assessment (LCA) community [12].
Despite these calls for more methodological rigor in EROI analyses, more consistent EROI
assessments still seem elusive.
This paper arises from this academic landscape: one in which the EROIs of most
major technologies are still debated and one in which the methods are often inconsistent.
Therefore, its goals are threefold: first, to perform a review of the literature to provide an
overview of the most recent estimates of the EROI ranges for various energy technologies
and energy carriers. Second, to harmonize those values so that they are comparable
across technologies. Third, to provide the data and recommendations to enable other net
energy researchers to harmonize future work in hopes to avoid spurious comparisons in
the literature.

2. Materials and Methods


2.1. EROI Definition
The most fundamental operational definition of EROI is listed by, among others,
Dale [13] as:
Gross Energy Output
EROI = (1)
∑ Energy Investments
By this formulation, the energy delivered by the technology is divided by the energy
invested to make such delivery possible (the “energy investment” is defined as the energy
used to harvest an energy resource from the environment and convert it into a usable energy
carrier). However, this equation is often manipulated by researchers to calculate values
based on both different data sources and different system boundaries and assessment goals.
One alternative operational definition of EROI entails the use of monetary values
to indirectly estimate energy investments. For instance, Court and Fizaine [14] use an
economic input-output framework to calculate EROI, and their equation is:

Eout,i
EROIi = (2)
Min,i EIi

where Min,i is the amount of money invested in global energy sector i, and EIi is the average
energy intensity of energy sector i, in exajoules per 1 USD. However, not only is this a much
more descriptive formulation than that provided by Equation (1), it is also a fundamentally
different calculation that opens the door to potential distortions, such as those which arise
due to the elasticity of the money-to-energy relationship or the fact that this value is really
a measure of the power flow, i.e., energy per unit time.
However, even foregoing the use of all economic calculations and sticking to strictly
physical units only as implied by Equation (1), the literature is still replete with incon-
sistent EROI estimates and mismatched comparisons which, over time, risk leading to
a devaluation of the very concept of EROI in general. Specifically, when reviewing the
literature, two main types of methodological inconsistencies emerge, as discussed in the
following sections.
Sustainability 2022, 14, 7098 3 of 20

2.2. Supply Chain Boundary Mismatch


A first kind of boundary mismatch occurs whenever a comparison is made between
energy carriers that are sampled at different stages of their supply chain and—most
importantly—which are not functionally equivalent [15]. An extreme—but unfortunately
not uncommon—example of this is when the EROI of crude oil at the well-head is compared
to the EROI of electricity entering the grid [16,17].
In the most general terms possible, a primary energy resource (PES) needs to be first
extracted from nature, and subsequently processed into a usable energy carrier (EC) and
transported to the end user. This is illustrated schematically in Figure 1.

Inv 1 Inv 2

Primary
energy
resource,
renewable
Preparation
Primary Energy
Extraction &
energy Distribution carrier

Primary energy
resource, non-
renewable

POINT OF EXTRACTION
POINT OF USE

Figure 1. Streamlined energy systems diagram of the exploitation of a primary energy resource
(PES) for the production of a useful energy carrier (EC). Inv1 = energy investment for resource
extraction (E); Inv2 = energy investment for resource processing and delivery (P&D). S = energy sink
(thermodynamic losses). Energy system diagram following the symbolic conventions introduced by
Odum [18].

Much of the early EROI literature has traditionally focused on the analysis of energy
resources at point of extraction [17,19], implicitly assuming that such initial stage of the
supply chain would always dominate the final EROI ranking of the various types of
energy that are ultimately delivered to society. However, more recently the careful analysis
of a number of fossil fuel supply chains has shown that in the real world the energy
investments for processing the extracted raw resources into usable energy carriers and for
their subsequent transportation to the end user (i.e., Inv2 in Figure 1) are often larger than
the initial energy investment for resource extraction (i.e., Inv1 in Figure 1) [15,20–25].
As the energy inputs to the EROI equation derive more and more from post-extraction
processes, the comparison of “crude oil at the well head” with any downstream energy
carrier, such as electricity, becomes increasingly misleading. What has become increasingly
important in the recent literature is the value of energy at the point of use, after accounting
for all of the energy processing inputs, since these are the energy carriers delivered to the
end user to perform actual work in society.
Sustainability 2022, 14, 7098 4 of 20

The importance of methodological consistency is exemplified in both Table 1 and


Figure 2, which report a streamlined EROI analysis of three fictional primary energy
resources (PES1, 2 and 3), “at point of extraction” and “at point of use”, respectively. Given
the non-linear nature of the EROI calculation (see Equation (1)), the processing energy
inputs that are often excluded in fossil fuel EROIs dramatically change the resulting EROIs.
For example, the PES1 has an EROI at point of extraction of 100 while the corresponding
EROI for PES3 is only 25, which might lead one to conclude that PES1 is four times better
than PES 3. Yet, if the same energy is assumed to be invested in the processing of all PES1,
PES2, and PES3 (in this case nine units), then the EROIs at the point of use are all very
similar, i.e., EROI for PES1 at point of use is 10 while that for PES3 is 7.7. In sum, an initial
EROI difference at the point of extraction (i.e., a factor of four) is reduced to being only
a very marginal difference at the point of use. Moeller and Murphy [26] reported similar
declines in EROI from extraction to point of use, reporting a point of extraction EROI
of 40 for natural gas production via hydraulic fracturing in Pennsylvania, and an EROI
of 10 when that gas is delivered to the grid as electricity. Brandt [20] calculated historical
EROI trends at point of extraction and point of use for domestic oil extraction in California,
finding a similar trend. Additionally, Raugei and Leccisi [22] and Raugei et al. [23] traced
the diminishing EROI of fossil fuels along their whole supply chains, respectively to the
UK and to Chile.

Table 1. Calculation of EROI “at point of extraction” vs. EROI “at point of use” for three fictional
primary energy resources, highlighting the strong non-linearity in the relation between the two.

EROI
EROI
Energy “at Point of
Inv1 PE Inv2 EC “at Point of Use”
Resource Extraction”
= EC/(Inv1 + Inv2 )
= PE/Inv1
PES1 1 100 100 9 100 10
PES2 2 100 50 9 100 9.1
PES3 4 100 25 9 100 7.7

Resource Harmonization
PES 1 None
POU
PES 2 None
POU
PES 3 None
POU
0.5 0.7 1 1.5 2 3 4 5 7 10 15 20 30 40 50 70 100 150 200 300
EROI

Figure 2. EROI “at point of extraction” vs. EROI “at point of use” for three fictional primary energy
resources, highlighting the strong non-linearity in the relation between the two.

Two clear take-home messages emerge from this simple exercise: (1) comparing EROI
values calculated at different stages of the supply chain (i.e., EROI “at point of extraction”
vs. EROI “at point of use”) is methodologically unsound and results in inconsistent “apples-
to-oranges” comparisons that are devoid of any real significance; and (2) possibly even
more importantly, potentially large differences in EROI values calculated “at point of
extraction” can often be misleading, as necessary investments to convert the raw resources
into usable energy carriers at point of use often negate large EROI differences at the “point
of extraction”.

2.3. Temporal Boundary Mismatch


The second main type of inconsistency is that of temporal mismatch. Most conven-
tional EROI analyses adopt an integrative modelling approach whereby all the energy
inputs and outputs to/from the system that occur over its full life cycle are considered.
Sustainability 2022, 14, 7098 5 of 20

Conversely, some authors have calculated “EROI” values using a different methodology,
using the energy inputs and outputs for an energy system for one year. The two calculation
approaches produce only marginally different results for those energy systems where the
energy investment (i.e., the denominator of the EROI ratio) is almost synchronous with the
energy return (i.e., the numerator). For instance, the exploitation of fossil fuel resources
requires energy investments for extraction, refining and delivery of the fuels throughout
the entire life of the resources. Conversely, most major renewable energy technologies, such
as PV and wind, mostly only require energy investments in construction, and then entail
very low to negligible additional energy investments throughout the remaining stages of
the life of the technology. In this latter case, the two calculation approaches lead to very
different EROI results. For instance, an “EROI” ratio calculated for the first year of PV or
wind will often be very low, which can be misleading if inconsistently compared to other
EROIs calculated using the more conventional integrative approach. Of course, the “EROI”
of a PV system calculated as the ratio of the electricity produced during the second year of
operation to the sole maintenance energy investment taking place in that same year would
then be extremely large, and likewise misleading, but for some reason this latter type of
calculation is rarely, if ever, encountered in the literature.
In order to avoid the ambiguity that may arise from the use of the same acronym
“EROI” for what are essentially two different metrics calculated using different temporal
boundaries, many authors have argued, as we do here, that when “EROI” is calculated
relative to only one year of operation, it should more accurately be referred to as Power
Return On Investment (PROI), because it in fact measures the ratio of two power flows (i.e.,
flows of energy per year) [12,13,27]. Despite such attempts at more rigorous definitions,
however, a number of recent influential works still conflate EROI with PROI (Brockway
et al. [28] and Court and Fizaine [2]).

2.4. Focusing on Net Energy and the “Cliff”


Net energy is defined as:

Net Energy = Gross Energy Output − ∑ Energy Investments (3)

Combining Equation (3) with Equation (1), one obtains that the mathematical relation
between EROI and net energy is:
  
1
Net Energy = Gross Energy Output 1 − (4)
EROI
 
Net Energy 1
NTG = = 1− (5)
Gross Energy Output EROI
Using Equation (5), the net-to-gross energy ratio (NTG) can be plotted against EROI,
resulting in the “net energy cliff” [29]—see Figure 3. Many authors have acknowledged the
importance of analyzing EROIs with respect to the net energy cliff by providing figures
that position EROI values on the cliff itself [15,22,28,29].
The relation of EROI and net energy is highly non-linear: an energy acquisition process
that has an EROI of 1 delivers 0% net energy, while one with an EROI of 2 already delivers
50% net energy, and so forth. At the other end of the scale, a technology that extracts energy
with an EROI of 10 will deliver 90% of its energy as net energy to society, and beyond that
any further increases in EROI will only produce comparatively marginal improvements in
the amount of net energy. In practical terms, what this means is that one needs to spend
much less time worrying about whether an EROI is 20, 30, 40 or even higher, but rather
simply assess whether or not it meets a given minimum acceptable EROI threshold.
There is considerable debate, of course, about what the threshold for “minimum
acceptable EROI” actually is. Authors have tried estimating the minimum EROI that will
provide enough net energy to sustain a modern society [1–3]. Not surprisingly, given the
nature of the net energy cliff, the “minimum EROIs” postulated in the literature generally
Sustainability 2022, 14, 7098 6 of 20

range from 3–10; in other words, the minimum EROI values are located along the portion
of the net energy cliff curve where the net energy delivered increases rapidly with EROI.
However, it should be acknowledged that setting any specific benchmark value for such
‘minimum’ EROI is intrinsically fraught with difficulties.

1.0
None None
POU None
0.9 POU
POU

0.8

0.7

0.6
Net Energy

0.5

0.4

0.3

0.2 Resource
PES 1
PES 2
0.1
PES 3

0.0
1 1.5 2 3 4 5 7 10 15 20 30 40 50 70 100 150 200 300

EROI

Figure 3. Net Energy Cliff diagram relating EROI and net energy expressed as proportion of the
Gross Energy Output that is delivered to society. Arrows shows how the estimates of EROI and net
energy change when extending from point-of-extraction to point-of-use.

Firstly, from a methodological point of view, the devil is in the details, and it has been
convincingly argued that the definitions of net energy given by Equations (3) and (4) are
only rigorously applicable if both the gross energy output and the energy investments are
measured by the same standard [30]. In other words, since the energy investments are
typically accounted for in terms of primary energy (i.e., in units of “oil equivalent”), the
gross energy output should also be measured in units of equivalent primary energy, if the
former are to be subtracted from the latter. For clarity, one can use the subscript “PE-eq” to
specify when this is done, i.e., (Gross Energy Output)PE-eq .
The most rigorous way to quantify such “primary energy equivalency” is to adopt the
replacement logic that is prevalent in LCA, whereby each unit of the output energy carrier
is assumed to be equivalent to X units of primary energy, where (1/X) is the overall energy
efficiency of the “average” supply chain of the energy carrier in question [11,31].
Such distinction was rarely made in the early EROI literature, which tended to focus
primarily on fossil fuels. Admittedly, for thermal fuel products like crude oil or coal, and
even refined oil fuels and gases, the (1/X) ratio is often sufficiently close to 1 to render
the numerical distinction between (Gross Energy Output) and (Gross Energy Output)PE-eq
inconsequential, in light of the inevitable uncertainties that these life-cycle calculations
entail. In fact, the global average value of (1/X) for all fossil fuels combined (oil + coal + gas)
can be estimated to be 0.96, by using the information in the latest IEA World balance Sankey
diagram [32].
Sustainability 2022, 14, 7098 7 of 20

Given the negligible effect that such methodologically rigorous “primary energy equiv-
alency” calculations would have on the numerical estimate of the EROI values for thermal
fuels at point of use (including also for all biofuels, which are functionally equivalent to
the corresponding refined fossil fuels that they are intended to replace), in the remainder
of this article all such EROI values are simply calculated as the straight ratio of the gross
energy output in the fuel itself to the energy investments (i.e., as per Equation (1)).
However, when instead the Gross Energy Output is provided in the form of a highly
processed energy carrier for which the average supply chain entails significant thermo-
dynamic losses (such as is the case for electricity, when the average grid mix comprises
thermal power plants, as it almost invariably does), (1/X) may be significantly lower than 1
(e.g., it is often close to 0.3 for grid mixes dominated by coal and gas electricity).
For the specific case of electricity, therefore, in the recent literature an alternative
definition of EROI has emerged (Equation (6)), which makes the distinction between (Gross
Energy Output) and (Gross Energy Output)PE-eq explicit [10,11]:

( Gross Electricity Output) PE−eq ( Gross Electricity Output)/ηG


EROIPE−eq = = (6)
∑ Energy Investments ∑ Energy Investments

where ηG is the life-cycle efficiency of the grid mix.


Of course, the logical corollary to the argument made above, about the need to calculate
Net Energy in terms of primary energy equivalents, is that whenever any EROI value for
electricity is discussed in relation to the concept of net energy, and more specifically when
such value is positioned on the “net energy cliff”, such EROI value should always be
calculated according to Equation (6) [22,33,34].
In the remainder of this article, then, all EROI values for electricity will be harmonized
to EROIPE-eq calculated according to Equation (6) above, and subject to a sensitivity analysis
on the value of ηG (which may vary significantly depending on the proportion of thermal
vs. renewable energy resources used to generate electricity).
One methodological complication of the alternative definition of the EROI of electricity
as given in Equation (6) is that it is no longer an absolute indicator of the energy perfor-
mance of the technology being analyzed, but instead it becomes a relative indicator of its
performance vs. that of the average grid mix into which it is assumed to be embedded. This
has important implications also in terms of how the resulting EROIPE-eq trends over time
are to be interpreted, as discussed in some of the literature [34–36]. In simple terms, the
value EROIPE-eq becomes closer and closer to the “straight” EROI (calculated as the ratio
of the output electricity to the investments), as the primary-to-electric energy conversion
efficiency of the grid mix as a whole improves. This is consistent with the replacement
logic that underpins the definition of “primary energy equivalent”; in fact, asymptotically,
if a grid mix achieved ηG = 1, then one unit of electricity would become equivalent to one
unit of primary energy.
A closely related important consideration, though, is that any assumed minimum
EROI threshold always implicitly rests on an assumed average efficiency for the down-
stream processes in which the various energy carriers are used. Historically, the most
commonly used energy carriers have been thermal fuels, whose conversion into useful
work is severely constrained by the Carnot ratio (ηmax = 1 − TC /TH ). However, in the
coming decades, “a massive cross-sector electrification and a concomitant shift away from
thermal processes [ . . . ] may open the door to achieving the required services with much
lower demand for primary energy, which in turn entails that a significantly lower EROI
than previously assumed may suffice” [37]. In other words, all else being equal, the reduced
reliance on thermal power plants in electricity grids will tend to increase ηG , and therefore
decrease the EROIPE-eq of electricity, but at the same time, the resulting lower EROIPE-eq
will still suffice to clear the correspondingly reduced “minimum EROI” threshold required
for the support of an increasingly electrified society.
Sustainability 2022, 14, 7098 8 of 20

2.5. Literature Review


The literature review targeted five major academic databases: Science Direct, SCOPUS,
Springer, JSTOR, and Google Scholar. The search was restricted to the years 2017 to
2020 with the twin intent to capture the latest trends in terms of which energy technologies
attract the most attention, and to exclude potentially obsolete results for those technologies
that are still undergoing rapid development. The search keywords used were: Energy
Return on Investment, EROI, Net Energy, and Net Energy Return Ratio. Overall, the search
returned 113 papers in total. Each returned paper was added to a master database in
Microsoft Excel that recorded the full bibliographic information, the resource type being
analyzed, and the published EROI for each energy resource type. A manual screening
process was then employed to discard those papers which did not report original EROI
values or which were found to be critically lacking in transparency; more detail about
the results of the screening process are given in Section 3.1. One exception to this search
was made, and that is the inclusion of Leccisi and Fthenakis [36], which was published
in 2021 and was included to represent the most recent values for PV, given the extremely
rapid pace of development for this technology.

2.5.1. Literature on EROI of Thermal Fuels


There are a range of resources and products in the oil industry, so to limit spurious
comparisons, this review was limited to conventional and shale oil production, inclusive of
both on-shore and off-shore production [23,38–46]. Studies that calculated oil and gas EROI
together were excluded since the individual values could not be separated, e.g., Brockway
et al. [28]. EROI estimates for oil sands operations were also included; it is noteworthy that
these represent a fundamentally different resource base and production process [47].
Ten papers made estimates of the EROI of natural gas. Estimates from both conven-
tional and shale gas were included in this analysis [14,27,28,39,41–43,46,48,49].
Data collected for coal EROI estimates come from nine papers, which include estimates for
both coal at point of extraction and coal-fired electricity generation [14,22,27,28,39,44,46,48–50].
Other papers analyze coal seam gas and coal-to-liquids technologies [43,51].
The literature on the EROI of biodiesel includes seven papers with estimates for differ-
ent feedstocks, including jojoba [52]; waste cooking oil [53,54]; soybean, palm, microalgae,
soybean, animal fat, palm oil [55]; African palm, pinion, bovine and swine fat [56]; photoau-
totrophic algae, hybrid biofuel [57]; and agroforestry and first-generation soybean with and
without co-products [58]. Three different papers presented EROI estimates for bioethanol.
Different feedstocks for the production of bioethanol include sugarcane, corn, wood [56];
almond shells [59]; and corn grain [55]. One paper provided EROI estimates for biogas [60],
and one for wood chips [23].

2.5.2. Literature on EROI of Electricity


Three research papers included EROI estimates from pressurized water nuclear
reactors [46,49,50].
The literature search returned three papers with estimates for the EROI of bioenergy
with carbon capture and sequestration (BECCS) systems. The feedstocks used in the papers
analyzed herein include microalgae, biodegradable waste, lignocellulosic biomass, wheat,
switchgrass, miscanthus, and willow [61–63].
There were four estimates of the EROI of geothermal energy sources from four separate
analyses. These analyses were limited to geothermal systems that produce electricity, and
do not include ground-source heat pump systems which are an often confused with the
former, but are an altogether different technology [46,48–50].
There were five papers that made estimates of the EROI of hydropower. Literature
collected included only conventional dam and run-of-river hydropower production, and
excluded oceanic power and hydrothermal liquefaction [46,48–50,64].
Eight papers estimated EROI values for solar PVs, encompassing mono-silicon, poly-silicon,
cadmium-telluride (CdTe), and copper indium gallium di-selenide (CIGS) [34,36,46,49,50,65–67].
Sustainability 2022, 14, 7098 9 of 20

Estimates for perovskite PVs were excluded since these technologies are not yet commer-
cially viable and much uncertainty remains on their durability.
One paper provided an EROI estimate for concentrating solar power (CSP) [35].
Ten studies made estimates of the EROI of wind power, both from conventional on-
shore and off-shore wind turbines [46,49,50,64,66,68–72]. Smaller (<1 MW) micro-wind
turbines were not analyzed.

2.6. EROI Harmonization


EROI values were often reported for each study for a number of different technolo-
gies. The first part of our harmonization process separated these energy technologies and
resources into two categories: (1) thermal fuels at the point of use, and (2) electricity at the
point of use. Thermal fuels at the point of use include oil fuels, natural gas, coal, biodiesel,
bioethanol and biogas. Electricity at the point of use includes natural gas electricity, coal-
fired electricity, nuclear electricity, biomass electricity (with and without carbon capture
and sequestration), geothermal electricity, hydropower, solar photovoltaics, concentrating
solar power and wind power.

2.6.1. Harmonization of EROI Values of Thermal Fuels


EROI values from the studies were made consistent with respect to the process chain
boundary, as discussed above in Section 2.2. Specifically, data was taken from the life-
cycle database Ecoinvent (v3.7 and v3.8) [73] to provide information on inputs to all the
individual supply chain processes that follow resource extraction (namely: preparation,
transmission, refining, purification, and distribution), as depicted in Figure 4 below. For
the specific case of oil-derived fuels, given the plurality of fuels that are simultaneously
produced at the refining stage, the decision was made to focus on petrol (i.e., gasoline)
as a representative case in point (energy investment allocation for refining was made on
the basis of the exergy content of the output fuels, so the results essentially hold for all
other co-products, too). For the coal supply chain, hard coal (as opposed to brown coal or
lignite) was assumed. Additionally, given the large variability in transport distances, the
associated investments for transmission and distribution were estimated on the basis of
global average distances for each fuel supply chain as reported by Ecoinvent.
In Table 2, data are presented on energy investments to each of the post-extraction unit
processes in Figure 4, expressed as % relative to the energy in the output fuel. Additionally,
Table 2 also reports the cumulative post-extraction energy investments in the supply chain
up to and including that process, and the associated maximum EROI at that point in the
chain (i.e., the EROI that would result at that point of the chain as a consequence of the
post-extraction energy investments only, even assuming an initial infinite EROI at point of
extraction). These data were used to harmonize results from studies that presented EROI
results at different points in the chain by adding in information from ‘missing’ processes
up to the point of use (POU). More detailed information with full calculations and links to
specific datasets from Ecoinvent can be found in the Supplementary Materials.
By including this information in the paper, any analyst in the net energy community
can, for example, plug their own EROI values at point of extraction in the appropriate cell
in the spreadsheet, and then quickly find a standardized value for the corresponding EROI
at point of use.
Note also that, even ignoring the extraction process completely, the downstream
processing chain sets an upper limit to the EROI value of the product at the point of
use. For example, bioethanol from maize (corn) would have an EROI of 1.6 even without
including any energy investments for crop production. This is due mainly to the enormous
energy penalty (over 60%) in the refining process. This can be contrasted to the refining
process for sugarcane, which requires an investment equivalent to only 2% of the energy
in the fuel. Additionally, it is important to point out that, irrespective of their initial
EROI at point of extraction, all conventional fuels derived from fossil resources end up
having a maximum EROI at point of use well below 10. It bears reiterating that this is
Sustainability 2022, 14, 7098 10 of 20

the sole consequence of the multiple unavoidable energy investments that are required
post-extraction, along the supply chain, to convert the “raw” fossil resources (e.g., crude
oil) into usable fuels at point of use (e.g., petrol or diesel at the pump, or heavy fuel oil at
point of delivery). While these findings may be surprising and perhaps counterintuitive to
some, they are actually in perfect alignment with a recent high-level study that used IEA
data and extended multi-regional input–output tables to estimate EROI at point of use for
all fossil fuels produced globally [28].

Inv 1 Inv 2 Inv 3 Inv 4 Inv 5 Inv 6


v

Primary
energy
resource,
renewable
Extraction Preparation Transmission Refining Purification Distribution
(1) (2) (3) (4) (5) (6)

Primary energy
resource, non-
renewable

POINT OF EXTRACTION
POINT OF USE

Figure 4. Process chain for thermal fuels from extraction (1) to point-of-use (6).

Table 2. For each stage (i) of the supply chain for each thermal fuel beyond extraction, the following
values are reported: energy investment required at that stage (Invi ), cumulative investment in the
upstream chain up to that stage, excluding the investment for extraction (∑2i Inv j ), and maximum
EROI at that stage (EROIi,MAX ), disregarding the investment for extraction (i.e., assuming infinite
EROI at point of extraction). Data taken from Ecoinvent v3.7 and v3.8 [73]. All investments are
expressed as % relative to the final energy “return” (i.e., the net available energy in the output fuel at
point of use). A value of 0 means that that specific supply chain stage does not apply to that fuel.

Supply
(2) Preparation (3) Transmission (4) Refining (5) Purification (6) Distribution
Chain Stage
2 3 4 5 6
Fuel Inv2 ∑2 Invj EROI2,MAX Inv3 ∑2 Invj EROI3,MAX Inv4 ∑2 Invj EROI4,MAX Inv5 ∑2 Invj EROI5,MAX Inv6 ∑2 Invj EROI6,MAX
Oil 0 0 ∞ 1.5% 1.5% 67 8.9% 10.4% 9.6 0 0 9.6 1.1% 11.5% 8.7
Gas 0 0 ∞ 7.7% 7.7% 13 0 7.7% 13 0 7.7% 13 10.2% 17.9% 5.6
Coal 4.2% 4.2% 24 5.6% 9.8% 10 0 9.8% 10 0 9.8% 10 0 9.8% 10
Bioethanol
0 0 ∞ 0 0 ∞ 61% 61% 1.7 2.2% 62.6% 1.6 1.5% 64.1% 1.6
(Maize)
Bioethanol
0 0 ∞ 0 0 ∞ 2.5% 2.5% 39 2.2% 4.7% 21 0 4.7% 21
(Sugarcane)
Bioethanol 0 0 ∞ 0 0 ∞ 33.6% 33.6% 3.0 2.2% 35.8% 2.8 0 35.8% 2.8
Biogas 0 0 ∞ 0.2% 0.2% 420 0 0.2% 420 15.3% 15.5% 6.4 0.4% 15.9% 6.3
Biodiesel 3.3% 3.3% 31 1.7% 4.9% 20 5.4% 10.3% 10 0 10.3% 10 0 10.3% 10
Wood Pellets 51% 51% 2.0 0 51% 2.0 0 51% 2.0 0 51% 2.0 11.7% 63% 1.6

2.6.2. Harmonization of EROI Values of Electricity


As expected, the reviewed literature was inconsistent in terms of whether the reported
EROI values had been calculated as “straight” ratios of electricity output to primary
energy investment, or “weighted” ratios, where the electricity output at the numerator is
“converted” into primary energy equivalents, based on some assumed coefficient.
In order to harmonize all the collected EROI values for electricity from nuclear, BECCS,
hydro, geothermal, oceanic, PV, CSP, and wind, the following process was therefore employed:
Sustainability 2022, 14, 7098 11 of 20

(i) all “straight” EROI ratios were consistently multiplied by the same fixed 1/ηG value,
thereby calculating the corresponding EROIPE-eq , as per Equation (6). Given the critical
sensitivity associated to ηG (as discussed in Section 2.3), a sensitivity analysis was
carried out by repeating such calculation twice, first by setting ηG = 0.3 (representative
of deployment in most grid mixes dominated by conventional thermal generators),
and then by setting ηG = 0.7 (representative of deployment in a typical “decarbonized”
grid mix with a significant penetration of renewable energies [7]).
(ii) All “weighted” EROI ratios were first divided by whatever weighting factor had
originally been assumed by the authors, thereby essentially undoing any such weight-
ing and reverting to the corresponding “straight” EROIs where the numerator is
simply the electricity output. Then, the same procedure as for (i) was applied, so as
to once again arrive at two sets of EROIPE-eq values, respectively based on assumed
ηG = 0.3 and ηG = 0.7 life-cycle primary-to-electricity conversion factors.
Additionally, EROI values for electricity from combustion of coal, natural gas, biogas,
and biomass (wood chips) were also estimated, by leveraging the EROI values “at point of
use” for the respective thermal fuels that were obtained from the previous harmonization
process described in Section 2.6.1, and then multiplying those values by the respective
Ecoinvent-sourced power plant heat rates (i.e., 0.34 for coal, 0.47 for gas combined cycles,
0.35 for biogas, and 0.24 for biomass). Finally, the resulting EROI values were multiplied
by the same fixed 1/ηG values of 0.3 or 0.7, respectively, as described at point (i) above, to
convert them to “primary energy equivalent” (EROIPE-eq ). It is noted that, technically, this
process fails to account for the additional energy investment for power plant construction
and maintenance, but the data has shown that the latter is negligible for large thermal power
plants when such investment is spread out over their long service life. For the specific
case of gas-fired electricity, the more modern and efficient combined-cycle operation was
assumed; additionally, the energy investment for gas distribution (stage 6 in Table 2)
was omitted.

3. Results
The main results of our literature search and harmonization analysis are (in no partic-
ular order, and with detailed explanations in the following sections):
- A total of 113 papers were found reporting EROI values, but, after screening them, the
harmonization used only 31 papers.
- Most thermal fuels, including biofuel, oil, and natural gas have EROIs well below
10 after accounting for the entire production chain to the point-of-use.
- EROIs from electricity production from hydro, wind, and PV are all at or above 10,
once they are consistently expressed as “primary energy equivalent” (EROIPE-eq ).

3.1. Literature Screening


Out of the total 113 papers, 84 reported EROI values for the energy resources listed in
Table 3. Of the remaining 29 papers, many were EROIs of technologies that fell outside the
intended scope of this paper (e.g., EROI of agroecological systems). The remaining 84 EROI
values were then screened to remove duplicates, i.e., EROI values that were in some way
derived from previous EROI values calculated in other papers. The post-screening tally
was reduced substantially to only 37 papers across all technologies (Table 3).
Sustainability 2022, 14, 7098 12 of 20

Table 3. Number of papers returned by the literature search, per resource type. The total number of
papers in this table is more than that reported in Table 2 because some papers estimated EROI values
for more than one technology or resource. BECCS = bioenergy with carbon capture and sequestration;
CSG = coal seam gas; CTL = coal to liquids; LNG = liquefied natural gas.

Papers by Resource Type Initial Tally Post-Screening Tally


Thermal fuels
Biofuels (including biodiesel, bioethanol, biogas) 8 3
Coal (including CSG and CTL) 9 2
Natural Gas (including shale gas and LNG) 10 5
Oil (including Oil Sands) 11 9
Electricity
BECCS 3 2
Biogas 4 0
Concentrated Solar Power 2 1
Geothermal 5 2
Hydropower 7 2
Oceanic 1 1
Nuclear Power 3 1
Photovoltaics 11 4
Wind Power 10 5

3.2. Harmonization Analysis


3.2.1. EROIs of Thermal Fuels at Point of Use
Using data from Table 2, EROI values from the different studies were compared on
both an ‘as-is’ (harmonization = None) basis and a harmonized (Harmonization = POU)
basis. These results are presented in Figure 5. Note that the horizontal axis for EROI is on a
logarithmic scale. The first thing to note is that there is large variability in results, but that
most of the energy products have EROI values below 10, when taken to POU. This is true
of both biofuels and, importantly, also of oil, coal, and gas.
For gaseous fuels, natural gas has the highest EROI. Shale gas EROI estimates fell from
83 at point of extraction to 5.2 at POU due mainly to the large energy investments in the
transmission and distribution stages. Many of the bio-based gas products suffer a large
penalty (15%) in the biogas-to-biomethane purification process.
For liquid fuels, none of the conventional oil products have an EROI above 10, and
together the estimates for oil have a median EROI value of 4.2 when harmonized to POU.
All other estimates for bioethanol, biodiesel, and petrol from oil sands have harmonized
EROIs below 5.
For solid fuels, hard coal has a harmonized EROI of 8.8. The very high value (32) for
solid biomass deserves special discussion, as this is only for the specific case of locally-
sourced woodchips, which entail a comparatively simple and low-energy intensive supply
chain. It should, however, be noted that the EROI of solid biomass fuels varies significantly
for other fuels at point of use that require more energy-intensive supply chains (e.g., the
maximum EROI for wood pellets would be only 1.6, as per Table 2). As a consequence, the
single value for woodchips reported here in Figures 5 and 6 should not be misconstrued to
be representative of the wider spectrum of solid biomass fuels.
The ‘net energy cliff’ for thermal fuels at POU is presented in Figure 6. As mentioned
previously, many of the energy products have EROIs below 10, meaning that less than
90% of the net energy content in the fuel is available to society. As noted above, liquid fuels
(circles) tend to have lower EROIs than solid (squares) or gaseous fuels (asterisks).
Sustainability 2022, 14, 7098 13 of 20

Fuel Carrier Resource Harmonization


Gas CSG None
POU
Maize None
POU
Nat. gas None
POU

Gas
Reed Canary Grass None
POU
Shale gas None
POU
Sorghum None
POU
Diesel African Palm None
POU
Bovine Fat None
POU
Pinion None
POU
Porcine Fat None
POU
Ethanol Maize None
POU
Liquid

Sugarcane None
POU
Wood None
POU
Gasoline CTL None
POU
LNG None
POU
Oil None
POU
Oil sands None
POU
Coal Coal None
Solid

POU
Woodchips Wood None
POU
0.5 1 2 5 10 20 50 100 200
EROI

Figure 5. EROI values for thermal fuels, respectively, as originally published (Harmonization = “None”)
and post-harmonization at point of use (Harmonization = “POU”). CSG = coal seam gas; CTL = coal
to liquids; LNG = liquified natural gas. Note use of logarithmic scale on horizontal axis, for a more
meaningful representation of the significance of the relative differences in terms of net energy (cf.
Section 2.4). * = gaseouas fuel; o = liquid fuel;  = solid fuel.

1.0

0.9

0.8

Fuel
0.7 Gas
Liquid
Solid
0.6 Resource
Net Energy

African Palm
Bovine Fat
0.5 Coal
CSG
CTL
0.4 LNG
Maize
Nat. gas
Oil
0.3 Oil sands
Pinion
Porcine Fat
0.2 Reed Canary Grass
Shale gas
Sorghum
0.1 Sugarcane
Wood

0.0
1 1.5 2 3 4 5 7 10 15 20 30 40 50 70 100 150 200 300

EROI

Figure 6. Harmonized EROI values for thermal fuels at point of use, plotted against their correspond-
ing net-to-gross energy output ratios (“net energy cliff”). CSG = coal seam gas; CTL = coal to liquids;
LNG = liquified natural gas. Note use of logarithmic scale on horizontal axis, for a more meaningful
representation of the significance of the relative differences in terms of net energy (cf. Section 2.4).
Sustainability 2022, 14, 7098 14 of 20

3.2.2. EROIs of Electricity at Point of Use


The results of the harmonization calculations for the EROI of electricity, described in
Section 2.5.2, are reported in Figure 7. For each of the following primary energy sources
(PES): nuclear, BECCS, hydro, geothermal, oceanic, PV, CSP, and wind, three values are
shown, i.e., the original non-harmonized EROI (“None”), and the two harmonized ratios
(“PE eta = 0.3” and “PE eta = 0.7”, respectively). For thermal electricity from coal, gas
(combined cycle), biogas, and biomass (wood chips), only the harmonized EROIPE-eq ratios
are reported, since these were calculated from the harmonized EROIs at point of use for the
respective fuels (cf. Section 2.6.2). Methodologically consistent internal comparability is
thus made possible among each of the individual sets of harmonized EROIPE-eq values.
By analyzing the obtained results, the following general observations can be made:
1. Hydroelectricity exhibits the highest EROIPE-eq results by far. The second highest-
ranking group of technologies in terms of harmonized EROIPE-eq comprise: nuclear,
wind, and—in some cases—PVs (see point 2. below for caveats on the latter). CSP
and geothermal electricity can then be grouped together as the third “block” of
results in descending order of EROIPE-eq . Broadly speaking, all electricity generation
technologies listed thus far are characterized by harmonized EROIPE-eq values greater
than 10, when calculated assuming a primary energy to electricity life-cycle conversion
factor ηG = 0.3. Oceanic electricity straddles this symbolic EROIPE-eq = 10 line.
2. The EROIPE-eq values for PV electricity (and to a lesser extent also for geothermal
electricity) span a fairly wide range. This appears to be primarily due to intrinsic
differences in the assessed supply chains and technologies. Specifically, for the case of
PV, the technological differences among the various technologies (sc-Si, mc-Si, CdTe
and CIGS) are compounded by the large effect of variations in assumed solar irradia-
tion (from approximately 1000 kWh·m−2 ·yr−1 for northern latitudes e.g., Germany, to
over 2300 kWh·m−2 ·yr−1 for southern latitudes e.g., Chile). However, the deliberate
choice was made not to attempt any harmonization for the latter, since it represents
a real-world variable and not a methodological inconsistency per se. The important
take-home message in these cases is that it is unreasonable to expect to arrive at a
single value (or a very tight range of estimates) for the EROIPE-eq of these technologies,
due to the intrinsic variability ranges that characterize them.
3. The EROIPE-eq values for thermal electricity from the combustion of fossil fuels (coal
and natural gas) are both in the range of 10–12, when calculated using ηG = 0.3.
4. Thermal electricity from biogas and BECCS is characterized by comparatively low
EROIPE-eq values of 2–5, when calculated using ηG = 0.3. While these results may
appear to contradict some higher estimates in the previous literature, it seems likely
that in those earlier studies some of the supply chain investments identified in Table 2
may have been missed. For instance, Raugei et al. [23] caveated their results for
biomass- and biogas-fired electricity by stating that “EROI results for these technolo-
gies are affected by a larger margin of uncertainty, due to a combination of older
inventory data and (for biomass and biogas) possible inaccuracies in the modelling of
the feedstock supply chains”.
5. Finally, the calculated EROIPE-eq values for biomass-fired electricity using wood chips
is comparatively high at 16 (assuming ηG = 0.3). However, as discussed in Section 3.2.1
for wood chips as a fuel stock, this result is only valid for this particular biomass fuel,
whereas it would be considerably lower if a blend of woodchips and wood pellets
were employed instead (as is the case in the UK, for instance [22]).
Sustainability 2022, 14, 7098 15 of 20

PES Harmonization
BECCS None
PE eta=0.3
PE eta=0.7
Biogas None
PE eta=0.3
PE eta=0.7
Biomass (woodchips) None
PE eta=0.3
PE eta=0.7
Coal None
PE eta=0.3
PE eta=0.7
CSP None
PE eta=0.3
PE eta=0.7
Geothermal None
PE eta=0.3
PE eta=0.7
Hydro None
PE eta=0.3
PE eta=0.7
Nat Gas (CC) None
PE eta=0.3
PE eta=0.7
Nuclear None
PE eta=0.3
PE eta=0.7
Oceanic None
PE eta=0.3
PE eta=0.7
PV None
PE eta=0.3
PE eta=0.7
Wind None
PE eta=0.3
PE eta=0.7
0.5 1 2 5 10 20 50 100 200
EROI

Figure 7. EROI values for electricity, respectively, as originally published (Harmonization = “None”),
and post-harmonization in terms of equivalent primary energy output, respectively assuming deploy-
ment in a thermal-dominated electricity grid mix (Harmonization = “PE eta = 0.3”), and deployment
in a de-carbonized electricity grid mix (Harmonization = “PE eta = 0.7”). BECCS = bioenergy with
carbon capture and sequestration; CSP = concentrating solar power; PV = photovoltaics. Note use of
logarithmic scale on horizontal axis, for a more meaningful representation of the significance of the
relative differences in terms of net energy (cf. Section 2.4).

In more general terms, the systematically lower EROIPE-eq values for all technologies,
when calculated using ηG = 0.7 should not surprise nor be a reason for concern. This is
simply the consequence of assuming deployment in a grid mix that is itself on average
significantly more efficient at converting primary energy into electricity over its whole
life cycle. As discussed in Section 2.4, while the individual EROIPE-eq for all technologies
would be reduced in such conditions, at the same time it is reasonable to expect that, in the
future, the same widespread deployment of low-cost renewable energies that will lead to a
higher ηG = 0.7 in the first place will also enable a higher degree of electrification across
multiple sectors and end uses, thereby essentially lowering the “minimum EROI” threshold
to above that which a healthy societal energy metabolism may be sustained.
At present, the EROIPE-eq values obtained by setting ηG = 0.3 may still be considered
the more representative ones, as the use of thermal technologies to generate electricity
is still prevalent globally. These values were therefore selected to be reported vs. the
corresponding NTG ratios (i.e., superimposed on the “net energy cliff”) in Figure 8. This
latter figure allows a clearer visualization of which electricity generation technologies can be
expected to generate sufficient net energy over their life cycles. Once again, the results show
that most renewable technologies actually lead to NTG > 0.9, meaning that over 90% of the
equivalent primary energy returned by them remains available for societal uses other than
Sustainability 2022, 14, 7098 16 of 20

supporting the energy sector itself. Overall, this is a reassuring result that should put to
rest many often-voiced concerns about the net energy viability of non-conventional and
renewable electricity.

1.0

0.9

0.8

0.7

0.6
Net Energy

0.5
PES
BECCS
0.4
Biogas
Biomass (woodchips)
Coal
0.3 CSP
Geothermal
Hydro
0.2 Nat Gas (CC)
Nuclear
Oceanic
PV
0.1
Wind

0.0
1 1.5 2 3 4 5 7 10 15 20 30 40 50 70 100 150 200 300

-0.1111
EROI -0.1111

Figure 8. Harmonized EROI values for electricity (assuming deployment in a thermal-dominated


electricity grid mix, ηG = 0.3), plotted against their corresponding net-to-gross energy output ratios
(“net energy cliff”). BECCS = bioenergy with carbon capture and sequestration; CSP = concentrating
solar power; PV = photovoltaics. Note use of logarithmic scale on horizonal axis, for a more
meaningful representation of the significance of the relative differences in terms of net energy (cf.
Section 2.4).

4. Discussion and Conclusions


The analysis performed herein represents a much-needed update and harmonization of
the EROI literature, and it advances the conversation surrounding the viability of renewable
resources in the energy transition process. A common argument is that the EROIs from
renewable energy technologies are supposedly lower than those provided by fossil fuels,
and that transitioning to RE technologies would therefore result in a large loss in net energy.
The results of this analysis rebuke that sentiment, noting that the three most important
technologies for the energy transition—wind, PV, and hydropower—all have EROIs at
or above 10 (even when the output is weighted in terms of primary energy equivalent
assuming a future-proof life-cycle grid efficiency of ηG = 0.7, i.e., 1 unit of electricity per
1.4 units of primary energy). This means that greater than 90% of the energy produced by
these technologies is delivered to society as net energy.
Perhaps more interesting still, the EROIs from liquid fuels, including the EROI from
conventional oil production, are less than 10 once the costs of refining and delivery to
the point-of-use are included. Oil is widely considered the most important fuel for the
economy, used mostly in the transportation sector. This means that oil delivers less net
energy to society for each unit invested in extraction, refining, and delivery than PV or
wind. The transition to electric vehicles, according to these results, will actually increase
the amount of net energy delivered to society (even more so when considering the higher
efficiency of electrical power trains vs. internal combustion engines).
Sustainability 2022, 14, 7098 17 of 20

It is clear from these results that EROI estimates at the point of extraction can be wildly
misleading. As a case in point, even if crude oil were measured to have an EROI of 1000 or
more at the point of extraction, the corresponding EROI at the point of use, using global
average data for the energy “cost” of the process chain, would still only be a maximum
of 8.7. Furthermore, as the quality of oil, gas and coal continue to decline in the future, the
energy “cost” of the associated process chains will increase, further reducing the EROIs.
On the other hand, as the technologies used to harness renewable energy improve, the
corresponding EROIs will continue to increase in the future.
Finally, it is also important to observe that, in the future, a significant increase in
the penetration of renewable technologies into the electricity grid mixes will have to be
accompanied by a concomitant deployment of electrical storage, to compensate for the
intrinsic intermittency or renewable energy availability and ensure the continued real-time
matching of the supply and demand curves. However, detailed scenario analyses of the net
energy performance of even highly decarbonized grid mixes relying heavily on PVs, based
on high temporal resolution grid balancing algorithms rather than blunt assumptions,
indicate that the additional energy investment for electrochemical energy storage does not
significantly affect the overall EROIPE-eq of the resulting electricity mix [7,11].

Supplementary Materials: The following are available online at https://www.mdpi.com/article/10


.3390/su14127098/s1: “EROI_harmonization.xls”, providing detailed supply-chain energy invest-
ment calculations for selected thermal fuels, with references.
Author Contributions: Data curation, B.R.E., D.J.M., M.R. and M.C.-D.; Formal analysis, D.J.M., M.R.
and M.C.-D.; Writing—original draft, D.J.M., M.R. and M.C.-D. All authors have read and agreed to
the published version of the manuscript.
Funding: This research received no external funding.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Conflicts of Interest: The authors declare no conflict of interest.

References
1. Hall, C.A.S.; Balogh, S.; Murphy, D.J.R. What Is the Minimum EROI That a Sustainable Society Must Have? Energies 2009, 2,
25–47. [CrossRef]
2. Fizaine, F.; Court, V. Energy Expenditure, Economic Growth, and the Minimum EROI of Society. Energy Policy 2016, 95, 172–186.
[CrossRef]
3. Brandt, A.R. How Does Energy Resource Depletion Affect Prosperity? Mathematics of a Minimum Energy Return on Investment
(EROI). BioPhysical Econ. Resour. Qual. 2017, 2, 2. [CrossRef]
4. Lambert, J.G.; Hall, C.A.S.; Balogh, S.; Gupta, A.; Arnold, M. Energy, EROI and Quality of Life. Energy Policy 2014, 64, 153–167.
[CrossRef]
5. Hall, C.A.S.; Lavine, M.; Sloane, J. Efficiency of Energy Delivery Systems: Part 1 An Economic and Energy Analysis. Environ.
Manag. 1979, 3, 493–504. [CrossRef]
6. Bhandari, K.P.; Collier, J.M.; Ellingson, R.J.; Apul, D.S. Energy Payback Time (EPBT) and Energy Return on Energy Invested
(EROI) of Solar Photovoltaic Systems: A Systematic Review and Meta-Analysis. Renew. Sustain. Energy Rev. 2015, 47, 133–141.
[CrossRef]
7. Raugei, M.; Peluso, A.; Leccisi, E.; Fthenakis, V. Life-Cycle Carbon Emissions and Energy Return on Investment for 80% Domestic
Renewable Electricity with Battery Storage in California. Energies 2020, 13, 3934. [CrossRef]
8. Ferroni, F.; Hopkirk, R.J. Energy Return on Energy Invested (ERoEI) for Photovoltaic Solar Systems in Regions of Moderate
Insolation. Energy Policy 2016, 94, 336–344. [CrossRef]
9. Prieto, P.A.; Hall, C. Spain’s Photovoltaic Revolution: The Energy Return on Investment; Springer: New York, NY, USA, 2013.
10. Raugei, M.; Frischknecht, R.; Olson, C.; Sinha, P.; Heath, G. Methodological Guidelines on Net Energy Analysis of Photovoltaic
Electricity; International Energy Agency: Paris, France, 2016.
11. Raugei, M.; Frischknecht, R.; Olson, C.; Sinha, P.; Heath, G. Methodological Guidelines on Net Energy Analysis of Photovoltaic
Electricity, 2nd ed.; International Energy Agency: Paris, France, 2021.
12. Murphy, D.J.; Carbajales-Dale, M.; Moeller, D. Comparing Apples to Apples: Why the Net Energy Analysis Community Needs to
Adopt the Life-Cycle Analysis Framework. Energies 2016, 9, 917. [CrossRef]
Sustainability 2022, 14, 7098 18 of 20

13. Carbajales-Dale, M. When Is EROI Not EROI? BioPhysical Econ. Resour. Qual. 2019, 4, 16. [CrossRef]
14. Court, V.; Fizaine, F. Long-Term Estimates of the Energy-Return-on-Investment (EROI) of Coal, Oil, and Gas Global Productions.
Ecol. Econ. 2017, 138, 145–159. [CrossRef]
15. Raugei, M. Net Energy Analysis Must Not Compare Apples and Oranges. Nat. Energy 2019, 4, 86–88. [CrossRef]
16. Hall, C.A.S.; Day, J.W. Revisiting the Limits to Growth After Peak Oil. Am. Sci. 2009, 97, 230–237. [CrossRef]
17. Murphy, D.J.; Hall, C.A.S. Year in Review—EROI or Energy Return on (Energy) Invested. N. Y. Ann. Sci. 2010, 1185, 102–118.
[CrossRef]
18. Odum, H.T. Systems Ecology: An Introduction; John Wiley and Sons, Inc.: New York, NY, USA, 1983.
19. Cleveland, C.J. Energy Quality and Energy Surplus in the Extraction of Fossil Fuels in the U.S. Ecol. Econ. 1992, 6, 139–162.
[CrossRef]
20. Brandt, A.R. Oil Depletion and the Energy Efficiency of Oil Production: The Case of California. Sustainability 2011, 3, 1833–1854.
[CrossRef]
21. Rahman, M.M.; Canter, C.; Kumar, A. Well-to-Wheel Life Cycle Assessment of Transportation Fuels Derived from Different North
American Conventional Crudes. Appl. Energy 2015, 156, 159–173. [CrossRef]
22. Raugei, M.; Leccisi, E. A Comprehensive Assessment of the Energy Performance of the Full Range of Electricity Generation
Technologies Deployed in the United Kingdom. Energy Policy 2016, 90, 46–59. [CrossRef]
23. Raugei, M.; Leccisi, E.; Fthenakis, V.; Escobar Moragas, R.; Simsek, Y. Net Energy Analysis and Life Cycle Energy Assessment of
Electricity Supply in Chile: Present Status and Future Scenarios. Energy 2018, 162, 659–668. [CrossRef]
24. Aguirre-Villegas, H.A.; Benson, C.H. Case History of Environmental Impacts of an Indonesian Coal Supply Chain. J. Clean. Prod.
2017, 157, 47–56. [CrossRef]
25. Yáñez, E.; Ramírez, A.; Uribe, A.; Castillo, E.; Faaij, A. Unravelling the Potential of Energy Efficiency in the Colombian Oil
Industry. J. Clean. Prod. 2018, 176, 604–628. [CrossRef]
26. Moeller, D.; Murphy, D. Net Energy Analysis of Gas Production from the Marcellus Shale. BioPhysical Econ. Resour. Qual. 2016, 1, 5.
[CrossRef]
27. Raugei, M.; Leccisi, E.; Azzopardi, B.; Jones, C.; Gilbert, P.; Zhang, L.; Zhou, Y.; Mander, S.; Mancarella, P. A Multi-Disciplinary
Analysis of UK Grid Mix Scenarios with Large-Scale PV Deployment. Energy Policy 2018, 114, 51–62. [CrossRef]
28. Brockway, P.E.; Owen, A.; Brand-Correa, L.; Hardt, L. Estimation of Global Final-Stage Energy-Return-on-Investment for Fossil
Fuels with Comparison to Renewable Energy Sources. Nat. Energy 2019, 4, 612–621. [CrossRef]
29. Murphy, D.J. The Implications of the Declining Energy Return on Investment of Oil Production. Philos. Trans. R. Soc. A 2014, 372,
20130126. [CrossRef]
30. Averson, A.; Hertwich, E.G. More Caution Is Needed When Using Life Cycle Assessment to Determine Energy Return on
Investment (EROI). Energy Policy 2015, 76, 1.
31. Frischknecht, R.; Stolz, P.; Heath, G.; Raugei, M.; Sinha, P.; de Wild-Scholten, M.; Fthenakis, V.; Kim, H.C.; Alsema, E.; Held, M.
Methodology Guidelines on Life Cycle Assessment of Photovoltaic Electricity; PVPS Task 12; International Energy Agency (IEA): Paris,
France, 2020.
32. IEA. Sankey Diagram; IEA: Paris, France, 2018.
33. Murphy, D.; Raugei, M. The Energy Transition in New York: A Greenhouse Gas, Net Energy, and Life-Cycle Energy Analysis.
Energy Technol. 2020, 8, 1901026. [CrossRef]
34. Raugei, M.; Kamran, M.; Hutchinson, A. A Prospective Net Energy and Environmental Life-Cycle Assessment of the UK
Electricity Grid. Energy Technol. 2020, 13, 2207. [CrossRef]
35. Raugei, M. Energy Pay-Back Time: Methodological Caveats and Future Scenarios. Prog. Photovolt. Res. Appl. 2013, 21, 797–801.
[CrossRef]
36. Fthenakis, V.; Leccisi, E. Updated Sustainability Status of Crystalline Silicon-Based Photovoltaic Systems: Life-Cycle Energy and
Environmental Impact Reduction Trends. Prog. Photovolt. Res. Appl. 2021, 29, 1068–1077. [CrossRef]
37. Raugei, M. Energy Return on Investment: Setting the Record Straight. Joule 2019, 3, 1810–1811. [CrossRef]
38. Tripathi, V.S.; Brandt, A.R. Estimating Decades-Long Trends in Petroleum Field Energy Return on Investment (EROI) with an
Engineering-Based Model. PLoS ONE 2017, 12, e0171083. [CrossRef]
39. Feng, J.; Feng, L.; Wang, J. Analysis of Point-of-Use Energy Return on Investment and Net Energy Yields from China’s Conven-
tional Fossil Fuels. Energies 2018, 11, 313. [CrossRef]
40. Huang, C.; Gu, B.; Chen, Y.; Tan, X.; Feng, L. Energy Return on Energy, Carbon, and Water Investment in Oil and Gas Resource
Extraction: Methods and Applications to the Daqing and Shengli Oilfields. Energy Policy 2019, 134, 110979. [CrossRef]
41. Salehi, M.; Khajehpour, H.; Saboohi, Y. Extended Energy Return on Investment of Multiproduct Energy Systems. Energy 2020,
192, 116700. [CrossRef]
42. Chen, Y.; Feng, L.; Tang, S.; Wang, J.; Huang, C.; Höök, M. Extended-Exergy Based Energy Return on Investment Method and Its
Application to Shale Gas Extraction in China. J. Clean. Prod. 2020, 260, 120933. [CrossRef]
43. Kong, Z.; Lu, X.; Dong, X.; Jiang, Q.; Elbot, N. Re-Evaluation of Energy Return on Investment (EROI) for China’s Natural Gas
Imports Using an Integrative Approach. Energy Strategy Rev. 2018, 22, 179–187. [CrossRef]
44. Qu, J.-L.W.-X.F.B.-Y.F. A Review of Physical Supply and EROI of Fossil Fuels in China. Pet. Sci. 2017, 14, 806–821. [CrossRef]
Sustainability 2022, 14, 7098 19 of 20

45. Solé, J.; García-Olivares, A.; Turiel, A.; Ballabrera-Poy, J. Renewable Transitions and the Net Energy from Oil Liquids: A Scenarios
Study. Renew. Energy 2018, 116, 258–271. [CrossRef]
46. King, L.C.; Van Den Bergh, J.C.J.M. Implications of Net Energy-Return-on-Investment for a Low-Carbon Energy Transition. Nat.
Energy 2018, 3, 334–340. [CrossRef]
47. Wang, K.; Vredenburg, H.; Wang, J.; Xiong, Y.; Feng, L. Energy Return on Investment of Canadian Oil Sands Extraction from 2009
to 2015. Energies 2017, 10, 614. [CrossRef]
48. Walmsley, M.R.W.; Walmsley, T.G.; Atkins, M.J. Linking Greenhouse Gas Emissions Footprint and Energy Return on Investment
in Electricity Generation Planning. J. Clean. Prod. 2018, 200, 911–921. [CrossRef]
49. Kis, Z.; Pandya, N.; Koppelaar, R.H.E.M. Electricity Generation Technologies: Comparison of Materials Use, Energy Return on
Investment, Jobs Creation and CO2 Emissions Reduction. Energy Policy 2018, 120, 144–157. [CrossRef]
50. Walmsley, T.G.; Walmsley, M.R.W.; Varbanov, P.S.; Klemeš, J.J. Energy Ratio Analysis and Accounting for Renewable and
Non-Renewable Electricity Generation: A Review. Renew. Sustain. Energy Rev. 2018, 98, 328–345. [CrossRef]
51. Kong, Z.; Dong, X.; Jiang, Q. The Net Energy Impact of Substituting Imported Oil with Coal-to-Liquid in China. J. Clean. Prod.
2018, 198, 80–90. [CrossRef]
52. Sandouqa, A.; Al-Hamamre, Z. Energy Analysis of Biodiesel Production from Jojoba Seed Oil. Renew. Energy 2019, 130, 831–842.
[CrossRef]
53. Farid, M.A.A.; Roslan, A.M.; Hassan, M.A.; Hasan, M.Y.; Othman, M.R.; Shirai, Y. Net Energy and Techno-Economic Assessment
of Biodiesel Production from Waste Cooking Oil Using a Semi-Industrial Plant: A Malaysia Perspective. Sustain. Energy Technol.
Assess. 2020, 39, 100700. [CrossRef]
54. Barbera, E.; Naurzaliyev, R.; Asiedu, A.; Bertucco, A.; Resurreccion, E.P.; Kumar, S. Techno-Economic Analysis and Life-Cycle
Assessment of Jet Fuels Production from Waste Cooking Oil via in Situ Catalytic Transfer Hydrogenation. Renew. Energy 2020,
160, 428–449. [CrossRef]
55. Sales, E.A.; Ghirardi, M.L.; Jorquera, O. Subcritical Ethylic Biodiesel Production from Wet Animal Fat and Vegetable Oils: A Net
Energy Ratio Analysis. Energy Convers. Manag. 2017, 141, 216–223. [CrossRef]
56. Chiriboga, G.; Rosa, A.D.L.; Molina, C.; Velarde, S. Energy Return on Investment (EROI) and Life Cycle Analysis (LCA) of
Biofuels in Ecuador. Heliyon 2020, 6, e04213. [CrossRef]
57. Carneiro, M.L.N.M.; Pradelle, F.; Braga, S.L.; Gomes, M.S.P.; Martins, A.R.F.A.; Turkovics, F.; Pradelle, R.N.C. Potential of Biofuels
from Algae: Comparison with Fossil Fuels, Ethanol and Biodiesel in Europe and Brazil through Life Cycle Assessment (LCA).
Renew. Sustain. Energy Rev. 2017, 73, 632–653. [CrossRef]
58. Pragya, N.; Sharma, N.; Gowda, B. Biofuel from Oil-Rich Tree Seeds: Net Energy Ratio, Emissions Saving and Other Environmental
Impacts Associated with Agroforestry Practices in Hassan District of Karnataka, India. J. Clean. Prod. 2017, 164, 905–917.
[CrossRef]
59. Kaur, M.; Kumar, M.; Sachdeva, S.; Puri, S.K. An Efficient Multiphase Bioprocess for Enhancing the Renewable Energy Production
from Almond Shells. Energy Convers. Manag. 2020, 203, 112235. [CrossRef]
60. Krzystek, L.; Wajszczuk, K.; Pazera, A.; Matyka, M.; Slezak, R.; Ledakowicz, S. The Influence of Plant Cultivation Conditions on
Biogas Production: Energy Efficiency. Waste Biomass Valorization 2020, 11, 513–523. [CrossRef]
61. Cheng, F.; Porter, M.D.; Colosi, L.M. Is Hydrothermal Treatment Coupled with Carbon Capture and Storage an Energy-Producing
Negative Emissions Technology? Energy Convers. Manag. 2020, 203, 112252. [CrossRef]
62. Melara, A.J.; Singh, U.; Colosi, L.M. Is Aquatic Bioenergy with Carbon Capture and Storage a Sustainable Negative Emission
Technology? Insights from a Spatially Explicit Environmental Life-Cycle Assessment. Energy Convers. Manag. 2020, 224, 113300.
[CrossRef]
63. Fajardy, M.; Dowell, N.M. The Energy Return on Investment of BECCS: Is BECCS a Threat to Energy Security? Energy Environ.
Sci. 2018, 11, 1581–1594. [CrossRef]
64. Trainer, T. Estimating the EROI of Whole Systems for 100% Renewable Electricity Supply Capable of Dealing with Intermittency.
Energy Policy 2018, 119, 648–653. [CrossRef]
65. Wang, C.; Cheng, X.; Shuai, C.; Huang, F.; Zhang, P.; Zhou, M.; Li, R. Evaluation of Energy and Environmental Performances of
Solar Photovoltaic-Based Targeted Poverty Alleviation Plants in China. Energy Sustain. Dev. 2020, 56, 73–87. [CrossRef]
66. Dupont, E.; Koppelaar, R.; Jeanmart, H. Global Available Solar Energy under Physical and Energy Return on Investment
Constraints. Appl. Energy 2020, 257, 113968. [CrossRef]
67. Zhou, Z.; Carbajales-Dale, M. Assessing the Photovoltaic Technology Landscape: Efficiency and Energy Return on Investment
(EROI). Energy Environ. Sci. 2018, 11, 603–608. [CrossRef]
68. Feng, J.; Feng, L.; Wang, J.; King, C.W. Evaluation of the Onshore Wind Energy Potential in Mainland China—Based on GIS
Modeling and EROI Analysis. Resour. Conserv. Recycl. 2020, 152, 104484. [CrossRef]
69. Palmer, G. A Framework for Incorporating EROI into Electrical Storage. BioPhys. Econ. Resour. Qual. 2017, 2, 6. [CrossRef]
70. Zhang, J.; Zhang, J.; Cai, L.; Ma, L. Energy Performance of Wind Power in China: A Comparison among Inland, Coastal and
Offshore Wind Farms. J. Clean. Prod. 2017, 143, 836–842. [CrossRef]
71. Walmsley, T.G.; Walmsley, M.R.W.; Atkins, M.J. Energy Return on Energy and Carbon Investment of Wind Energy Farms: A Case
Study of New Zealand. J. Clean. Prod. 2017, 167, 885–895. [CrossRef]
Sustainability 2022, 14, 7098 20 of 20

72. Huang, Y.-F.; Gan, X.-J.; Chiueh, P.-T. Life Cycle Assessment and Net Energy Analysis of Offshore Wind Power Systems. Renew.
Energy 2017, 102, 98–106. [CrossRef]
73. Ecoinvent Centre for Life Cycle Inventories. EcoInvent Life Cycle Inventory Database; Ecoinvent: Zurich, Switzerland, 2021.

You might also like