Download as pdf or txt
Download as pdf or txt
You are on page 1of 64

5 July 2023

U.S. Environmental Protection Agency


EPA Docket Center, Office of Air and Radiation Docket, Mail Code 28221T
1200 Pennsylvania Avenue, NW
Washington, DC 20460

Docket ID No. EPA-HQ-OAR-2022- 0829

RE: International Council on Clean Transportation (ICCT) comments on EPA proposed rule,
titled, “Multi-Pollutant Emissions Standards for Model Years 2027 and Later Light-Duty and
Medium-Duty Vehicles”

Dear Michael Olechiw:

The International Council on Clean Transportation (ICCT) welcomes the opportunity to


provide comments on the EPA’s proposed Multi-Pollutant Emissions Standards for Model Years
2027 and Later Light-Duty and Medium-Duty Vehicles. The ICCT is an independent nonprofit
organization founded to provide unbiased research and technical analysis to governments in
major vehicle markets around the world. Our mission is to improve the environmental
performance and energy efficiency of road, marine, and air transportation in order to benefit
public health and mitigate climate change.

We commend EPA on its continuing efforts to reduce greenhouse gas (GHG) and
conventional pollutant emissions from light- and medium-duty vehicles. EPA’s standards have
paved the way for cleaner vehicles and delivered enormous climate and health benefits for the
United States. This proposed rulemaking is one more step in EPA’s progress in reducing climate
and air pollution from transportation.

These comments provide technical observations on EPA’s proposal that the agency may
consider in finalizing its rulemaking. We would be glad to clarify or elaborate on any points made
in these comments. EPA staff can feel free to contact our U.S. Passenger Vehicle Program
Lead, Pete Slowik (peter.slowik@theicct.org) with any questions.

Sincerely,

Rachel Muncrief, PhD


Acting Executive Director
International Council on Clean Transportation
TABLE OF CONTENTS
Summary of ICCT Comments .................................................................................................2
Market readiness......................................................................................................................5
Alignment with greenhouse gas emission reduction targets and climate goals ......... 11
Life-cycle comparison of combustion and electric vehicles ............................................. 14
International policy comparison ......................................................................................... 15
Greenhouse gas emissions standards ............................................................................. 15
Electric vehicle uptake ...................................................................................................... 17
Comparison with automaker commitments ...................................................................... 18
Battery electric vehicle cost ................................................................................................ 21
Combustion vehicle efficiency potential and cost-effectiveness ................................... 30
Gasoline Direct Injection (GDI) ......................................................................................... 31
Cylinder deactivation (DEAC) ........................................................................................... 31
Cooled Exhaust Gas Recirculation (CEGR) ..................................................................... 32
Atkinson cycle engine (ATK) ............................................................................................ 32
Miller cycle engine (MIL) ................................................................................................... 33
Mild hybrid (MHEV) ........................................................................................................... 34
Strong hybrid (HEV) .......................................................................................................... 37
Negative valve overlap in-cylinder fuel reforming (NVO) ................................................. 38
Passive prechamber combustion (PPC) .......................................................................... 39
High energy ignition (HEI)................................................................................................. 39
Transmissions ................................................................................................................... 39
Lightweighting ................................................................................................................... 40
Potential technology penetration impacts of improved ICE technology adoption ........... 40
Off-cycle credits .................................................................................................................... 42
Criteria pollutant emissions ................................................................................................ 43
NMOG and NOx emissions .............................................................................................. 43
International comparison on NMOG + NOx emission standards ..................................... 44
Evaporative emission ........................................................................................................ 46
Particulate matter emission .............................................................................................. 46
Benefits and cost-effectiveness of the proposal............................................................... 54
Proposed stringency ............................................................................................................ 54
Observations about proposed footprint curves ............................................................... 55
Plug-in hybrid electric vehicles........................................................................................... 57
Utility factor ....................................................................................................................... 57
Incorporating PHEVs in the proposal analysis ................................................................. 58
Medium-duty vehicles .......................................................................................................... 59
Stringency ......................................................................................................................... 59
Fuel neutrality.................................................................................................................... 61
MDV data and transparency ............................................................................................. 62
1
SUMMARY OF ICCT COMMENTS
ICCT strongly supports EPA’s proposed multi-pollutant emission standards for light- and
medium-duty vehicles for model years 2027-2032. The proposed standards are critical to
achieving the pace and scale of needed transportation emission reductions in the United States,
where there is a clear and urgent need to rapidly transition to cleaner vehicles. Continued and
strengthened standards are necessary to protect public health and deliver on national
environmental obligations. We support the proposed standards that would dramatically reduce
climate and air pollution from new passenger and medium-duty vehicles and deliver trillions of
dollars in net benefits.

ICCT recommends the adoption of Alternative 1, the most stringent option EPA presents in its
proposal for the light-duty vehicle (LDV) GHG standards. We believe even greater GHG
reductions than laid out in Alternative 1 would be feasible because today’s technology, policy,
and market landscape have primed the market for a rapid transition to cleaner vehicles.
Substantial public and private sector investments and a comprehensive package of federal and
state level policies make the timing and stringency of the proposed rule achievable, feasible,
and cost-effective. At the federal level, the combination of substantial consumer and industry
incentives from the $370 billion allocated to climate and clean energy investments through the
Inflation Reduction Act of 2022 (IRA) will accelerate the shift to electric vehicles while supporting
a domestic supply chain and charging infrastructure buildout. In parallel, the Bipartisan
Infrastructure Law complements the IRA by investing $7.5 billion in electric vehicle charging
infrastructure, $10 billion in clean transportation, and more than $7 billion in battery
components, critical minerals, and materials.

At the state level, policymakers are charging ahead with their own zero-emission vehicle
regulations, investments, consumer incentives, planning, and infrastructure deployment.
California’s Advanced Clean Cars II (ACC II) regulations will require dramatic reductions in light-
duty vehicle emissions to 100% zero-emissions by 2035 through the Zero-Emission Vehicle
Regulation and the Low-emission Vehicle Regulations. Many other U.S. states follow
California’s leadership on automotive emissions regulations. As of May 2023, 7 states have
adopted ACC II and it is likely that others will follow. By adopting its proposal, EPA can build on
ACC II and expand access to cleaner vehicles more broadly across the U.S.

Globally, automakers have already announced over $1.2 trillion in investments in electrification.
These investments will lead to greatly expanded model line-ups and production volumes,
technological advancements, and reduced costs. Battery mineral resources and production
capacity are sufficient to meet the standards. There are enough mineral resources available to
support a global transition to EVs and there are substantial ongoing investments in new projects
along the mineral supply chain in the U.S. and in friendly countries. Significant and growing
battery recycling capacity, along with battery technology diversification and improving EV
efficiency, will reduce pressure on EV mineral demand. Domestic battery production capacity is
quickly ramping up, already increasing by more than one-third since the IRA was passed. With
major announced public, private, and utility investment in chargers, the infrastructure is also
being built to meet the BEV expansion projected by EPA.

The United States is not alone in its commitment to transition to cleaner cars and trucks. EPA’s
proposal is in line with regulations in leading countries and automaker commitments. If

2
implemented, the proposed rule would put the U.S. on track to catch up with Europe and China,
which have led in the global transition to zero-emissions through 2022. Canada, the European
Union, the United Kingdom, and seven U.S. states have committed to entirely phase out the
sale or registration of new internal combustion engine vehicles by 2035. EPA’s proposal also
aligns with announcements from auto manufacturers; Ford, General Motors, Mercedes-Benz,
Audi, and others have committed to selling 100% zero-emission vehicles globally or in leading
markets by 2035. With the rapid development of ZEV technologies, automakers are constantly
updating their targets. In fact, ICCT reviewed EPA’s compilation of automaker electric vehicle
sales targets and announcements in the proposal and identified several additional
announcements that had been made through 2022 beyond those in EPA’s assessment,
including announcements of expanded and accelerated electric vehicle targets by General
Motors, Ford, Stellantis, Mercedes-Benz, Toyota, and Hyundai-Kia. These developments further
reflect industry commitment to and desirability of ZEV technologies.

ICCT recommends EPA finalize Alternative 1 for the greenhouse gas (GHG) emissions
standards for light-duty vehicles (LDV) and believes the GHG standards for medium-duty
vehicles (MDV) could also be strengthened. These recommendations are based on thorough
and clear evidence (detailed in the sections below) that additional GHG reductions could be
possible through continued and cost-effective internal combustion engine (ICE) technology
improvements and battery electric vehicle (BEV) penetration at lower costs than EPA estimates
in its modeling.

There is potential for significant additional GHG savings from ICE vehicles beyond what EPA
has modeled for its proposal. Many existing and recently announced ICEV technology
improvements have ample room for increased application throughout the ICEV fleet. ICCT has
identified several technologies, including advanced mild hybrids and plug-in hybrid EVs
(PHEVs) that are commercially available and could significantly and cost-effectively improve
ICE vehicle efficiency that EPA has not included in its modeling. We have also identified other
technologies for which EPA has overestimated the costs in its modeling, such as Miller cycle
engines and mild hybrid technology. If EPA made these changes to its model, we believe the
agency would project lower overall compliance costs and lower BEV shares to achieve the
same level of GHG reductions compared to its proposal.

ICCT also believes that EPA has overestimated battery electric vehicle (BEV) costs. In in-depth
study, we find that incremental BEV costs relative to their combustion vehicle counterparts are
much lower than the estimates by EPA by about $2,300 (pickups) to about $5,300 (crossovers)
in 2027. For 2032, EPA finds that BEVs have incremental costs of about $3,000 to $5,000;
during this same timeframe, the ICCT study finds that BEVs will be cheaper to purchase than
comparable gasoline vehicles. This difference is the result of several factors. We find that on
average, battery pack costs in the ICCT analysis are about $3,700 less than in EPA’s analysis
in 2027 and about $5,200 less than in EPA’s analysis in 2032. We believe EPA has
overestimated pack size, BEV energy consumption, and pack-level costs ($/kWh). If EPA were
to update their analysis with the data and evidence regarding BEV cost and technical
specifications based on the ICCT cost study, BEV costs would be lower, the costs of
compliance would be lower, and the net benefits would be greater.

ICCT believes there is even greater opportunity to capture technology potential and cost
reductions for medium-duty vehicles (MDVs) in EPA’s modeling. While ICCT believes EPA
underestimated the potential for ICE vehicle improvements for LDVs in its modeling, EPA did
not model any ICE vehicle improvement technologies for MDVs. All of the above findings and
recommendations about combustion and EV technology potential and cost for LDVs apply to

3
MDVs. In addition, EPA has not included any kind of hybrid technology options in its modeling
for MDVs and has omitted other ICE technology options it included for LDVs. All of the LDV
technology options we review in these comments can be applied to MDVs and can deliver cost
savings in meeting the proposed GHG standards. We strongly recommend EPA add ICE and
PHEV technology options and incorporate the latest evidence on declining ICE and BEV
technology costs into its modeling for MDVs. We also recommend EPA consider increasing the
stringency of the MDV GHG standards since this could be achieved at lower costs when
considering the evidence presented here on technology options.

For particulate matter emissions, ICCT strongly support’s EPA’s proposed PM limit. Fine
particulate matter (PM2.5) emissions from vehicles are a major environmental health hazard, and
setting a more stringent PM emissions limit is critical for protecting public health. We support
EPA’s proposed multipollutant rule that sets the PM limit at 0.5 mg/mile on all test cycles with
the addition of the cold temperature FTP test cycle, which would help to greatly reduce tailpipe
PM emissions from new gasoline vehicles. Based on our review of recent evidence of
increasing PM emissions from recent model year gasoline light-duty vehicles and trucks, we find
an urgent need to implement a stronger PM standard. Due to the urgent need to address the
rise in PM emissions with recent gasoline vehicles, we recommend EPA adopt the accelerated
phase-in pathway for the PM standard.

ICCT supports EPA’s proposed improvements on a number of technical design elements that
will strengthen the overall rule. ICCT strongly supports EPA’s proposal to phase out additional
credits given to automakers for technologies that are not captured in the test cycle (off-cycle
credits; OC) and that previously were considered novel but now are fully commercialized. In
particular, ICCT supports EPA’s proposal to sunset or eliminate off-cycle (OC) credits for air-
conditioning (AC), to phaseout off-cycle (OC) credits by MY2031, to eliminate the off-menu OC
credit option starting MY2027, and to limit OC credits to ICE vehicles. On footprint curves, ICCT
commends EPA for its proposed standards becoming increasingly flat, and we see clear
evidence that the car and truck footprint curves could be flattened even further, as discussed
below. On PHEVs, ICCT applauds EPA for and its proposal to adjust the PHEV utility factor
(UF) curve to better fit real-world PHEV usage data.

ICCT strongly supports the proposed multi-pollutant emission standards for light- and medium-
duty vehicles for model years 2027-2032 and recommends its finalization as quickly as possible.
Doing so will provide a clear long-term signal that automakers, suppliers, charging companies,
and utilities need to make needed investments with confidence.

Our detailed comments below provide further background, analysis and references on these
points.

4
MARKET READINESS
The global transition to zero-emission vehicles is accelerating. The U.S. is the third largest
electric vehicle market, behind China and Europe. In 2022, new U.S. plug-in electric vehicle
sales reached nearly 1 million units, representing over 7% of new light-duty vehicle sales.1 This
growth is expected to continue, as evidenced by the more than $1.2 trillion in announced
automaker spending on electric vehicles, widespread consumer demand, and billions of dollars
in new federal tax credits, incentives, and investments.2 Many electric vehicles have multi-month
long wait times, indicating that demand is outpacing supply, and consumer research surveys
show that demand for BEVs in the United States continues to increase and that 30% of licensed
drivers aren’t even considering gasoline vehicles for their next purchase or lease.3 Substantial
public and private sector investments and state and federal policies have primed the market for
a rapid shift to electric vehicles.

The Inflation Reduction Act (IRA) of 2022 will accelerate electric vehicle sales in the United
States across all vehicle types. The $370 billion allocated to climate and clean energy
investments dramatically expands tax credits and incentives to deploy more clean vehicles,
including commercial vehicles, while supporting a domestic EV supply chain and charging
infrastructure buildout. IRA transportation sector provisions will accelerate the shift to zero-
emission vehicles by combining consumer and manufacturing policies. Consumer tax credits for
new and used EVs and tax credits for commercial EVs, along with individual and commercial
charging infrastructure tax credits, will increase sales. Domestic supply chain incentives and
investments will boost EV manufacturing and battery production. Critical mineral mining and
refining incentives will bolster industrial development.

An ICCT and Energy Innovation study assesses the future impact of the IRA on electrification
rates for LDV sales in the United States through 2035.4 We analyze the value of the personal
and commercial EV tax credits, factoring in the various supply chain, income, and price caps on
new EVs, and combine this with new estimates of future light-duty EV cost declines. We find
that, on average over the period 2023–2032, the IRA tax credits will reduce EV purchase costs
by $3,400 to $9,050 and accelerate the timing for price parity with combustion vehicles. Using
methodologies from the Energy Policy Simulator, we project how these changing costs and
incentives over time will affect the LDV markets in the United States.

Figure 1 summarizes the results from the ICCT and Energy Innovation IRA study. It shows the
findings of estimated new electric vehicle sales shares for different IRA scenarios depending on
how certain provisions are implemented and how the value of incentives is passed on to
consumers. The figure shows our modeled projection of how the IRA will accelerate
electrification. By providing thousands of dollars in financial incentives, the IRA unlocks

1 Based on data from EV-Volumes. (2023). https://www.ev-volumes.com/datacenter/


2 Lienert, P. (2022, October 25). Exclusive: Automakers to double spending on EVs, batteries to $1.2 trillion by 2030.
Reuters. https://www.reuters.com/technology/exclusive-automakers-double-spending-evs-batteries-12-trillion-by-
2030-2022-10-21/
3 Kiai, E. (2023, May 6). Waiting times for new electric car deliveries down by 42% since October peak. Electrifying.

https://www.electrifying.com/blog/article/waiting-times-for-new-electric-car-deliveries-down-by-42-since-october-
peak; Consumer Reports (2023). Automakers must increase production of electric vehicles or risk losing
customers, new analysis finds. https://advocacy.consumerreports.org/press_release/new-analysis-of-consumer-ev-
demand/
4 Slowik, P., Searle, S., Basma, H., Miller, J. Zhou, Y., Rodriguez, F., Buysse, C., Kelly, S., Minjares, R., Pierce, L.

(ICCT) Orvis, R., and Baldwin, S. (Energy Innovation). (2023). Analyzing the impact of the inflation reduction act
on electric vehicle uptake in the United States. https://theicct.org/publication/ira-impact-evs-us-jan23/
5
widespread consumer benefits. We find rapid projected EV uptake when considering both
expected manufacturing cost reductions and the IRA incentives, as well as state policies. By
2030, we find a range of a 48%–61% projected EV sales share, increasing to 56%–67% by
2032, the final year of the IRA tax credits.

Figure 1. Baseline, Low, Moderate, and High projections of EV sales share for light-duty
vehicles, considering ACC II adoption in only California versus increased states 5

In parallel, states are also adopting their own zero-emission vehicle regulations, investments,
consumer incentives, planning, and infrastructure deployment. California’s Advanced Clean
Cars II (ACC II) regulations will require dramatic reductions in light-duty vehicle emissions to
100% zero-emissions by 2035 through the Zero-Emission Vehicle Regulation and the Low-
emission Vehicle Regulations.6 Many other U.S. states follow California’s leadership on
automotive emissions regulations. As of May 2023, California, Massachusetts, New York,
Oregon, Vermont, Virginia, and Washington have adopted ACC II. It is likely that many other
states will continue to follow California’s leadership and adopt the new ACC II Program to
benefit from the anticipated emissions reduction and health benefits of the program.7 Many
additional states currently follow the Advanced Clean Cars regulations through model year
2026; as of May 13th, 2022, 17 U.S. states have adopted all or part of California’s low-emission

5 ICCT and Energy Innovation. (2023). Analyzing the impact of the inflation reduction act on electric vehicle uptake in
the United States. https://theicct.org/publication/ira-impact-evs-us-jan23/
6 California Air Resources Board. (2023). Advanced Clean Cars II Regulations: All new passenger vehicles sold in

California to be zero emissions by 2035. https://ww2.arb.ca.gov/our-work/programs/advanced-clean-cars-


program/advanced-clean-cars-ii
7 Houk, J., Huang, J., and Sussman. (2023). Benefits of adopting California’s Advanced Clean Cars II standards in

sixteen U.S. states. Sonoma Technology Inc. Retrieved from https://theicct.org/publication/benefits-of-state-level-


adoption-of-california-acc-ii-regulations/
6
and zero-emission vehicle regulations, and about 37% of national new light-duty vehicle sales
meet California’s emission standards.8 Acknowledging the significant fraction of new U.S. light-
duty vehicles subject to California’s regulations, ICCT recommends EPA align with California’s
BEV and PHEV provisions on battery durability, warranty, state of health, and other measures.
Such an alignment not only ensures adequate performance of plug-in vehicles for purposes of
emissions reductions, but it would protect all consumers purchasing EVs and would support the
development of secondary EV markets.

The U.S. share of global automaker electric vehicle investments is increasing, largely driven by
these state and federal policies and investments. Research from January 2023 estimates that
$210 billion in automaker electric vehicle manufacturing and $54 billion in battery production
investments had been announced for the U.S.9 These electric vehicle and battery manufacturing
investments will lead to greatly expanded model line-ups and production volumes, technological
advancements, and reduced costs.

EPA’s proposed rule, along with sustained commitments will help build out the charging
infrastructure needed to support accelerated electrification.10 Norway – the world’s electric
vehicle sales share leader – achieved nearly 80% EV sales in 2022 with one public charger for
every 26 EVs.11 Significant resources are already being dedicated to charging in the United
States, including the $7.5 billion allocation from the IIJA as well as several billions of dollars
in power utility and private sector investment.12 For example, British Petroleum announced
plans to invest $1 billion in EV charging in the U.S. by 2030.13 Automakers are investing too
– GM, working with its dealers, aims to install up to 40,000 public charging stations across the
U.S. and Canada.14 Globally, Bloomberg New Energy Finance expects $100 billion to be spent
to grow charging infrastructure in the next 3 years alone. 15 In parallel, state infrastructure
planning is underway, with all 50 states, DC, and Puerto Rico submitting and receiving approval

8 California Air Resources Board. (May 13, 2022). States that have adopted California’s vehicle standards under
Section 177 of the Federal Clean Air Act. https://ww2.arb.ca.gov/sites/default/files/2022-
05/%C2%A7177_states_05132022_NADA_sales_r2_ac.pdf
9 Gabriel, N. (2023, January 12). $210 billion of announced investments in electric vehicle manufacturing headed for

the U.S. Atlas EV Hub. https://www.atlasevhub.com/data_story/210-billion-of-announced-investments-in-electric-


vehicle-manufacturing-headed-for-the-u-s/
10 Searle, S., Kodjak, D., and Slowik, P. (2023, April 26). Infrastructure and supply chains won’t hold up EPA’s

proposed light and medium-duty vehicle standards. International Council on Clean Transportation.
https://theicct.org/infrastructure-and-supepas-proposed-ldv-mdv-standards-apr23/
11 Johnson, P. (2023, January 2). This is the Norway – nation hits record EV share in 2022 on its way to ending gas

car sales. Electrek. https://electrek.co/2023/01/02/norway-hits-record-ev-share-in-2022/; Kok, I., and Hall, D.


(2023). Battery electric and plug-in hybrid electric vehicle uptake in European cities. International Council on Clean
Transportation. https://theicct.org/publication/bev-phev-european-cities-mar23/
12 The White House. (2023a). Fact sheet: Biden-Harris administration announces new standards and major progress

for Made-in-America National network of Electric vehicle chargers. https://www.whitehouse.gov/briefing-


room/statements-releases/2023/02/15/fact-sheet-biden-harris-administration-announces-new-standards-and-
major-progress-for-a-made-in-america-national-network-of-electric-vehicle-chargers/; Electric utility filings. Atlas
EV Hub. https://www.atlasevhub.com/materials/electric-utility-filings/
13 British Petroleum (2023, February 15). BP plans to invest $1 billion in EV charging across US by 2030, helping to

meet demand from Hertz’s expanding EV rentals. https://www.bp.com/en_us/united-states/home/news/press-


releases/bp-plans-to-invest-1-billion-in-ev-charging-across-us-by-2030-helping-to-meet-demand-from-hertzs-
expanding-ev-rentals.html
14 General Motors. (2022, December 7). GM advances dealer community charging program.

https://news.gm.com/newsroom.detail.html/Pages/news/us/en/2022/dec/1207-charging.html
15 Bloomberg NEF. (2023, January 20). Next $100 Billion EV-Charger spend to be super fast.

https://about.bnef.com/blog/next-100-billion-ev-charger-spend-to-be-super-fast/
7
for their National Electric Vehicle Charging Network plans in 2022.16 New research by WSP and
EDF find that the expected growth in public charging from announced deployments will provide
at least 70% of the infrastructure needed by 2030 to meet the BEV projections in EPA’s
proposal; for DC fast chargers, the announced investments sum up to over 100% of the amount
needed by 2030.17 These announcements and investments indicate that the pace and scale of
U.S. charging infrastructure deployment is on track to support the projected BEV market shares
in EPA’s proposal.

In addition to all this industry investment in EV manufacturing, battery production, and the
battery material supply chain, substantial investments are underway in the electric power sector.
From 2012 through mid 2022, about $3.6 billion in utility transportation electrification investment
plans have been approved across the country.18 These investments include upgrades in grid
capacity, safety, resilience, and managed charging. The U.S. already has enough power
generation and transmission capacity to fuel the EV expansion over the next few years.19
Meeting the 2050 demand requires about 1% per year growth in electricity production, well
below the 3.2% average annual growth rate for the electricity generation over the past 70
years.20

There is evidence that it should be possible to quickly scale up investments into mining and
battery production with careful planning and investment. EPA’s proposal lays out the much
needed roadmap for the mineral mining, refining, and battery production supply chain to plan
and invest around. There are more than enough minerals available for a global transition to
EVs, and the amount of raw material reserves domestically and in friendly countries greatly
exceeds what is needed to meet EPA’s proposal.21 With the IRA’s Advanced Manufacturing
Production Tax Credit and the domestic content provisions in the Clean Vehicle Tax Credit, the
U.S. directly incentivizes mining, recycling, and battery production on U.S. soil, and further
supports establishing resilient material supply chains from friendly countries. The manufacturing
subsidy of $45/kWh cuts about one third of total battery costs (global average of $151 in 2022),
making battery production in the U.S. even cheaper than in China. 22 This support showed an
immediate effect. In response to the IRA, we saw a significant uptick by more than one-third in

16 U.S. Department of Transportation. (2022, September 27). President Biden’s Bipartisan Infrastructure Law provides
$5 billion to help States install EV chargers along interstate highways. https://highways.dot.gov/newsroom/historic-
step-all-fifty-states-plus-dc-and-puerto-rico-greenlit-move-ev-charging-networks
17 U.S. Public Electric Vehicle (EV) Charging Infrastructure Deployment, WSP for EDF (July 2023). [Link forthcoming]
18 Lepre, N. (2022, September). Electric utility filing annual update. Atlas Public Policy. https://atlaspolicy.com/wp-

content/uploads/2022/09/Electric-Utility-Filing-Brief-July-2021-through-June-2022-v2-1.pdf
19 Houston, S. (2022, September 12). Can the electric grid handle EV charging? Union of Concerned Scientist.

https://blog.ucsusa.org/samantha-houston/can-the-electric-grid-handle-ev-charging/
20
Harto, C. (2023, May 10). Blog: Can the grid handle EVs? Yes! Consumer Reports.
https://advocacy.consumerreports.org/research/blog-can-the-grid-handle-evs-yes/; Miller, T., and Bischof, A.
(2020, November 20). Electricity Demand’s COVID comeback. Morningstar investor.
https://www.morningstar.com/stocks/electricity-demands-covid-comeback
21 Slowik, P., Lutsey, N., and Hsu, C-W. (2020). How technology, recycling, and policy can mitigate supply risks to the

long-term transition to zero-emission vehicles. International Council on Clean Transportation.


https://theicct.org/publication/how-technology-recycling-and-policy-can-mitigate-supply-risks-to-the-long-term-
transition-to-zero-emission-vehicles/ and Oge, M. (2023, May 16). History shows EPA’s proposed vehicle
emissions rule can be done – it’s worth trillions! Forbes.
https://www.forbes.com/sites/margooge/2023/05/16/history-shows-epas-proposed-vehicle-emissions-rule-can-be-
done---its-worth-trillions/?sh=ea3150cc6275
22 Bloomberg New Energy Finance. (2022, December 6). Lithium-ion battery pack prices rise for the first time to an

average of $151/kWh. https://about.bnef.com/blog/lithium-ion-battery-pack-prices-rise-for-first-time-to-an-average-


of-151-kwh/
8
announced plans for battery production facilities, catching up with Europe.23 Longer-term
forecasts now indicate U.S. battery production capacity at 1 TWh by 2030, approaching
forecasted battery demand of 1.2 TWh taking the proposed EPA standards into account.24 A
mapping compilation of U.S. EV supply chain investment from December 2022 shows that there
are several dozen investments in minerals, battery production, recycling, and other electric
vehicle facilities across the country.25 A separate study by the U.S. Department of Energy
compiled U.S. battery supply chain announcements and found over $100 billion announced as
of May 2023 with over 160 new or expanded minerals, materials processing, and manufacturing
facilities, and over 70,000 new jobs.26 Specific recent examples from June 2023 include a $9.2
billion federal loan to Ford to greatly expand U.S. electric vehicle and battery production, and
Hyundai’s announcement to increase its EV investments to $28 billion while reducing China
operations.27

Battery re-use and recycling will be important to limit unnecessary mining and domestically
maintain critical materials needed for new batteries. 28 Several automakers – BMW, Ford, Geely,
General Motors, Honda, Hyundai-Kia, Stellantis, Tesla, and Toyota – have announced one or
more battery recycling or repurposing project in the United States.29 Figure 2 shows draft ICCT
estimates of how the installed U.S. recycling capacity as of 2022 and the announced 2030
capacity compares with the projections for end-of-life batteries that could become available for
recycling in the U.S. It shows that the installed capacity in 2022 (76,000 tons) is sufficient to
process end-of-life batteries up to the year 2035. When also considering the recycling plants
announced as of 2022, bringing the total recycling capacity to 325,000 tons, sufficient capacity
is available to recycle end-of-life batteries until 2039. After which, much more recycling capacity
is needed as the mass of end-of-life batteries becoming available for recycling increases swiftly
between 2030-2050. Not shown in this figure are the additional tons of material from battery
production scrap that become available for recycling.

23 Benchmark Minerals. (2022, November 23). IRA lifts US battery pipeline growth in second half of the year,
outpacing Europe. https://source.benchmarkminerals.com/article/ira-lifts-us-battery-pipeline-growth-in-second-half-
of-the-year-outpacing-europe
24 Benchmark Minerals. (2023, April 14). US EPA emissions rules ratchet up pressure on the battery supply chain.

https://source.benchmarkminerals.com/article/us-epa-emissions-rules-ratchet-up-pressure-on-the-battery-supply-
chain
25 EV supply chain and investment map. (2022, December 21). Zero Emission Transportation Association.

https://www.zeta2030.org/education-fund/investments
26 U.S. Department of Energy. (2023, May 23). Investments in American-Made Energy: Battery supply chain

investments. https://www.energy.gov/investments-american-made-energy
27 Rathi, A., Natter, A., and Naughton, K. (2023, June 22). Ford gets $9.2 billion to help US catch up with China’s EV

dominance. Bloomberg. https://www.bloomberg.com/graphics/2023-ford-ev-battery-plant-funding-biden-green-


technology/ and Yim, H., and Yang, H. (2023, June 20). Hyundai raises EV investment to $28 billion, to reduce
China operations. Reuters. https://www.reuters.com/business/autos-transportation/hyundai-motor-invest-8541-
billion-by-2032-accelerate-ev-plans-2023-06-20/
28 Tankou, A., Bieker, G., and Hall, D. (2023). Scaling up reuse and recycling of electric vehicle batteries: assessing

challenges and policy approaches. International Council on Clean Transportation.


https://theicct.org/publication/recycling-electric-vehicle-batteries-feb-23/
29 Shen, C., Fadhil, I., Yang, Z., Searle, S. (2023). The global automaker rating 2022: who is leading the transition to

electric vehicles? International Council on Clean Transportation. https://theicct.org/publication/the-global-


automaker-rating-2022-may23/
9
Figure 2. Estimates of installed battery recycling capacity (data circles) and mass of end-of-life
electric vehicle batteries (bars)

The shift to cleaner cars and trucks can strengthen domestic manufacturing and supply chains,
increase industrial competitiveness, and create good-paying jobs.30 Research about the U.S. job
potential from electric vehicle manufacturing finds that transitioning 50% of sales to battery
electric vehicles (BEVs) can lead to 150,000 net job creation in the auto sector, including auto
parts and auto assembly, if coupled with an increase in domestic content for vehicles sold in the
U.S., and market share growth for U.S. made vehicles.31 The growth in electric vehicle charging
infrastructure can also lead to domestic job creation. Research quantifying the number of jobs
associated with EV infrastructure from assembly, deployment, and maintenance appears
limited. Draft work by the ICCT is underway that quantifies the number of jobs needed to
support this rapidly growing infrastructure network.

Table 1 shows the draft results of the number of infrastructure-related jobs created from the
charging infrastructure sector in 2035: more than 183,000. The light-duty vehicle EV
infrastructure industry would generate in total close to 160,000 jobs that involve the public,
workplace, and home charging infrastructure assembly, deployment, and maintenance, and the
MDHV EV infrastructure industry will generate close to 23,700 jobs. Around 36% of these jobs
come from electrical work, followed by 29% from assembly, 17% from maintenance, 10% from
planning and design, 6% from general labor, and 3% from administration and legal. These jobs
would further support supply chain jobs in assembly components and materials for EV charging
equipment, which are not included in our estimates but can be significant, especially with
increase in domestic production.

30 The White House (2023b). Fact sheet: Biden-Harris Administration announces new private and public sector
investments for affordable electric vehicles. https://www.whitehouse.gov/briefing-room/statements-
releases/2023/04/17/fact-sheet-biden-harris-administration-announces-new-private-and-public-sector-investments-
for-affordable-electric-vehicles/; Naimoli, S., Kodjak, D., German, J., and Schultz, J. (2017, May 23). International
competitiveness and the auto industry: what’s the role of motor vehicle emission standards? International Council
on Clean Transportation. https://theicct.org/publication/international-competitiveness-and-the-auto-industry-whats-
the-role-of-motor-vehicle-emission-standards/
31 Barrett, J., and Bivens, J. (2021, September 22). The stakes for workers in how policymakers manage the coming

shift to all-electric vehicles. Economic Policy Institute. https://www.epi.org/publication/ev-policy-workers/


10
Table 1. Estimate of light-duty and medium- and heavy duty vehicle charging infrastructure-
related jobs in 2035
MHDV
LDV chargers
chargers
Multi- Single- Total
Job type Public Workplace family family Public
home home
DC fast Level 2 Level 2 Level 2 Level 2 DC fast
Assembly 290 440 1,400 2,360 43,700 4,230 52,420
Planning and design 260 1,420 4,510 7,610 3,890 17,690
General labor 200 890 2,810 4,740 2,910 11,550
Electrical work 370 1,820 5,770 9,740 41,900 5,500 65,090
Administration and Legal 100 410 1,320 2,220 1,510 5,550
Maintenance 630 4,010 8,710 11,890 5,670 30,910
Total 1,840 8,990 24,520 38,560 85,600 23,710 183,220

The number of chargers needed for light-duty vehicles and the associated findings for U.S. job
growth potential are based on an analysis of charging needs based on achieving Biden
administration’s goal of 50% EV sales by 2030, which is assumed to increase to 80% sales by
2035. For MHDV, the model a scenario where EV sales achieve 30% for tractor truck, 50% for
buses and rigid truck by 2030, and 100% for all segments in 2040, which builds on the
manufacturers’ ambitions and existing legislation in California and several other states.32 The
analysis is being updated based on the projections in EPA’s new multi-pollutant standards
proposal. Because the EPA proposal estimates that 60% of new light-duty vehicle sales would
be battery electric by 2030 and 67% by 2032 – greater than Biden’s 50% by 2030 target –
updating our analysis for increased BEV penetration would increase the number of chargers
needed and thus increase the estimates of the number of jobs in this industry through 2035
compared to what is shown in Table 1.

ALIGNMENT WITH GREENHOUSE GAS EMISSION


REDUCTION TARGETS AND CLIMATE GOALS
ICCT strongly supports the proposed Multi-Pollutant Emissions Standards for Model Years 2027
and Later Light-Duty and Medium-Duty Vehicles and recommends its finalization. This
rulemaking is critical to achieving the pace and scale of needed transportation emission
reductions in the United States. There is a clear and urgent need to rapidly transition the
transportation sector to zero-emission vehicles. Continued and strengthened standards are
important to protect public health and deliver on national air quality and climate change goals.

On his first day in office, President Biden issued an executive order for the United States to
rejoin the Paris Climate Agreement.33 With this executive order, the U.S. joined nearly every
country on earth with the shared goal of limiting global warming to well below 2 degrees
Celsius.34 Delivering on the goals of this Agreement will require a significant reduction in
greenhouse gas emissions. The transportation sector is the largest contributor to climate

32 International Council on Clean Transportation. (2022a). ICCT comments on EPA proposed HDV rule.
https://theicct.org/comments-epa-proposed-hdv-rule-may22/
33 U.S. Department of State. (2021, February 19). The United States officially rejoins the Paris Agreement.

https://www.state.gov/the-united-states-officially-rejoins-the-paris-agreement/
34 Denchak, M. (2021, February 19). Paris Climate Agreement: Everything you need to know. Natural Resources

Defense Council. https://www.state.gov/the-united-states-officially-rejoins-the-paris-agreement/


11
pollution in the United States, with light-duty vehicles—passenger cars, SUVs, and pickup
trucks—accounting for more than half of these emissions. Therefore, a shift to zero-emission
vehicles like battery electric vehicles, along with increased efficiency of new gasoline vehicles,
is needed.

The ICCT’s global modeling assessment for the Zero Emission Vehicle Transition
Council shows that for the U.S. to be on a path to limiting global warming to well below 2° C, at
least 67% of new passenger vehicle sales will need to be plug-in electric by 2030, with the vast
majority (65%) being battery electric vehicles (BEVs).35 In parallel, new gasoline vehicles would
need to improve at about 3.5% per year for the U.S. to be on a path to be Paris-compatible.
Previous ICCT research estimated that the 2030 U.S. EPA standard would need to be about 57
grams CO2/mile by 2030 to achieve this combination of new gasoline vehicle improvement with
greatly increased BEV penetration that aligns with the Paris Agreement. 36

The figure below summarizes the path from the current standard of 161 gCO2 per mile in 2026
to 57 gCO2e per mile in 2030 for the whole light duty fleet (top panel). This represents a
combination of continued gasoline vehicle efficiency improvements (middle panel) and
accelerated EV uptake to reach 67% sales in 2030 (bottom panel), including 65% BEVs and
about 2% PHEVs. This assumed composition of battery electric vs. plug-in hybrid electric
vehicles is based on BEVs’ substantial economic and environmental benefits relative to PHEVs.
Of course, if plug-in hybrids make up a greater share of EV sales than assessed here, then the
fleetwide standards would have to be more stringent than 57 gCO 2e per mile to be Paris-
compatible.37

35 Sen, A., and Miller, J. (2022). Emissions reduction benefits of a faster, global transition to zero-emission vehicles.
International Council on Clean Transportation. https://theicct.org/publication/zevs-global-transition-benefits-mar22/;
ZEV Transition Council. https://zevtc.org/
36 Slowik, P., and Miller, J. (2022, December 19). Aligning the U.S. greenhouse gas standard for cars and light trucks

with the Paris climate agreement. International Council on Clean Transportation. https://theicct.org/us-ghg-
standard-paris-agreement-dec22/
37 Slowik. P., Isenstadt, A., Pierce, L, Searle, S. (2022). Assessment of light-duty electric vehicle costs and consumer

benefits in the United States in the 2022-2035 time frame. International Council on Clean Transportation.
https://theicct.org/publication/ev-cost-benefits-2035-oct22/; Bieker, G. (2021). A global comparison of the life-cycle
greenhouse gas emissions of combustion engine and electric passenger cars. International Council on Clean
Transportation. https://theicct.org/publication/a-global-comparison-of-the-life-cycle-greenhouse-gas-emissions-of-
combustion-engine-and-electric-passenger-cars/
12
Figure 3. Stringency of model year 2030 GHG targets to be Paris-compatible

This Paris-compatible path is ambitious yet achievable. As the middle panel shows, the 3.5%
annual improvement for gasoline vehicles beyond 2026 is less than the rate of improvement
needed to meet the current standards through 2026. As discussed above, EV uptake consistent
with this scenario is possible through a combination of the new federal electric vehicle
incentives and investments from the Inflation Reduction Act and state policies.

Although reaching 57 gCO2e per mile by 2030 requires a steeper decline in GHG emissions
compared to the current EPA standards through 2026, ICCT’s analysis finds this rate of
improvement to be cost-effective for combustion and battery electric vehicles alike.38 In addition,

38 Lutsey, N., Meszler, D., Isenstadt, A., German, J., and Miller, J. (2017). Efficiency technology and cost assessment
for U.S. 2025-2030 light-duty vehicles. International Council on Clean Transportation.
https://theicct.org/publication/efficiency-technology-and-cost-assessment-for-u-s-2025-2030-light-duty-vehicles/;

13
the associated consumer benefits grow well into the tens of billions of dollars annually by 2027
due to greatly reduced upfront electric vehicle prices and lower fuel and maintenance costs
compared to fossil fuel vehicles.39

Life-cycle comparison of combustion and electric vehicles


The climate benefits of electric vehicles compared to internal combustion engine vehicles are
clear and growing. Numerous studies show that BEVs are much cleaner than combustion
vehicles over their lifetime. In fact, the most recent ICCT life-cycle analysis shows that only
battery electric and hydrogen fuel cell electric vehicles have the potential to achieve the
magnitude of life-cycle greenhouse gas emissions reductions needed to meet Paris Agreement
goals.40 Specifically, the assessment finds that the life-cycle emissions over the lifetime of BEVs
registered in 2021 in the United States are lower than comparable gasoline vehicles by 60%-
68%. For new vehicles registered in 2030, the life-cycle emissions gap between BEV and
gasoline vehicles increase to 62%-76%.

Figure 4. Life-cycle GHG emission of average medium-size gasoline internal combustion


engine and battery electric vehicles registered in the United States in 2021 and projected to be
registered in 2030.41

There are a few reasons our study finds such high life-cycle GHG benefits of BEVs compared to
gasoline passenger vehicles. The foremost reason is that electric motors are far more efficient

Slowik. P., Isenstadt, A., Pierce, L, Searle, S. (2022). Assessment of light-duty electric vehicle costs and consumer
benefits in the United States in the 2022-2035 time frame. International Council on Clean Transportation.
https://theicct.org/publication/ev-cost-benefits-2035-oct22/
39 Ibid
40 Bieker, G. (2021). A global comparison of the life-cycle greenhouse gas emissions of combustion engine and

electric passenger cars. International Council on Clean Transportation. https://theicct.org/publication/a-global-


comparison-of-the-life-cycle-greenhouse-gas-emissions-of-combustion-engine-and-electric-passenger-cars/
41 Bieker, G. (2021). A global comparison of the life-cycle greenhouse gas emissions of combustion engine and

electric passenger cars. International Council on Clean Transportation. https://theicct.org/publication/a-global-


comparison-of-the-life-cycle-greenhouse-gas-emissions-of-combustion-engine-and-electric-passenger-cars/
14
than internal combustion engines. Another important reason is that we expect continued
reductions in the U.S. electricity generation mix as renewables such as solar and wind become
more economically competitive (these sources also continue to be supported by fiscal incentives
through the IRA). Our analysis accounts for this expected reduction in the GHG emissions from
U.S. electricity throughout the 18-year lifetime of a passenger vehicle registered in the U.S.
before it is deregistered. A third reason why the life-cycle GHG emissions from BEVs are so low
is that estimated GHG emissions from battery production have declined significantly over the
past few years as we have collected new data from commercial-scale battery production plants.

These findings are aligned with other life-cycle analysis of electric and combustion vehicles in
the U.S. A 2022 UCS study found that electric vehicles produce fewer emissions than new
gasoline vehicles everywhere in the U.S., and that on average EVs produce emissions
equivalent to 91-mile-per-gallon gasoline cars.42 The same study found that BEVs’ higher
manufacturing emissions are quickly paid off after about 12,600 to 21,300 miles of driving,
which is typically reached in about 1-2 years for most U.S. drivers.

For combustion vehicles, there is no realistic pathway for deep decarbonization, even when the
potential for biofuels and biogas are taken into account. For gasoline cars, we account for the
current ~10% blend of corn ethanol and its associated GHG emissions. Blending corn ethanol
does not significantly reduce the life-cycle GHG emissions from combusting gasoline because
the GHG emissions from producing corn ethanol are also quite high when including EPA’s
estimated GHG emissions from land use change – only around 20% lower than the lifecycle
GHG emissions from producing and combusting gasoline. The production and consumption
within the U.S. of ultra-low GHG biofuels that can be blended into the gasoline mix (mainly
cellulosic ethanol) remains relatively very low.43 Given the greater incentives for second
generation biofuels used in aviation compared to the road sector in the IRA, it seems unlikely
that biofuels will significantly reduce the overall GHG intensity of the U.S. gasoline mix over the
timeframe of EPA’s proposal.

INTERNATIONAL POLICY COMPARISON


The United States is not alone in its commitment to transition to cleaner cars. The number of
national and subnational governments around the world that are committed to cleaner vehicles
continues to rise. The sections below summarize and compare the U.S. EPA proposal with the
standards in other leading markets around the world in terms of greenhouse gas emissions
requirements, projections for electric vehicle uptake, and criteria pollutant emissions standards.

Greenhouse gas emissions standards


Figure 5 illustrates the U.S. trajectories of average CO2 emissions performance and CO2
emissions targets for new passenger cars and light trucks over the years, in comparison with
other global vehicle markets that have CO2 emissions targets at least as of 2022. For a
consistent basis of comparison, CO2 emissions values for all markets have been converted to
the same U.S. test cycles (referred to as the CAFE cycle), where needed, using the ICCT cycle

42 Reichmuth, D., Dunn, J., Anair, D. (2022). Driving cleaner: Electric cars and pickups beat gasoline on lifetime
global warming emissions. Union of Concerned Scientists. https://www.ucsusa.org/sites/default/files/2022-
09/driving-cleaner-report.pdf
43 U.S. EPA (2022 December 13). Public data for the renewable fuel standard. https://www.epa.gov/fuels-registration-

reporting-and-compliance-help/public-data-renewable-fuel-standard
15
conversion factors44. The CO2 emissions targets are shown for both enacted and proposed
targets for each country, where applicable including for the U.S.

300 8
CO2 emissions (g/mile) for passenger cars, normalized to CAFE cycle

7
250

Fuel consumption (l/100 km gasline equivalent)


6
Brazil 2022:
201
200
India 2022: 5
179

China 2025: 153


150 4
Chile 2030: 135
Mexico 2025: 143

Canada 2026: 128 Japan 2030: 124


S Korea 2030: 113 3
100
New Zealand 2027: 106

2
US 2032: 73

50
1
Historical performance
Enacted targets UK 2035: 0
Proposed targets
EU 2035: 0
0 0
2010 2015 2020 2025 2030 2035
Note: U.S. 2026 target is adjusted in proposed standards for 2027 and later vehicles to reflect differences in various credits between the current
rule and the proposal; UK fleet-average targets estimated based on non-ZEV CO2 emissions and ZEV mandate.

Figure 5. Global comparison of historical data and standards for passenger car CO2 emissions
and fuel consumption, normalized to CAFE cycle.

44 Kühlwein, J. et al. (2014). Development of test cycle conversion factors among worldwide light-duty vehicle CO2
emission standards. International Council on Clean Transportation. https://theicct.org/publication/development-of-
test-cycle-conversion-factors-among-worldwide-light-duty-vehicle-co2-emission-standards/; Yang, Z. (2014).
Improving the conversions between the various passenger vehicle fuel economy/CO2 emission standards around
the world. International Council on Clean Transportation. https://theicct.org/improving-the-conversions-between-
the-various-passenger-vehicle-fuel-economy-co2-emission-standards-around-the-world/
16
400

10
CO2 emissions (g/mile) for light trucks, normalized to CAFE cycle

350

Fuel consumption (l/100 km gasline equivalent)


300 8

250
Mexico 2025: S Korea 2030: 235
228 6

Japan 2022: 211


200
Canada 2026: 190

150 4
Chile 2030: 137
New Zealand 2027: 144

100
US 2032: 89
2

50
Historical performance
Enacted targets UK 2035: 0
Proposed targets
EU 2035: 0
0 0
2010 2015 2020 2025 2030 2035
Note: U.S. 2026 target is adjusted in proposed standards for 2027 and later vehicles to reflect differences in various credits between the current
rule and the proposal; UK fleet-average targets estimated based on non-ZEV CO2 emissions and ZEV mandate.

Figure 6. Global comparison of historical data and standards for light truck CO 2 emissions and
fuel consumption, normalized to CAFE cycle.

The historical trend shows that the U.S. CO2 emissions were historically one of the highest
across the global markets, which began to reduce substantially through stringent GHG
emissions standards starting from 2012. While the GHG emissions standards have significantly
contributed to controlling emissions of the fleet over the years, the recent U.S. CO2 emissions
and enacted targets are still higher than the average trajectories for a few markets such as the
EU, Japan, South Korea, and UK for passenger cars and Chile, Japan, and UK for light trucks.

The proposed standards for 2027 and later model year vehicles make the U.S. one of the most
ambitious, forward-looking markets for CO2 standards. The proposed standard levels are more
stringent compared to most of the global markets, except for the EU and UK. The EU has
already adopted 0 g CO2 per km standard by 2035. The UK targets are still in the proposal
phase; however, the UK has set strong ZEV mandates for 100% ZEV sales of light-duty
vehicles by 2035.

Thus, the proposed U.S. targets are more ambitious compared to the current trajectories for
most of the global markets, yet not stringent enough to align with the most ambitious
international targets.

Electric vehicle uptake


Figure 7 shows the historical EV sales shares from 2015 through 2022 (solid lines) and
estimated future sales shares based on regulatory targets (dotted lines) for light-duty vehicles in
17
the United States compared with other major markets. As shown, about 7% of new light-duty
vehicle sales in the US. Were plug-in electric in 2022, compared to 8% in Canada, 20% in
California, 23% in the United Kingdom and Europe, and 26% in China. Several jurisdictions
have set a 100% ZEV target or a zero gram CO2/mile target by 23035, including California,
Canada, European Union, and the United Kingdom (UK). 45 The figure shows how the estimated
67% EV sales share in 2032, the last year of the proposed EPA standard, somewhat lags
compared to the estimated uptake from regulations in California, Canada, Europe, and the UK.
Not shown is the 46% EV sales share EPA projects for medium-duty vehicles in 2032.

Figure 7. Historical electric vehicle sales shares (solid lines) and regulatory targets or projected
ZEV shares consistent with regulatory targets (dotted lines) for light-duty vehicles in major
markets

As discussed above and shown in Figure 7, at the state-level, California’s Advanced Clean Cars
II regulation will lead to 68% electric vehicle sales by 2030 and 100% by 2035. At least a dozen
additional states are considering adopting California’s new rules. As of May 2023, California,
Massachusetts, New York, Oregon, Vermont, Virginia, and Washington have adopted ACC II. It
is likely that many other states will continue to follow California’s leadership and adopt the new
ACC II Program to benefit from the anticipated emissions reduction and health benefits of the
program.46 Many additional states currently follow the Advanced Clean Cars regulations through
model year 2026; as of May 13th, 2022, 17 U.S. states have adopted all or part of California’s
low-emission and zero-emission vehicle regulations, and about 37% of national new light-duty
vehicle sales meet California’s emission standards.47

COMPARISON WITH AUTOMAKER COMMITMENTS


In the proposed rule, EPA highlighted the proliferation of announcements by automakers in the
past couple of years signaling a market shift away from internal-combustion technologies toward

45 European Union and UK’s proposed standards do not include PHEV in ZEV definition
46 Houk, J., Huang, J., and Sussman. (2023). Benefits of adopting California’s Advanced Clean Cars II standards in
sixteen U.S. states. Sonoma Technology Inc. Retrieved from https://theicct.org/publication/benefits-of-state-level-
adoption-of-california-acc-ii-regulations/
47 California Air Resources Board. (May 13, 2022). States that have adopted California’s vehicle standards under

Section 177 of the Federal Clean Air Act. https://ww2.arb.ca.gov/sites/default/files/2022-


05/%C2%A7177_states_05132022_NADA_sales_r2_ac.pdf
18
plug-in electric vehicles. The rapid growth in the global vehicle market shows that more
automakers are committed to developing and investing in ZEV technologies at a global and
regional level. ICCT recognizes the increasing number of commitments made by automakers
and the speed at which the automotive industry is electrifying its fleet.

ICCT previously collected information and analyzed announcements made by automakers on


ZEV targets through the end of 2022. The analysis, from ICCT’s 2022 Global Automaker Rating
report,48 shows that 19 out of 20 top global automakers have made specific commitments on
their pace of electrification.49 For example, General Motors and Ford Motors (the second and
third largest automakers in the U.S by 2022 sales), as well as Mercedes-Benz have committed
to cease internal combustion engine in leading markets by 2035 as COP26 ZEV declaration
signatories.50

With the rapid development of ZEV technologies, automakers are constantly updating their
targets. Consequently, ICCT recommends the EPA to incorporate more recent ZEV target
announcements in Table 2 (right column) in addition to those outlined in page 24-28 of section II
of the proposed rule (left column). We believe these developments further reflect the desirability
of ZEV technologies in the market.

Table 2. Announcements ZEV sales target by automakers51


Automaker Announcements outlined in the proposed rule Additional announcements as of 2022
In November 2021, GM signed the COP26 ZEV
GM In January 2021, General Motors announced plans to
declaration aiming for a 100% ZEV sales target
shift its light-duty vehicles entirely to zero-emissions
for LDVs in leading markets by 2035 including
by 2035.
the U.S.52
In March 2021, Volvo announced plans to make only
Volvo Cars No update
electric cars by 2030.
In March 2021, Volkswagen announced that it
VW expected 50% of its U.S. sales to be all-electric by No update
2030.
In April 2021, Honda announced a full electrification
plan to take effect by 2040, with 40% of North
Honda No update
American sales expected to be ZEV by 2030, 80% by
2035 and 100% by 2040.
In November 2021, Ford signed the COP26
In May 2021, Ford announced that they expect 40% ZEV declaration aiming for a 100% ZEV sales
Ford
of their global sales will be all-electric by 2030. target for LDVs in leading markets by 2035
including the U.S.53

48 Shen, C., Fadhil, I., Yang, Z., Searle, S. (2023). The global automaker rating 2022: who is leading the transition to
electric vehicles? International Council on Clean Transportation. https://theicct.org/publication/the-global-
automaker-rating-2022-may23/.
49 The top 20 automakers are selected based on the total light-duty vehicle sales in 2022 across six major markets:

China, United States, Europe, Japan, India, and South Korea. About 90% of their global LDV sales were delivered
in these six major markets.
50 Ford Motors. (2021). Ford statements on signing the ambitious route zero initiative at COP26. Ford: Cologne,

Germany. November 2021. https://media.ford.com/content/fordmedia/feu/en/news/2021/11/10/ford-statements-on-


signing-the-ambitious-routezero-initiative-at.html
51 Announcements are targets reported through the end of 2022 and cover ZEV sales target in the U.S., leading

markets, and/or global sales.


52 COP26 ZEV declaration. (2021, November). Automotive manufacturer signatories. (COP26: Glasgow).

https://cop26transportdeclaration.org/en/signatories/automotive-manufacturers/.
53 COP26 ZEV declaration. (2021, November). Automotive manufacturer signatories. (COP26: Glasgow).

https://cop26transportdeclaration.org/en/signatories/automotive-manufacturers/.
19
In June 2021, Fiat announced a move to all electric
In March 2022, Stellantis announced a 50%
vehicles by 2030, and in July 2021 its parent
Stellantis ZEV target by 2030 for LDVs in the United
corporation Stellantis announced an intensified focus
States.54
on electrification across all of its brands.
In July 2021, Mercedes-Benz announced that all its In November 2021, Mercedes-Benz signed the
Mercedes- new architectures would be electric-only from 2025, COP26 ZEV declaration aiming for a 100% ZEV
Benz with plans to become ready to go all-electric by 2030 sales target for LDVs in leading markets by
where possible. 2035 including the U.S.55
In December 2021, Toyota is committed to
In December 2021, Toyota announced plans to
Toyota selling 3.5 million BEVs globally by 2030. This
introduce 30 BEV models by 2030 worldwide.
is roughly 32% of its LDV sales.56
In March 2022, Hyundai57 and Kia58 are
Hyundai- Hyundai and Kia announced a global ZEV target of
committed to 36% and 30% global ZEV sales
Kia 50% and 45% by 2030 respectively.
by 2030.
Mazda announced a global ZEV target of 25% by
Mazda No update
2030.
Nissan announced a ZEV target of 40% by 2030 in
Nissan No update
the United States.
Jaguar announced a global ZEV target of 100% by
Jaguar
2025.
No update
Land Rover announced a global ZEV target of 60%
Land Rover
by 2030.

The increased ambition in announced ZEV targets among automakers is observed to be in


parallel with the growing number of regulations proposed by governments across the globe
pledging for zero-emission transportation policies and transitioning to a ZEV fleet. The
implementation of regulations in these jurisdictions that establish more stringent CO2 standards
or require higher ZEV penetration can steer automakers to prioritize investments and ZEV
deployment in these markets to comply. Thus, regulatory mechanisms can be one of the driving
factors to accelerate ZEV penetration in the new vehicle fleet.

In the case of the European Union, the CO2 emission standards59, adopted in March 2023, sets
a more stringent CO2 emission performance standards for LDVs, with a requirement for zero-
CO2 LDVs in 2035. Aligned with this, automakers operating in the region have pledged more

54 Stellantis. (2022, March 1). Long-term strategic plan: Dare forward 2030. Stellantis: Amsterdam.
https://www.stellantis.com/content/dam/stellantis-corporate/investors/events/strategic-plan-
2030/2022_03_01_Strategic_Plan.pdf
55 COP26 ZEV declaration. (2021, November). Automotive manufacturer signatories. (COP26: Glasgow).

https://cop26transportdeclaration.org/en/signatories/automotive-manufacturers/.
56 We estimate the ZEV sales share of the total LDV in 2030 based on Toyota’s absolute ZEV sales target number of

3.5 million BEVs by 2030 and project the global LDV sales in 2030 based on Toyota’s 2011-2022 annual CAGR
from its worldwide production. Retrieved from https://global.toyota/en/company/profile/production-sales-
figures/202212.html
57 Hyundai. (2022, March). 2022 CEO Investor day. Hyundai: Seoul.

https://www.hyundai.com/content/hyundai/ww/data/ir/calendar/2022/0000000352/files/2022-ceo-investor-day-eng-
220620.pdf.
58 Kia Media. (2022, March). Kia CEO Investor Day presents 2030 roadmap to become global sustainability mobility

leader. Kia: Seoul. https://press.kia.com/eu/en/home/media-resouces/press-releases/2022/Kia-CEO-Investor-


Day.html.
59 EU Press release. (2023). Fit for 55: Council adopts regulation on CO2 emissions for new cars and vans. March

2023. EU; Brussels. https://www.consilium.europa.eu/en/press/press-releases/2023/03/28/fit-for-55-council-


adopts-regulation-on-co2-emissions-for-new-cars-and-vans/
20
ambitious ZEV targets or commitment to electrify their new sales60. For example, Volkswagen
(VW) and Stellantis, the two biggest automakers in Europe (by sales), have pledged to sell
80%61 and 100% ZEVs, respectively, by 2030. This sales target is more ambitious than their
targets in other regions including the U.S in which VW and Stellantis are committed to a 50%
ZEV target by 2030 for its new passenger cars and light-duty vehicles respectively. Another
example is Toyota, which aims for 100% CO2 emission reductions in all new vehicle sales in
Europe by 2035 with at least 50% ZEV mix in 2030.62 In addition to the global ZEV target sales,
Europe is the only region where Toyota is ready to commit to cease its ICE sales.

The trend observed among automakers provides evidence of the suitability of ZEVs as
emissions-reducing technologies. The developments outlined above further underscore the fact
that ZEV technology is feasible and desirable as evidenced by the increasing fleet penetration.
Automakers will need to transition toward ZEVs to stay on track as government policies
continued to push for regulations that align with the Paris Agreement climate goals and as
market forces continue to push in this direction.

This changing reality is already reflected in new LDV sales. The share of EVs has been growing
rapidly in leading markets such as China and Europe at 24% and 21%, respectively in 2022.
The demand for these vehicles in the U.S., although lagging, also increased rapidly, even under
the current standards. In 2022, PEV accounted for approximately 7.3% of total LDVs sold 63, an
increase from 4.8% in 2021 and 2.6% in 2020. 64 In terms of absolute number, total PEV sales in
2022 nearly reached 1 million with 80% of sales being BEVs, and 20% PHEVs. With the right
supporting regulatory mechanism, the proposed regulation will enable the U.S. to achieve the
ZEV penetration rate needed to achieve the emission reductions projected under the proposed
scenario.

In conclusion, a strong market force to advance ZEV technologies, as evidenced by the


increasing commitment among automakers and a more stringent standards in place such as the
proposed rule, can push the automotive industry to accelerate ZEV technology penetration in
the fleet. ICCT supports the proposition of EPA to advance technologies that will further enable
emission reductions and consequently the proposed rule needed to achieve those targets.

BATTERY ELECTRIC VEHICLE COST


While ICCT strongly supports the proposed rule, we believe there is evidence available to
support lower BEV costs than EPA has modeled. EPA analysis shows clear and significant
benefits associated with achieving a reduced combined fleet gram CO2/mile standard and an
associated increase in battery electric vehicle penetration. However, the cost of compliance may

60 Yang, Z. (2022). Beyond Europe: are there ambitious electrification targets across major markets? International
Council on Clean Transportation. https://theicct.org/global-oem-targets-cars-ldvs-nov22/
61 Earlier this year, VW announced a higher target of 80% by 2030 (https://europe.autonews.com/automakers/vw-

brand-targets-ev-sales-80-percent-2030).
62 Toyota Europe Newsroom. (2021). Toyota Motor Europe outlines its path to 100% CO2 reduction by 2035.

December 2021. Toyota: Brussels. https://newsroom.toyota.eu/toyota-motor-europe-outlines-its-path-to-100-co2-


reduction--by-2035/#
63 Fadhil, I., Shen, C. (2023). Electric vehicles market monitor for light-duty vehicles: China, Europe, United States,

and India, 2022. International Council on Clean Transportation. https://theicct.org/publication/ev-ldv-major-


markets-monitor-2022-jun23/.
64 Fadhil, I., Chu, Y., Jin, L. (2023). Electric Vehicle Market Monitor for light-duty vehicles: China, Europe, United

States, and India, 2020 and 2021. International Council on Clean Transportation. https://theicct.org/publication/ev-
ldv-major-markets-monitor-jan23/.
21
be overstated due to the use of outdated technology data and information. Electric vehicle and
battery technology has been improving rapidly and technology costs have been greatly reduced.
Automakers are investing heavily in BEV R&D and manufacturing capacity and are achieving
higher production volumes with more advanced technologies at lower costs. We reviewed
EPA’s analysis on electric vehicle and battery costs and identified a few key parameters within
EPA’s analysis that we believe could be updated to better reflect the latest evidence and
analysis. We make observations below that would reduce the incremental BEV costs and
accelerate the timing for cost parity, which further strengthen the case for adopting the proposed
rule and would result in relatively greater benefits at lower cost.

New 2022 ICCT research assessed light-duty electric vehicle costs and consumer benefits in
the United States in the 2022-2035 time frame and found that without any federal, state, utility,
or local incentives, BEV purchase price parity is coming before 2030 for BEVs with up to 300
miles of range across all light-duty car, crossover, SUV, and pickup truck classes.65 Continued
technological advancements and increased battery production volumes mean that pack-level
battery costs are expected to decline to about $105/kWh by 2025 and $74/kWh by 2030. These
developments are critical to achieving electric vehicle initial price parity with conventional
vehicles, which the 2022 ICCT analysis finds to occur between 2024 and 2026 for 150- to 200-
mile range BEVs, between 2027 and 2029 for 250- to 300-mile range BEVs, and between 2029
and 2033 for 350- to 400-mile range BEVs. These results—along with others discussed below—
from ICCT’s 2022 EV cost parity study are aligned with those found in similarly recent studies of
EV cost parity.66 The cost parity findings are further reinforced by new Energy Innovation and
Consumer Reports research showing that in 2023 most new electric vehicles are already
cheaper to own than gasoline-powered vehicles in the United States from day one.67

EPA’s OMEGA model also includes an analysis of light-duty electric vehicle costs. ICCT
examined the BEV cost data from EPA’s OMEGA output files and compared the overall costs
and the costs of various battery and vehicle components and technical specifications with those
applied in the 2022 ICCT U.S. EV cost study for industry-average 300-mile range BEVs.
Specifically, we compared the component-level costs and specifications broadly categorized
into powertrain direct, non-powertrain direct, and indirect costs; doing so allowed for the most
appropriate and direct comparisons with the findings in the ICCT study. The Figure 8 flowchart
below illustrates the BEV cost comparison approach based on EPA data from the OMEGA
output file and the ICCT data.

65 Slowik. P., Isenstadt, A., Pierce, L, Searle, S. (2022). Assessment of light-duty electric vehicle costs and consumer
benefits in the United States in the 2022-2035 time frame. International Council on Clean Transportation.
https://theicct.org/publication/ev-cost-benefits-2035-oct22/
66 H. Saxena, V. Nair, S. Pillai, “Electrification Cost Evaluation of Light-Duty Vehicles for MY 2030,” 2023. Roush.

https://www.edf.org/sites/default/files/2023-05/Electrification_Cost_Evaluation_of_LDVs_for_MY2030_Roush.pdf
and Transport and Environment. (2021). Hitting the EV inflection point. Retrieved from
https://www.transportenvironment.org/discover/hitting-the-ev-inflection-point/
67 Orvis, R. (2022). Most electric vehicles are cheaper to own off the lot than gas cars. Energy Innovation.

https://energyinnovation.org/wp-content/uploads/2022/05/Most-Electric-Vehicles-Are-Cheaper-Off-The-Lot-Than-
Gas-Cars.pdf and Consumer Reports (2023). Electric vehicles save consumers money.
https://advocacy.consumerreports.org/research/cr-fact-sheet-electric-vehicles-save-consumers-money/
22
EPA data (from output file) ICCT data

Electrified driveline cost ÷ 1.5 other powertrain direct cost

e-machine cost ÷ 1.5 BEV motor direct cost

Powertrain
battery cost ÷ 1.5 direct cost BEV battery direct cost

driveline cost ÷ 1.5 driveline direct cost

engine cost ÷ 1.5 engine direct cost

initial MSRP ÷ RPE -


glider non-structure cost ÷ 1.5 Non- powertrain direct
Powertrain
structure cost ÷ 1.5 direct cost

other indirect
powertrain cost × 0.5

Indirect cost OEM profit margin


non-powertrain cost × 0.5

Dealer markup

Manufacturing
Total cost
cost (output
=
datapoint)
Vehicle price

Figure 8. Illustration of methodology for EPA vs. ICCT BEV-300 cost comparison

To calculate EPA's modeled total manufacturing cost – including all indirect costs using the 1.5
retail price equivalent (RPE) factor – the total costs of individual vehicle elements were
summed. These individual vehicle elements are electrified driveline, e-machine, battery,
driveline (conventional vehicle), engine, glider (non-structure), structure (i.e., chassis/frame).
The first five cost elements are powertrain-specific. The last two are non-powertrain. Some
vehicles, such as hybrids, have all five powertrain cost elements. All vehicles have both non-
powertrain cost elements.

23
To determine the industry average total costs for each cost component, the individual vehicle
model costs were averaged, weighted by model sales. Critically, BEV and ICE platforms are
kept separate by only considering costs associated with designated ICE or BEV vehicles, as
appropriate. For instance, in calculating the average ICE vehicle battery cost, only ICE vehicle
models were considered in the average calculation. Once the average total costs by component
are obtained, they are summed into powertrain and non-powertrain (described above), and then
summed into full vehicle total costs. Lastly, the direct costs for each cost component and for the
full vehicle can be calculated by dividing total costs by 1.5.

While the OMEGA model uses “total manufacturing costs” described above as an input to
calculate final vehicle price, for the purpose of comparison with the ICCT 2022 EV cost analysis,
these total costs represent the equivalent of “price” as defined and analyzed by ICCT (see the
flowchart above).

Figure 9 illustrates the key findings from our comparison of the EPA and ICCT BEV cost
analyses. The figure shows the upfront incremental BEV costs for cars, crossovers, SUVs, and
pickup trucks relative to their combustion counterparts for a 300-mile range BEV. The top figure
shows the incremental BEV costs in 2027 and the bottom figure shows the incremental BEV
costs in 2032. The blue bars represent EPA’s findings of BEV incremental costs and the orange
bars represent the ICCT findings of incremental costs. As shown, ICCT finds that incremental
BEV costs are much lower than estimates by EPA by about $2,300 (pickups) to about $5,300
(crossovers) in 2027. As battery and electric vehicle costs fall and technology improves, the
incremental costs are reduced. For 2032, the EPA finds that BEVs have incremental costs of
about $3,000 to $5,000; during this same timeframe, the ICCT study finds that BEVs will be
cheaper to purchase than comparable gasoline vehicles, which is indicated by the negative
incremental costs shown in the bottom figure.

24
$7,000
2027
Incremental BEV cost (2027)

$6,000

$5,000

$4,000
EPA
$3,000 ICCT

$2,000

$1,000

$0
Car Crossover SUV Pickup
$8,000
2032
Incremental BEV cost (2032)

$6,000

$4,000

$2,000
EPA
$0 ICCT
Car Crossover SUV Pickup
-$2,000

-$4,000

-$6,000
Figure 9. Comparison of EPA and ICCT (2022) findings of incremental BEV costs compared to
ICEV costs, for 2027 (top) and 2032 (bottom)

The differences in overall incremental cost findings can be explained by comparing the
assessments of combustion and electric vehicle costs. For combustion vehicles, on average,
the costs in the ICCT analysis are about $500 less than EPA for 2027, and about $1,500 more
than EPA for 2032. This shift is for two reasons: 1) the ICCT study applies cost-effective fuel
savings technology to combustion vehicle models that improve the new vehicle fleet by about
3.5% per year and increase powertrain direct costs by about 1% per year, and 2) the average
combustion vehicle prices in the EPA analysis decline by about $1,400 from 2027 to 2032,
along with their reduced efficiency and increase in per-mile emissions. For battery electric
vehicles, the average BEV costs in the ICCT analysis are about $5,000 less than EPA for 2027
and about $7,000 than EPA for 2032. The ICCT analysis finds that BEV costs can be reduced
faster than the EPA projections. We further assessed BEV costs by powertrain direct, non-
powertrain direct, and indirect cost elements (Figure 8).

Total battery pack costs on average in the ICCT analysis are about $3,700 less than in EPA’s
analysis in 2027 and about $5,200 less than in EPA’s analysis in 2032. This is due to several
factors, including average BEV pack size (kWh), efficiency (kWh/mile) and pack-level cost
($/kWh). Table 3 summarizes the findings of the underlying technical specifications for 300-mile
range BEVs in 2027, 2030, and 2032 in the EPA and ICCT analyses. As shown, on average in
the ICCT analysis the pack size is smaller for the same range, the efficiency is higher, and the

25
per-pack costs are lower. As shown in the bottom row, the total pack cost in the ICCT study is
about $3,200 to about $4,400 less than EPA in 2027, about $2,800 to $4,100 in 2030, and
about $4,200 to $6,000 in 2032. The pack cost per kWh in the EPA analysis increase from 2030
to 2032 along with the phaseout of the 45X battery production tax credit. The pack cost per kWh
in the ICCT analysis do not include the 45X tax credit; applying the 45X tax credit to the ICCT
battery costs would further reduce BEV costs and the gap between the findings of incremental
costs in the ICCT and EPA analyses (Figure 9).

Table 3. Summary of key BEV technical specifications in EPA and ICCT (2022)
2027 2030 2032
EPA ICCT Difference EPA ICCT Difference EPA ICCT Difference
Car 87 71 16 86 64 22 86 63 23
Pack size Crossover 103 78 25 103 67 35 103 67 36
(kWh) SUV 125 93 32 125 82 43 124 81 44
Pickup 135 117 18 135 104 31 135 103 32
Car 0.27 0.24 0.02 0.27 0.22 0.04 0.27 0.22 0.04
On-road Crossover 0.33 0.27 0.06 0.33 0.24 0.09 0.33 0.23 0.10
efficiency
(kWh/mile) SUV 0.40 0.32 0.08 0.40 0.28 0.12 0.40 0.28 0.12
Pickup 0.44 0.40 0.04 0.44 0.36 0.08 0.45 0.36 0.09
Car 108 87 21 86 71 15 96 65 31
Pack cost Crossover 103 86 17 82 71 11 93 65 28
($/kWh) SUV 98 84 14 78 69 9 89 63 26
Pickup 96 82 14 76 67 9 88 61 27
Car $9,398 $6,224 $3,174 $7,410 $4,549 $2,861 $8,290 $4,097 $4,193
Total pack Crossover $10,633 $6,766 $3,866 $8,436 $4,783 $3,653 $9,537 $4,308 $5,229
cost (kWh
* $/kWh) SUV $12,249 $7,851 $4,397 $9,745 $5,648 $4,097 $11,097 $5,087 $6,010
Pickup $12,951 $9,564 $3,387 $10,322 $6,983 $3,340 $11,864 $6,290 $5,575
Note. Numbers in table are rounded

The combination of reduced per kWh costs and smaller pack size for the same range (due to
improved efficiency) in the ICCT analysis is the driver for lower costs compared to EPA. The
ICCT conducted a thorough battery cost review and applied the best available data and analysis
of current and future electric vehicle and battery technical specifications. 68 Figure 10 shows the
ICCT’s 2022 battery cost review, which was based on expert sources, research literature
projections, and automaker announcements. Our battery cost review includes the most recent
projections by expert sources including the California Air Resources Board (2022), Roush
Industries Inc. (see Saxena, Stone, Nair, & Pillai, 2023), Bloomberg New Energy Finance (2020,
2021), UBS (2020) and technical research studies, including Mauler, Lou, Duffner, and Leker
(2022), Nykvist, Sprei, and Nilsson (2019), Penisa et al. (2020), Hsieh, Pan, Chiang, and Green
(2019), and Berckmans et al. (2017). The automaker announcements shown include
Volkswagen for $135 per kilowatt-hour in 2021–2022 (Witter, 2018), Tesla for $55/kWh in 2025
(Tesla, 2020), and Renault and Ford for $80/kWh in 2030 (Automotive News, 2021a, 2021b;
Ford, 2021). Not shown due to uncertainties related to timing, General Motors in 2020
announced continued improvements toward below $100/kWh at the cell level, and Volkswagen

68 Slowik. P., Isenstadt, A., Pierce, L, Searle, S. (2022). Assessment of light-duty electric vehicle costs and consumer
benefits in the United States in the 2022-2035 time frame. International Council on Clean Transportation.
https://theicct.org/publication/ev-cost-benefits-2035-oct22/
26
in 2021 announced developments toward “significantly below” $100/kWh at the pack-level
(General Motors, 2020; Volkswagen, 2021).69

Figure 10. Electric vehicle battery pack costs from technical studies and automaker statements

Figure 10 also shows the battery pack cost per kWh applied in EPA’s OMEGA analysis,
illustrated by the brown hashed line. The EPA (2023) cost curve in the figure is based on the
volume weighted average pack marked-up cost from EPA Figure 2-27 from the DRIA with the
value of the 45X battery production tax credit incentive removed so that it can be directly
compared to the ICCT’s 2022 battery cost review.70 The EPA (2023) battery cost curve clearly
stands out with much higher per-kWh costs than nearly every other study identified in the ICCT
battery cost review. On average over the 2023-2035 timeframe, the annual per kWh battery
pack costs in the EPA model are about 55% greater than those applied in ICCT (2022). To
provide context to these values, industry surveys showed that the volume-weighted average
global BEV pack-level prices were approximately $126 per kWh in 2020 and $118 per kWh in
2021.71 EPA is applying about $126 kWh in 2029 and about $115 per kWh in 2030, which is
about nine years later than the 2020 and 2021 volume-weighted average global pack prices
reported by BNEF.

For BEV efficiency, the technical specifications applied in the ICCT study are based on official
electric vehicle range and efficiency values from the U.S. Department of Energy and reflect
consumer label efficiency.72 The ICCT study applies a charging efficiency factor of 93% for all

69 Complete referencing is available at https://theicct.org/publication/ev-cost-benefits-2035-oct22/


70 EPA (2023). Multi-pollutant emissions standards for model years 2027 and later light-duty and medium-duty
vehicles. Draft Regulatory Impact Analysis. https://nepis.epa.gov/Exe/ZyPDF.cgi?Dockey=P10175J2.pdf
71 Bloomberg New Energy Finance. (2020). Battery pack prices cited below $100/kWh for the first time in 2020, while

market average sits at $137/kWh. Retrieved from https://about.bnef.com/blog/battery-pack-prices-cited-below-100-


kwh-for-the-first-time-in-2020-while-market-average-sits-at-137-kwh/ and Bloomberg New Energy Finance. (2021).
Battery pack prices fall to an average of $132/kWh, but rising commodity prices start to bite. Retrieved from
https://about.bnef.com/blog/battery-pack-prices-fall-to-an-average-of-132-kwh-but-rising-commodity-prices-start-to-
bite/
72 United States Department of Energy. (2022b). “Download fuel economy data”. Retrieved from

https://www.fueleconomy.gov/feg/download.shtml
27
years. A useable-to-total battery pack size ratio is also applied based on average high-volume
MY 2022 electric vehicles such that BEVs can use 92% of the kWh, which increases for new
vehicles by less than 1% per year through 2030, based on the best available models from 2022.
For context, several BEV models including the BMW i4, Chevrolet Bolt EV, Chevy Bolt EUV,
Hyundai Ioniq 5, Nissan Leaf, Polestar 2, and Volvo C40 and XC40 have a useable-to-total
battery ratio of 96% or greater in 2022.

The initial 2022 electric vehicle efficiencies applied in the ICCT study are based directly on
existing MY 2022 BEV and PHEV models, accounting for increased electricity-per-mile for
longer-range electric vehicles. We apply average technical specifications based on several high-
volume MY2022 electric vehicle models within each class. For example, our BEV crossover
efficiency is based on the Tesla Model Y, Ford Mach-e, Volkswagen ID 4, Hyundai Kona, Kia
Niro, Kia EV6, and Volvo XC40. Electric vehicle efficiency improves annually due to electric
component (battery, motor, power electronic) and vehicle-level (mass reduction, aerodynamic,
and tire rolling resistance) improvements. The 2030–2035 electric vehicle efficiencies are based
on modeling by CARB (2022), accounting for range and adjusting for charging losses. 73
Between 2022 and 2030, we apply an average annual improvement that links the high-volume
2022 average electric vehicle model specifications with the 2030 CARB values. By 2030, the
efficiencies are somewhat better than those of the “best in class” models from 2022. For
example, our representative 350-mile range battery electric car is 0.23 kWh/mile compared to
the 358-mile range 2022 Tesla Model 3 at 0.26 kWh/mile. Our representative 350-mile range
crossover in 2030 is 0.25 kWh/mile, compared to the 330-mile range 2022 Tesla Model Y at
0.28 kWh/mile.

In addition to the powertrain battery costs, we found that the non-battery powertrain costs in the
ICCT analysis are about $600 to $900 less than the EPA modeling in 2027, and that gap
declines to about $400 to $600 less than EPA in 2032. In the ICCT analysis, non-battery costs
scale with power and are based on several sources. Non-battery powertrain costs are assessed
primarily based on a teardown analysis by UBS (2017) and the National Academies of Sciences
Engineering and Medicine fuel economy technology assessment (NASEM, 2021).74 Virtually all
electric vehicles equipped with AWD do so with additional motors, rather than electronic AWD or
another AWD system used on combustion vehicles. By matching electric and combustion
vehicle power, combined motor power for electric vehicles with multiple motors is the same as
the power for single motor vehicles. With additional motors, costs for high voltage cables and
motor cooling increase. It is unclear from literature whether motor costs include driveshaft,
which would also increase with the number of motors. According to NASEM, future permanent
magnet motor costs are expected to decline due to reduced magnetic material requirements.
These future costs scale proportionally with motor power, suggesting that certain cost elements
that increase with motor number are not included. Further investigation into the true costs of
BEV AWD is beyond the scope of this paper. However, manufacturers may opt for induction
motors as a second motor in AWD configurations. Absent permanent magnets, induction motors
have the potential to decrease AWD costs further, even below the future permanent magnet
motor costs shown in NASEM.

73 California Air Resources Board. (2022). Advanced Clean Cars II Rulemaking: ZEV Cost Modeling Workbook March
2022. https://ww2.arb.ca.gov/our-work/programs/advancedclean-cars-program/advanced-clean-cars-ii
74 UBS. (2017). UBS evidence lab electric car teardown: Disruption ahead? [Q-Series newsletter]. Retrieved from

https://neo.ubs.com/shared/d1ZTxnvF2k/ and National Academies of Sciences, Engineering, and Medicine.


(2021). Assessment of Technologies for Improving Light-Duty Vehicle Fuel Economy—2025-2035. Washington,
DC: The National Academies Press. https://doi.org/10.17226/26092
28
Table 4 summarizes the findings of the underlying technical specifications for motor power in
kilowatts for 300-mile range BEVs and combustion vehicles in 2027, 2030, and 2032 in the EPA
and ICCT analyses. As shown, the BEV motor power assumed in the ICCT analysis is
significantly less than what is applied by EPA. In the ICCT study, matching electric and
combustion vehicle power means that the combined motor power for electric vehicles with
multiple motors is the same as the power for single motor vehicles. From the table below, EPA
applies BEV motor power that is much greater than comparable combustion vehicles. We
compared EPA’s assumed motor power on a model-by-model basis for BEVs and their
combustion counterparts and found that on average BEV motor power is at least 38% (cars) to
45% (pickups) greater than the assumed motor power of the ICE variants. Because the non-
battery powertrain costs in EPA's analysis directly scale with motor power, reducing the
assumed BEV motor power (kW) in EPA’s analysis to be similar to the motor power applied for
combustion vehicles – based on evidence from the ICCT analysis – would reduce the BEV non-
battery powertrain costs in EPA’s analysis.

Table 4. Summary of fleet average motor power (kW) applied in EPA and ICCT (2022)
2027 2030 2032
EPA ICCT Difference EPA ICCT Difference EPA ICCT Difference
Car 210 153 58 205 153 52 204 153 51
Crossover 230 146 84 227 146 81 227 146 81
BEV
SUV 310 227 82 311 227 84 311 227 84
Pickup 364 253 111 362 253 109 361 253 108
Car 159 153 7 160 153 7 150 153 -2
Crossover 175 146 29 171 146 25 169 146 23
ICE
SUV 258 227 31 271 227 44 266 227 39
Pickup 283 253 31 268 253 15 265 253 12

Very recent BEV-related automaker announcements and developments from automotive


suppliers reflect the continued rapid pace of innovation in BEV technologies. For example,
Toyota recently announced its 2026 plans to produce battery packs capable of over 600 miles of
range at lower cost than its current packs.75 Toyota also plans to produce solid-state battery
packs with 20%-50% more range, and bipolar lithium iron phosphate packs at 40% lower cost
starting around 2028.76 Beyond reducing costs through battery improvements and vehicle
efficiency increases, Toyota plans to reduce vehicle manufacturing costs through improvements
in factory operation, giga-casting to reduce parts and complexity, and reduced development
times.77 Similarly, Volvo announced its plans to reduce EV costs through cell-to-body battery
packaging, integrated motor, inverter, transmission units, and giga-castings.78 Meanwhile, the
advent of two-speed transmissions for BEVs from automotive suppliers increases motor
efficiency and performance, allowing battery and motor downsizing.79 Another supplier
innovation that improves EV efficiency especially for induction motors is Tula’s Dynamic Motor

75 Greimel, H. (2023, June 12). Newly revealed Toyota EV plans include batteries with 900-plus miles of range.”
Automotive News. https://www.autonews.com/mobility-report/toyota-future-ev-plans-include-batteries-900-mile-
range
76 Ibid.
77 Ibid.
78 Karkaria, U. (2023, June 14). Volvo EX30 electric crossover will approach price parity. Automotive News.

https://www.autonews.com/retail/how-volvos-electric-ex30-crossover-will-approach-price-parity
79 Brooke, L. (2021, February). Gearing EVs for greater efficiency.” SAE Automotive Engineering.

https://www.nxtbook.com/smg/sae/21AE02/index.php#/p/16
29
Drive.80 Still other areas of innovation include higher voltage EV platforms, axial flux motor
improvements, and current-source inverters.81 These recent examples go beyond the
technology improvements assumed in the 2022 ICCT EV cost study. With these innovations,
BEVs become even more cost-effective as a technology option for manufacturers.

Based on the above analysis, we believe that the relative BEV costs may be overstated in the
EPA proposal and accompanying OMEGA analysis. If EPA were to update the analysis with the
data and evidence regarding BEV cost and technical specifications based on the ICCT cost
study, BEV costs would be lower, the costs of compliance would be lower, and the net benefits
would be greater.

The above analyses and comparisons were conducted almost entirely relying on OMEGA
output files, which can obscure important modeling assumptions. To allow for more complete
independent comparisons among the proposal, prior EPA analyses, and third-party studies,
ICCT recommends EPA expand its documentation of assumptions and cost breakdowns of
underlying technologies or cost components. For example, EPA could provide average or
exemplary vehicle and component costs across various body styles or classes over time, similar
to tables in prior rulemakings. The general lack of clarity on cost applies to certain combustion
technologies, too, which are identified and discussed in the next section.

COMBUSTION VEHICLE EFFICIENCY POTENTIAL AND COST-


EFFECTIVENESS
While ICCT strongly supports the proposed rule, the cost of compliance may be overstated due
to the use of outdated internal combustion engine vehicle (ICEV) technology data and
information. ICEV technology has been consistently improving for decades. While automakers
are investing heavily in BEV development, the substantial progress that has been—and
continues to be—made in ICE technology has yet to saturate the market. That is, many existing
and recently announced ICEV technology improvements have ample room for increased
application throughout the ICEV fleet. As many ICE vehicles are still to be sold in the MY2027-
2032 timeframe, the proposed rule is an opportunity to maximize their efficiency and minimize
their tailpipe emissions, while providing substantial consumer fuel savings.

ICCT commented extensively on recent ICEV technology improvements in its 2018 comments
on the SAFE NPRM for 2021-26 cars and light trucks (ICCT 2018 comments)82, its study of
LPM and OMEGA modeling of the 2018 Camry (ICCT 2018 Camry)83, its supplemental
comments responding to Toyota comments on ICCT’s study of LPM and OMEGA modeling

80 Wolfe, M. (2022, December). Tula’s DMD promises gains in EV efficiency. SAE Automotive Engineering.
https://www.nxtbook.com/smg/sae/22AE12/index.php#/p/14
81 Madasamy, K. (2023, May 7). Guest commentary: The next frontier of EV innovation: The powertrain. Automotive

News. https://www.autonews.com/guest-commentary/ev-powertrain-next-frontier-innovation
82 ICCT Comments on the Safer Affordable Fuel-Efficient (SAFE) Vehicles Rule for Model Years 2021-2026

Passenger Cars and Light Truck. (2018, October 26). https://theicct.org/news/comments-safe-regulation-2021-


2026 (ICCT 2018 comments)
83 German J. (2018, February 21). How things work: OMEGA modeling case study based on the 2018 Toyota Camry.

https://theicct.org/publications/how-things-work-omega-modeling-case-study-based-2018-toyota-camry (ICCT
2018 Camry)
30
of the 2018 Camry (ICCT 2019 comments)84, and its 2021 comments on the Revised 2023 and
Later Model Year Light-Duty Vehicle Greenhouse Gas Emissions Standards (ICCT 2021
comments).85 Much of the content of these prior comments are reiterated or summarized in the
following subsections, as appropriate and relevant for this proposed rule. Moreover, recent
reports demonstrate that further technology improvements are coming that can boost ICE
vehicle efficiency levels well beyond that of even the highly-efficient Atkinson cycle engine
efficiency levels assumed in this proposal,86,87 as well as show the declining costs of 48-volt mild
hybrid systems.88,89

As documented in the following subsections, the efficiency potential of ICE technology has
continued to improve, while costs have remained lower than previously estimated. Thus, if
technology costs and benefits were updated with the latest information, it would show that the
proposed standards are even more feasible and lower-cost than EPA’s analysis indicates. The
following subsections discuss various ICE technologies and compare the assumptions about
cost and efficiency potential within EPA’s OMEGA analysis with independent research by the
ICCT and other experts.

Gasoline Direct Injection (GDI)


Cost
Based on the DRIA Table 2-30, GDI direct manufacturing costs in EPA’s OMEGA modeling are
between $55-$81 per cylinder. ICCT submitted direct injection cost data in our 2018 comments
based on a 2016 FEV teardown cost study (FEV 2016)90, which found per-cylinder costs to be
about $40 per cylinder. For additional information see:
• ICCT 2021 comments page 6
• ICCT 2018 comments pages I-69–I-70
• FEV 2016

Cylinder deactivation (DEAC)


Application in OMEGA

84 Supplemental Comment from the International Council on Clean Transportation. (2019, April 28). Docket #NHTSA-
2018-0067-12387. https://www.regulations.gov/comment/NHTSA-2018-0067-12387, #NHTSA-2018-0067-12388
https://www.regulations.gov/comment/NHTSA-2018-0067-12388 (ICCT 2019 comments)
85 ICCT comments on the Revised 2023 and Later Model Year Light-Duty Vehicle Greenhouse Gas Emissions

Standards. (2021, September 29). Docket ID EPA-HQ-OAR-2021-0208, https://www.regulations.gov/docket/EPA-


HQ-OAR-2021-0208, Comment ID EPA-HQ-OAR-2021-0208-0522, https://www.regulations.gov/comment/EPA-
HQ-OAR-2021-0208-0522 (ICCT 2021 comments)
86 AVL Webinar on Passenger Car powertrain 4.x – Fuel Consumption, Emissions, and Cost. (2020, June 2).

https://www.avl.com/-/passenger-car-powertrain-4.x-fuel-consumption-emissions-and-cost (Slides available at


https://www.regulations.gov/comment/EPA-HQ-OAR-2021-0208-0522) (AVL 2020)
87 Roush report on Gasoline Engine Technologies for Improved Efficiency (Roush 2021 LDV)

https://www.regulations.gov/comment/EPA-HQ-OAR-2021-0208-0210
88 Roush report on 48V and BEV costs (Roush 2021 48V) https://www.regulations.gov/comment/EPA-HQ-OAR-2021-

0208-0210
89 Dornoff, J., German, J., Deo, A. (ICCT), Dimaratos, A. (DITENCO). (2022). Mild-hybrid vehicles: a near term

technology trend for CO2 emissions reduction. https://theicct.org/publication/mild-hybrid-emissions-jul22/ (ICCT


2022 MHEV)
90 David Blanco-Rodriguez, 2025 Passenger car and light commercial vehicle powertrain technology

analysis. FEV GmbH. (2016, November 21). https://www.theicct.org/publications/2025-passenger-car-and-light-


commercial-vehicle-powertrain-technology-analysis (FEV 2016)
31
Based on the response surface equations (RSE) input file
(simulated_vehicles_rse_ice_20221021_debug_noP2.csv), there are no technology
packages/cost curve classes with DEAC on a turbocharged engine in EPA’s OMEGA modeling.
While adding DEAC to a turbocharged engine has smaller pumping loss reductions than for
naturally aspirated engines, DEAC still has significant pumping loss reductions and has the
additional benefit of enabling the engine to operate in a more thermal efficient region of the
engine fuel map. As described in the National Academies of Sciences’ 2021 report on light-duty
vehicle fuel economy (NAS 2021)91, turbocharged engines with DEAC are already in production
(NAS 2021, section 4.1.3). EPA could consider adding DEAC option to turbocharged cost curve
classes.

Cooled Exhaust Gas Recirculation (CEGR)


Application in OMEGA
Similar to the application of DEAC, based on the RSE input file, there are no cost curve classes
with CEGR on a base turbo engine in EPA’s OMEGA modeling (i.e. “TDS” within the input file).
As reported in NAS 2021 (section 4.1.3) turbocharged engines with CEGR are already in
production. NAS 2021 also provides estimates for the efficiency benefit of including CEGR on a
base turbo engine. EPA could consider adding CEGR option to turbocharged cost curve
classes.

Atkinson cycle engine (ATK)


Application in OMEGA
EPA appears to have excluded the modeling and application of ATK from pickups and other
body-on-frame vehicles (DRIA page 1-10 and RSE input file). However, as detailed in ICCT
2021 comments (pages 14-16), this exclusion could be lifted, allowing all vehicle classes to
adopt ATK in the OMEGA model.

To briefly summarize those ICCT 2021 comments, engines in pickup trucks and high-
performance vehicles are sized and powered to handle higher peak loads. This means larger
engines that operate at lower loads relative to their maximum capacity on the 2-cycle test – and
during most real-world driving. This, in turn, means that pickup trucks and high-performance
vehicles will spend more time in Atkinson Cycle operation than lower performance vehicles on
both the test cycles and in the real world. This includes time spent towing, which represents a
very small fraction of light-duty pickup usage.92,93 Altogether, the large majority of pickup trucks
spend the vast majority of driving at low loads relative to the engine’s capability, where Atkinson
Cycle engines are very effective. In other words, ATK is likely a highly cost-effective technology
for pickup trucks, which may be the most challenging to electrify, as evidenced by the low BEV
share of pickups in the 2027-2031 time frame compared to other body styles in EPA’s modeling
(Preamble Table 80). Furthermore, the claim that an Atkinson Cycle engine that switches to
Otto cycle on demand cannot provide the additional torque reserve is not accurate (a claim

91 National Academies of Sciences, Engineering, and Medicine. (2021). Assessment of Technologies for Improving
Light-Duty Vehicle Fuel Economy—2025-2035. The National Academies Press. https://doi.org/10.17226/26092
92 Berk, B. (2019, March 13). You Don’t Need a Full-Size Pickup Truck, You Need a Cowboy Costume.

Thedrive.com.
https://www.thedrive.com/news/26907/you-dont-need-a-full-size-pickup-truck-you-need-a-cowboy-costume
93 Chase, W., Whalen, J., Muller, J. (2023, January 23). Pickup Trucks: from workhorse to joyride. Axios.

https://www.axios.com/ford-pickup-trucks-history
32
previously used to justify blocking ATK on pickups in prior rulemakings, see ICCT 2021
comments).

Moreover, Atkinson Cycle engines have been used on the Toyota Tacoma pickup V6 engine
since 2017, illustrating that Atkinson Cycle engines are cost-effective for use on pickups.

For additional information see:


• ICCT 2021 comments pages 14-16
• ICCT 2018 comments pages I-2–I-12
• ICCT 2018 Camry study

Miller cycle engine (MIL)


Cost
It is unclear from the DRIA how MIL costs were developed and how OMEGA calculates MIL
costs. Nevertheless, based on analysis of output file engine costs, MIL costs appear too high,
especially as compared to base turbo costs (TDS).

Table 5 below is an excerpt from the central analysis of the proposed standards
(2023_03_14_22_42_30_central_3alts_20230314_Proposal_vehicles.csv). It highlights 9 ICE
vehicles which changed from being equipped with MIL to being equipped with TDS only. Note
that the change from a more advanced engine (MIL) to a less advanced engine (TDS) is an
artefact of the modeling and is unlikely to occur in the real world as such a change may come at
the expense of reduced performance, fuel economy, or other consumer-valued function.
Comparing solely the engine costs associated with a specific vehicle model (assigned to a base
year vehicle ID in the table), the cost of MIL appears to be $600-$2400 more expensive than
TDS. As MIL costs very little compared to a turbocharged, downsized engine such as that used
in the proposal analysis to represent TDS (2016 Honda 1.5L L15B7) (see NAS 2021), the
incremental costs shown in the table suggest the MIL costs are far too high. ICCT 2021
comments, ICCT 2018 comments, and NAS 2021 explain that incremental MIL costs range from
$0–$250.

Table 5. Comparison of Miller (MIL) and Turbo-downsized (TDS) engine costs


Base
Model No. of Displacement engine Cost of MIL
year Cost curve class
year cylinders (L) cost over TDS
veh ID
2026 2 0.6 MIL_TRX22_SS0 $4,657
696 $608
2032 2 0.6 TDS_TRX21_SS0 $4,049
2025 4 1.6 MIL_TRX12_SS1 $9,200
393 $1,854
2031 4 1.6 TDS_TRX22_SS0 $7,346
2025 4 1.6 MIL_TRX22_SS1 $9,160
10 $1,849
2031 4 1.6 TDS_TRX10_SS0 $7,311
2024 4 1.97 MIL_TRX12_SS1 $9,358
540 $2,057
2030 4 1.97 TDS_TRX22_SS0 $7,301
2022 4 2 MIL_TRX22_SS1 $11,808
9 $1,513
2023 4 2 TDS_TRX10_SS0 $10,295
11 2022 4 2 MIL_TRX22_SS1 $11,787 $1,519

33
2023 4 2 TDS_TRX10_SS0 $10,267
2024 4 1.6 MIL_TRX12_SS1 $9,338
539 $1,943
2025 4 1.6 TDS_TRX22_SS0 $7,395
2023 4 2.8 MIL_TRX12_SS0 $11,948
610 $2,408
2024 4 2.8 TDS_TRX21_SS0 $9,540
2023 4 2.8 MIL_TRX12_SS0 $11,949
611 $2,395
2024 4 2.8 TDS_TRX21_SS0 $9,555

Column Column Column Column Column


Column DH Calculation
AF F DV DT AV
Note. The data in each column above was taken from the following columns in
2023_03_14_22_42_30_central_3alts_20230314_Proposal_vehicles.csv

In prior rulemakings, EPA included the cost of ATK in the cost of MIL which led to unnecessarily
high MIL costs. While it is unclear if ATK costs are included in MIL in the current proposal
analysis, such an inclusion would contribute to high MIL costs.

For additional information see:


• ICCT 2021 comments page 7

Mild hybrid (MHEV)


Cost
The costs for position 0 (P0) MHEV are determined according to the equations in DRIA Table 2-
36. In Table 6 below, by comparing these MHEV costs to those found by ICCT in 201694 and
more recently in 202295, ICCT finds the EPA MHEV costs to be 2x-3x more expensive.
Consequently, ICCT recommends EPA reassess and adjust its MHEV costs to better reflect the
most recent data, which are summarized below. One possible source of added cost in the
proposal is the identical calculation of MHEV and HEV costs (scaled by motor power). This
approach may lead to an overestimate of MHEV costs because the 48V electrical systems of
MHEVs do not require the same safety and electrical hardware as higher voltage HEVs. ICCT
recommends adjusting MHEV cost calculations according to the table below.

Table 6. Comparison of EPA and ICCT P0 mild hybrid (MHEV) direct manufacturing costs
P0 MHEV ICCT 2022 ICCT 2016
EPA Table 2-36
cost
(2019 USD) (2020 USD) (2019 USD)
component
battery
0.53 0.53 0.53
(kWh)
power (kW) 15 15 15
Non-battery
component

94 Isenstadt, A., German, J. (ICCT), Dorobantu, M. (Eaton), Boggs, D. (Ricardo), Watson, T. (JCI). (2016).
Downsized, boosted gasoline engines. https://theicct.org/publication/downsized-boosted-gasoline-engines-2/
95 Dornoff, J., German, J., Deo, A. (ICCT), Dimaratos, A. (DITENCO). (2022). Mild-hybrid vehicles: a near term

technology trend for CO2 emissions reduction. https://theicct.org/publication/mild-hybrid-emissions-jul22/ (ICCT


2022 MHEV)
34
Motor +
$362 $300
Inverter
Alternator ** -- -$65
component
DC-DC
$139 costs not $150
converter
estimated
HV cables $190 $0
brakes &
$200 $78
actuators
Non-battery
$891 $323 $462
subtotal
Battery
$880 $314 $449
subtotal
Total $1,771 $637 $911

** Alternator removal costs represented by negative cost (cost savings)


For further information, see Dornoff, J., German, J., Deo, A. (ICCT), Dimaratos, A.
(DITENCO). (2022). Mild-hybrid vehicles: a near term technology trend for CO 2 emissions
reduction. https://theicct.org/publication/mild-hybrid-emissions-jul22/ and
Isenstadt, A., German, J. (ICCT), Dorobantu, M. (Eaton), Boggs, D. (Ricardo), Watson, T.
(JCI). (2016). Downsized, boosted gasoline engines.
https://theicct.org/publication/downsized-boosted-gasoline-engines-2/

Beyond P0 MHEV architectures, there are substantial CO2 reduction benefits achievable by
implementing P1-P4 architectures, representing placement of the motor/generator in positions
of increasing distance from the engine along the driveline. While such systems cost more than
P0, they are more cost-effective in that they have lower cost per percent reduction in CO2. Thus,
ICCT recommends EPA consider including in its modeling more advanced MHEV architectures
beyond P0. Additional discussion on MHEV effectiveness is in the following section. Table 7
below replicates Table 18 in ICCT 2022 MHEV. As shown in the table, P1-P4 MHEV
architectures with specifications similar to P0 MHEV can increase cost by at most 53% (P4+P0
for FWD) with P4+P0 for AWD decreasing costs vs P0. At the same time, P2-P4 architectures
can more than double P0 effectiveness. Combining the “Total” cost scaling shown in the below
table with the ICCT P0 cost in the table above, all architectures have lower cost than the P0
MHEV cost used in the proposal.

Table 7. Mild hybrid architecture cost in 2020 (ICCT 2022 MHEV, Table 18)
System
Cost normalized to P0 Effectiveness
specifications
Architecture
Non- WLTP CO2 Cost per % CO2 reduction
Motor Battery Battery Total
battery reduction normalized to P0
P0 16 kW 800 Wh 1.00 1.00 1.00 6.6% 1.00
P1 15 kW 800 Wh 1.00 1.29 1.15 8.5% 0.89
P2 side
16 kW 800 Wh 1.00 1.51 1.27 11.9% 0.70
mounted
P2 coaxial 15 kW 800 Wh 1.00 1.62 1.32 14.8% 0.59
P3 16 kW 800 Wh 1.00 1.65 1.34 15.3% 0.58
P4+P0 vs. 15
800 Wh 1.00 2.01 1.53 15.5% 0.65
FWD kW+4kW

35
P4+P0 vs. 15
800 Wh 1.00 0.82 0.91 23.9% 0.25
AWD kW+4kW
For further information, see Dornoff, J., German, J., Deo, A. (ICCT), Dimaratos, A. (DITENCO). (2022). Mild-hybrid vehicles: a near term
technology trend for CO2 emissions reduction. https://theicct.org/publication/mild-hybrid-emissions-jul22/

Effectiveness
Outlined in the table above, MHEV architectures beyond P0 can have substantial CO2-reduction
benefits. However, the benefits of more advanced MHEV architectures are expected to exceed
those illustrated in the table, through the implementation of higher power systems (20kW-
30kW). Roush 2021 LDV96 describes the additional benefits offered by higher power MHEV
systems, including advancements in electric boosting, high energy ignition systems (see section
below), accessory electrification, and electrically heated catalysts. Enabling electrically heated
catalysts in particular permits further fuel economy optimization through, for example,
aggressive stop-start strategies.

For additional information, see:


• Roush 2021 LDV page 11 and pages 38-40
• Roush 2021 48V97 pages 11-23
• AVL 2020 slide 6298

Roush 2021 LDV provides specific example applications of high power MHEV systems and the
associated fuel efficiency improvements on pickups and SUVs. These examples, which have
not previously been considered by either EPA or ICCT, are excerpted below:

Pickup/full-size SUV GHG reduction: As ICCT previously commented, Roush 2021 states “Two
powertrain configurations are recommended for study and could support future rulemaking. The
first option synergistically combines available technologies (without a major redesign of the
underlying engine architecture) to give maximum fuel economy benefit for a relatively low cost,
hence high effectiveness. It combines a naturally aspirated DI engine with advanced cylinder
deactivation and a 30kW 48V P2 mild hybrid system. The 48V hybrid system is used to actively
smooth out crankshaft torque pulsations to enable aggressive cylinder deactivation strategies
(advanced deac – like the Tula Skipfire System). Such a system will also enable start-stop,
electric creep, regen braking, slow-speed electric driving, and a heated catalyst. Depending on
system integration factors Roush estimates a reduction in GHG emissions of 20% or more,
compared to a baseline naturally aspirated direct-injection V8.” (Roush 2021 LDV page 13).
Additional information can be found at:
• Roush 2021 LDV Section 13.1 page 65

Compact SUV GHG Reduction: Relatedly, as ICCT previously commented, Roush 2021 states,
“A 30kW 48-volt P2 system mated to a low bore-to-stroke ratio Miller cycle engine with
electrified boosting, advanced cylinder deactivation, cooled EGR and a heated catalyst can
provide a fuel economy benefit close to a full high voltage hybrid powertrain at a much lower
cost. The 48V electric motor can supplement the engine torque under low- speed high load

96 Roush report on Gasoline Engine Technologies for Improved Efficiency (Roush 2021 LDV)
https://www.regulations.gov/comment/EPA-HQ-OAR-2021-0208-0210
97 Roush report on 48V and BEV costs (Roush 2021 48V) https://www.regulations.gov/comment/EPA-HQ-OAR-2021-

0208-0210
98 AVL Webinar on Passenger Car powertrain 4.x – Fuel Consumption, Emissions, and Cost. (2020, June 2).

https://www.avl.com/-/passenger-car-powertrain-4.x-fuel-consumption-emissions-and-cost (Slides available at


https://www.regulations.gov/comment/EPA-HQ-OAR-2021-0208-0522) (AVL 2020)
36
conditions, thereby avoiding this knock-prone area of the engine map. Also, the use of an
advanced boosting system, combining a turbocharger and a 48V electric supercharger, will
reduce engine backpressure (larger turbine) and improve scavenging, reduce combustion
residuals, and reduce the propensity for knock. This combination enables the use of a higher
compression ratio, thereby increasing engine efficiency. A combination of a high-energy ignition
system (high energy spark plug/plasma ignition) and fuel reforming by pilot fuel injection during
NVO can be used to increase cEGR tolerance at low loads. The initial part of such a project
would include engine and combustion modeling, followed by prototype engine testing. The
overall GHG reduction potential will require modeling and optimization of engine design,
calibration parameters, and boosting system sizing and control. Roush estimates a reduction in
GHG emissions exceeding 30% is possible compared to a level 1 (NHTSA) turbocharged
engine.” (Roush 2021 LDV page 14).
Additional information can be found at:
• Roush 2021 LDV Section 2.3 pages 23-25 on higher compression ratios and higher
Miller/Atkinson ratios.
• Roush 2021 LDV Sections 2.4 and 2.5 pages 26-28 on low bore-to-stroke ratio
benefits
• Roush 2021 LDV Section 13.2 page 66

Strong hybrid (HEV)


Cost
EPA’s estimated HEV costs also appear to be overestimated. Due to the challenge of
disentangling the costs and effects of both electrified and conventional powertrain component
changes from one redesign to the next, ICCT examined total powertrain costs from the central
analysis output file (2023_03_14_22_42_30_central_3alts_20230314_Proposal_vehicles.csv).
Total powertrain costs are calculated as the sum of battery cost, electrified driveline cost, e-
machine cost, driveline cost, and engine cost. For redesigns that occur between MY2023-2027,
the difference in cost between the HEV powertrain and non-HEV powertrain ranges from
approximately $3,400 to over $9,000. For redesigns that occur after MY2027, the HEV
powertrain cost difference ranges between approximately $2,700 to over $10,000. Cost
changes due to learning notwithstanding, as analyzed in ICCT’s 2015 report on hybrids,99 these
levels of cost premiums are not plausible.

As with MIL costs, it is not clear precisely how HEV engine (MIL, DHE, and DHE2) costs are
calculated. Regardless, as HEV engines are modeled as either Atkinson or Miller cycle engines,
their costs ought to be very similar to the same types of engines on non-HEV models. As ATK
and MIL are fairly inexpensive as compared to a sufficiently advanced engine (NAS 2021),
modeled HEV engine costs ought not to be significantly more expensive than their non-HEV
counterparts. In fact, due to the capacity of HEV motor to take up low-speed, high torque
demand and transient response, HEV engines can be optimized to a narrower operating range
than non-HEV engines. This can enable higher compression ratios, increase EGR dilution, and
potentially decrease costs. Especially in the case of a serial hybrid or range extended PHEV,
the engine is effectively decoupled from the drivetrain, permitting deep optimization, with up to
40% engine cost reduction depending on electrification.100

99 German, J. (2015). Hybrid vehicles: Trends in technology development and cost reduction. International Council on
Clean Transportation.
https://www.theicct.org/hybrid-vehicles-trends-technology-development-and-cost-reduction
100 SAE 2021. (2021). Optimizing hybrids for cost and efficiency. SAE Automotive Engineering. Page 18.

https://www.nxtbook.com/smg/sae/21AE04/index.php#/p/18
37
Effectiveness
ICCT commends EPA’s incorporation of advanced Atkinson and Miller cycle engines. However,
the notion of a dedicated hybrid engine (DHE) extends beyond the engine maps used during
ALPHA HEV simulation. ICCT recommends EPA consider even further optimized/efficient
dedicated hybrid engines, both for HEV applications and for PHEVs. As described in ICCT’s
2021 comments and in SAE (2021),101 “EPA should focus on the expanded application of
energy management capabilities in full hybrid powertrains to also minimize operation under the
low-speed high torque areas of the engine which are prone to knocking by torque augmentation
with the electric motor. The instantaneous torque capability of the electric motor can effectively
support transient torque demand. This will allow both naturally aspirated and turbocharged
engines that are part of a hybrid powertrain to be optimized for a narrow operating range
incorporating higher compression ratios and increased EGR dilution (maintaining stoichiometric
operation), thereby prioritizing efficiency over peak torque at low engine speeds and transient
response.” (Roush 2021 LDV page 12)
For additional information, see:
• Roush 2021 LDV Section 7.0 pages 41-44
• AVL 2020 slide 24: BSFC for Lambda=1
• AVL 2020 slides 25-26: Dedicated Hybrid Engine Efficiency Roadmaps (45%
Lambda=1, 51% ideal)
• AVL 2020 slides 35-42: WLTP CO2 reduction potential of various hybrid
configurations
• AVL 2020 slide 43: Relative comparison of attributes for three powertrain
architectures
• AVL 2020 slide 62: WLTP % CO2 reduction and slide 63: cost per % FC reduction

Such dedicated hybrid engines can achieve 45% brake thermal efficiency (BTE) at
stoichiometric air-fuel ratio using known technologies,102 or 50% BTE in a serial/range-extender
with pre-chamber ignition, ultra-high pressure injection, and reduced intake air temperatures
(SAE 2021).

Negative valve overlap in-cylinder fuel reforming (NVO)


Effectiveness
As ICCT previously commented, Roush states, “In-cylinder fuel reforming by using pilot fuel
injection during NVO has shown to significantly improve cooled EGR (cEGR) tolerance,
combustion stability, and engine efficiency. Such a system can have wide application in
turbocharged and NA engines across different vehicle segments with minimal hardware
requirements. Depending on the base engine, Roush estimates an efficiency improvement, and
the corresponding reduction in GHG emissions, in the range of 5 to 10% is possible and low
cost, therefore correspondingly high effectiveness.” (Roush 2021 LDV page 14).
Additional information can be found at:
• Roush 2021 LDV Section 10.0 pages 50-52
• Roush 2021 LDV Section 13.3 page 66

101 Ibid.
102 Visnic, B. (2022, April). Keeping combustion in the conversation. SAE Automotive Engineering. Page 18.
https://www.nxtbook.com/smg/sae/22AE04/index.php#/p/18
38
Passive prechamber combustion (PPC)
Effectiveness
As ICCT previously commented, Roush states, “Prechamber combustion systems are one of
the most promising technologies for improving the dilution limit of engines, thereby improving
system efficiency. It can also enable extremely fast burn rates increasing the knock tolerance of
turbocharged engines, allowing higher compression ratios and the associated efficiency
improvements. The Maserati Nettuno engine in the 2021 Maserati MC20 will be the first
application of a passive prechamber engine in production. However, the primary objective in the
MC20 is high performance. It would be very valuable to study the effect of the system on knock
tolerance, burn rates, dilution tolerance (EGR and air), and emissions. The effort should focus
on quantifying possible efficiency gains in a non-performance application.” (Roush 2021 LDV
pages 14-15). In a dedicated hybrid engine developed by Mahle, pre-chamber combustion
enabled a CO2 emissions reduction of over 5%.103
Additional information can be found at:
• Roush 2021 LDV Section 13.4 page 67
• AVL 2020 slides 28, 31, and 33

High energy ignition (HEI)


Effectiveness
As ICCT previously commented, Roush states, “High energy volume ignition systems can
enable combustion of dilute (cEGR or air diluted) in-cylinder mixtures resulting in a step-change
in engine efficiency compared to conventional spark plugs. Such systems can be a drop-in
replacement for a spark plug, thereby representing a cost-effective GHG improvement option.
Such systems should be evaluated for maximum efficiency potential, in conventional, 48V mild
hybrid, and full HV hybrid applications. Roush estimates that systems such as plasma ignition
can support good combustion stability with high amounts of cooled EGR, thereby achieving
engine efficiency improvements in the range of 5-10% over a baseline turbocharged DI, dual
VVT engine. Microwave ignition systems, on the other hand, have the potential to achieve levels
consistent with prechamber ignition systems. This would enable lean-burn engines with low
engine-out NOx emissions which can achieve brake thermal efficiency which exceeds 45% in
light-duty vehicle applications, compared to a level of 36-38% for a baseline turbocharged DI,
dual VVT engine.” (Roush 2021 LDV page 15). Additional information can be found at:
• Roush 2021 LDV Section 11.0 pages 53-62
• Roush 2021 LDV Section 13.5 page 67

Transmissions
Application in OMEGA
In this proposal, EPA did not consider the application of automatic transmissions with 9 or more
gear ratios. This is unrealistic, as nearly 40% of pickups were equipped with such transmissions
in 2021, and EPA expects more than a quarter of all vehicles to have such transmissions in
MY2022.104 Without including the costs and benefits of transmissions with additional gears, EPA
is missing important fuel-savings technology. Consequently, ICCT recommends EPA
incorporate in its analysis transmissions with 9 or more gears.

103 Birch, S. (2019, November). Mahle reveals modular, scalable integrated hybrid powertrain. SAE Automotive
Engineering. Page 14. https://www.nxtbook.com/nxtbooks/sae/19AUTP11/index.php#/p/14
104 EPA. (2022). Automotive Trends Report [detailed automotive trends data]. https://www.epa.gov/automotive-

trends/explore-automotive-trends-data
39
Lightweighting
Cost
In the proposal, the only lightweighting option is the switch to an aluminum body. This is
certainly a viable lightweighting option, but it is not the only one. Manufacturers have many
avenues for lightweighting with various degrees of mass reduction and associated cost (NAS
2021 and ICCT 2018 comments).

EPA should consider adjusting its lightweighting options in its analysis to incorporate varying
levels of mass reduction at vary levels of cost per unit mass saved. This methodology has been
used in prior rulemakings and can fit into EPA’s existing mass and cost calculations.
Alternatively, EPA can add a single, discrete, intermediate lightweighting option (between the
base steel body and lightweighted aluminum body) that mimics the same format as the two
existing options. This intermediate option would be composed of primarily ultra-, advanced- and
high strength steels (as opposed to conventional or mild steel). Such steels with optimized
design can offer mass reductions on the order of 10%-15% (higher for specific parts), at costs
comparable to existing steel costs.105 More recent steel developments indicate further mass
reductions are possible with both steel and better design optimization, with high strength steel
costs similar to mild steel costs up to half the cost of lightweighting with Aluminum.106

Potential technology penetration impacts of improved ICE technology adoption


With the above recommended adjustments to ICE technology costs and effectiveness, potential
implementation of such technology on the MY2027-2032 ICEV fleet can result in significant
reduction in tailpipe CO2 emissions from ICE vehicles. In its study of updated light-duty vehicle
costs,107 ICCT assumed updated ICE technology adoption as described above would continue

105 Isenstadt, A. and German, J. (ICCT); Piyush Bubna and Marc Wiseman (Ricardo Strategic Consulting);
Umamaheswaran Venkatakrishnan and Lenar Abbasov (SABIC); Pedro Guillen and Nick Moroz (Detroit
Materials); Doug Richman (Aluminum Association), Greg Kolwich (FEV). Lightweighting technology development
and trends in U.S. passenger vehicles, (2016, December 19). http://www.theicct.org/lightweighting-technology-
development-and-trends-us-passenger-vehicles
106 Brooke, L. (2019, May). The economics of materials selection. SAE Automotive Engineering.

https://www.sae.org/publications/magazines/automotive-engineering/past-issues; Gehm, R. (2019, September).


Latest mass-reducing innovations honored by Altair. SAE Automotive Engineering.
https://www.sae.org/publications/magazines/automotive-engineering/past-issues; Visnic, B., Brooke, L. (2019,
September). Stuck on structural adhesives. SAE Automotive Engineering.
https://www.sae.org/publications/magazines/automotive-engineering/past-issues; Weissler, P. (2019, October).
Cutting weight seen as less vital for automated and shared vehicles. SAE Automotive Engineering.
https://www.sae.org/publications/magazines/automotive-engineering/past-issues; Gehm, R. (2020, September).
Altair honors lightweight advances. SAE Automotive Engineering.
https://www.sae.org/publications/magazines/automotive-engineering/past-issues; Macek. B. (FCA), Lutz, J. (US
Steel). (2020, September). Virtual and physical testing of Third-generation High Strength Steel. SAE Automotive
Engineering. https://www.sae.org/publications/magazines/automotive-engineering/past-issues; Vartanov, G.
(2021, June). Lightweight steel on a (cold) roll. SAE Automotive Engineering.
https://www.sae.org/publications/magazines/automotive-engineering/past-issues; Gehm, R. (2021, September).
Altair honors weight-saving innovations. SAE Automotive Engineering.
https://www.sae.org/publications/magazines/automotive-engineering/past-issues; Brooke, L. (2022, September). A
materials lesson in Civics. SAE Automotive Engineering. https://www.sae.org/publications/magazines/automotive-
engineering/past-issues; Gehm, R. (2022, September). Altair honors innovations in sustainability and
lightweighting. SAE Automotive Engineering. https://www.sae.org/publications/magazines/automotive-
engineering/past-issues; Brooke, L. (2023, March). Battle for the box. SAE Automotive Engineering.
https://www.sae.org/publications/magazines/automotive-engineering/past-issues
107 Slowik. P., Isenstadt, A., Pierce, L, Searle, S. (2022). Assessment of light-duty electric vehicle costs and

consumer benefits in the United States in the 2022-2035 time frame. International Council on Clean
Transportation. https://theicct.org/publication/ev-cost-benefits-2035-oct22/
40
to provide ICEV improvement equivalent to 3.5% per year through at least 2032. Fuel savings
with that level of improvement more than offset incremental technology cost. Starting from a
2022 fuel economy value of 27 mpg (328 g/mi), the overall real-world efficiency of the
combustion light-duty vehicle fleet was 32 mpg (275 g/mi) in 2027 and 36 mpg (247 g/mi) in
2030 and 39 mpg (230 g/mi) in 2032. In total, over MY2027-2032, the net reduction in ICE
tailpipe emissions is 16.4%. Applying that same annual improvement to the industry average car
and truck ICE fleets as modeled for this proposed rule results in a total 16% reduction in ICE
tailpipe emissions, reducing the reliance on BEVs as a compliance pathway. Specifically, we
find that improvements in ICE efficiencies would reduce the projected BEV shares from 60% in
2030 to about 54%, and from 67% in 2032 to about 59%. In other words, the same GHG levels
as proposed can be reached through improved ICE and fewer BEVs.

Table 8 summarizes these and additional results under several ICE improvement pathways.
The ICE fleet, as modeled in the proposal, shows a net increase in 2-cycle tailpipe emissions in
MY2032 versus MY2027. The share of advanced ICE technology, especially mild and strong
hybridization, changes little or even decreases within the ICE fleet. Updating the OMEGA model
to reflect the full technology options and lower costs we detail above would lead to higher
tailpipe GHG reductions from ICE vehicles and, consequently, lower shares of BEVs needed for
any particular GHG standards scenario. As just one example, considering the low share of
hybrids in the ICE fleet as modeled, there is considerable room for powertrain electrification.
With the MHEV and HEV developments discussed above, deeper application of hybridization
could yield total ICE improvements of 15%-30%, easily matching 3.5%/year or even 5%/year.
The MY2032 ICE vehicles in these scenarios would emit 16%-22% lower CO2 emissions than
the ICE vehicles modeled in the proposal. The resulting estimated BEV shares in 2032 could be
12%-20% lower than proposed. The availability of hybrid and other ICE technologies explored
above further strengthens the case that automakers have a variety of technological pathways to
pursue to meet the proposed standards and realize substantial ICE improvements. The potential
impact on the future fleet composition in Table 8 is illustrative of this variety of pathways
automakers can choose to meet the proposed targets. Incorporation of the above ICE
technology changes into OMEGA and subsequent modeling runs are needed for precise
estimates of projected impacts of the above-listed technology updates.

Table 8. Illustrative compliance pathways to meet MY2027-2032 proposed standards


Proposal Constant MY2027 1%/year 3.5%/year 5.0%/year
Assumed ICE
annual emissions -0.8% ** 0%/year 1%/year 3.5%/year 5.0%/year
reductions
Average 2-cycle
255 246 234 206 191
ICE g/mi by 2032
Net MY27-32 ICE
-4.1% ** -0.4% ** 4.5% 15.9% 22.0%
improvement *
Projected MY2032
67% 66% 64% 59% 55%
BEV share
* Net improvement is fleet average, weighted by the share of ICE cars and ICE trucks.
** Negative values indicate the ICE fleet is, on average, emitting more in MY2032 than in MY2027

As explained above and shown in the table, the ICE fleet emits more in 2032 than in 2027. In
fact, based on analysis of OMEGA results, the car fleet has higher 2-cycle tailpipe emissions in
MY2032 than in MY2022, with its lowest 2-cycle emissions levels occurring in MY2027.108
Similarly, the truck fleet shows its lowest GHG emissions levels in MY2028, with a net increase
in 2-cycle emissions by MY2032. Much of this backsliding is explained by individual ICE models

108 Based on analysis of 2023_03_14_22_42_30_central_3alts_20230314_Proposal_vehicles.csv


41
getting worse over time. Of all the 719 ICE models, 438 show higher emissions in MY2032 than
in MY2027, and 333 show worse emissions in MY2032 than in MY2022. This modeling result
presented in EPA’s proposal highlights the risk that backsliding on GHG emissions could occur
in real-world compliance with EPA’s standards. If BEV penetration occurs faster than modeled
by EPA, there is risk of further ICE backsliding on GHG emissions. ICCT recommends EPA
include in its regulation a mechanism to prevent the backsliding of new ICE vehicles or the new
ICE fleet.

OFF-CYCLE CREDITS
ICCT strongly supports EPA’s proposal to sunset or eliminate air-conditioning (AC) and off-cycle
(OC) credits.

Eliminating the AC leakage credit is reasonable, since at least 95% of new LD vehicles now use
HFO-1234yf,109 which is a low-GWP refrigerant. Since virtually all vehicles have this low-GWP
refrigerant, this low-GWP status is now the baseline and there is no need to grant further AC
leakage credits. Regarding AC efficiency credits, ICCT supports EPA limiting these credits to
ICEVs, as the credits are based on ICE tailpipe emissions reductions for AC-system
improvements. While BEVs also benefit from improved AC system efficiency, BEVs do not
require the additional incentive provided by AC credits. BEVs already are granted compliance
emissions values of 0 g/mi, and advances in AC efficiency are inherent to BEV development, as
passenger and battery heating/cooling loads can significantly impact BEV range and battery
size requirements.

ICCT supports the proposal to phaseout OC credits by MY2031, to eliminate the off-menu OC
credit option starting MY2027, and to limit OC credits to ICE vehicles. As with AC credits, a
large portion of the fleet already incorporates the technologies that are granted OC credits.
According to the 2022 Automotive Trends Report data, MY2021 cars averaged 5.1 g/mi in OC
credits (51% of the 10 g/mi cap) and trucks averaged 10.2 g/mi in OC credits (102% of the 10
g/mi cap).110 With these averages as a proxy for the share of the car and truck fleets with OC
technology, this technology is already widespread in the baseline and requires no further
incentivization. Evidence suggests that the menu OC credit values (such as solar and thermal
load control) overestimate the real-world impact of OC technologies.111 Moreover, the menu
credits are defined in terms of absolute g/mi reductions, rather than relative or percentage-
based reductions, as virtually all on-cycle technologies are defined. Because of this
inappropriate definition, as vehicles become increasingly efficient, these absolute credit values
represent unrealistically large shares of vehicles’ overall emissions improvement. Additionally,
as OC credits are based on reduced tailpipe emissions from ICE vehicles, they are not
applicable to BEVs. As with AC efficiency improvements, any innovation that reduces real world
energy consumption in BEVs is inherently incentivized by the reduced battery capacity
requirements of incorporating such innovations. Relatedly, ICCT supports the proposal to scale
PHEV OC credits by the (newly proposed) utility factor.

109 EPA. (2022). Automotive Trends Report. https://www.epa.gov/automotive-trends/download-automotive-trends-


report#Full%20Report
110 EPA. (2022). Automotive Trends Report [data]. https://www.epa.gov/automotive-trends/explore-automotive-trends-

data
111 Lutsey, N., Isenstadt, A. (2018). How will off-cycle credits impact U.S. 2025 efficiency standards? International

Council on Clean Transportation. https://theicct.org/publication/how-will-off-cycle-credits-impact-u-s-2025-


efficiency-standards/
42
ICCT recommends EPA adjust its modeling of OC credits to more accurately reflect the trend of
the car and truck fleets taking full advantage of the OC credit cap. As currently modeled, both
the car and truck fleets use the full value of their respective AC efficiency credit cap.112 The OC
credits, however, are not modeled as reaching their cap for either car or truck fleet in any model
year. As evidenced in the 2022 Automotive Trends Report data discussed above, the car fleet is
currently receiving at least 50% of the full OC credit cap, while the truck fleet is receiving at least
100% of the full OC credit cap. ICCT suggests EPA modify its inputs to OMEGA to capture the
real-world trend of manufacturers taking full advantage of the off-cycle credit cap. (The AC
efficiency credits are appropriately modeled at the full value of the cap, which accurately
represents the trend of the fleet receiving nearly 90% of the AC credit cap in MY2021.) If EPA
incorporates the maximal use of OC credits during years that the OC cap is nonzero, then the
standards could be made more stringent to offset this higher OC credit usage and maintain the
same GHG emissions impacts.

CRITERIA POLLUTANT EMISSIONS

NMOG and NOx emissions


ICCT supports EPA’s inclusion of the new provisions on vehicles’ operating conditions for light-
duty vehicles that align with the CARB ACC II program. The additional operating conditions
include high power cold starts in plug-in hybrid vehicles, SFTP early drive-away, and SFTP
intermediate soak mid-temperature starts (called partial soak standard in ACC II). The bin-
specific standards are mostly congruent with those in the ACC II program with some slight
difference. Instead of having Bin 125, 25 or 15, the proposal includes Bin 10 which has more
stringent NMOG + NOx standards compared to CARB’s.

ICCT supports EPA’s move to set more stringent NMOG + NOx emission limit for light-duty
vehicle bins and the fleet average. We have observed that these proposed standards are more
stringent than the Euro 7 emission regulations for cars and vans. The emission limit for all cars
and vans in Euro 7, converted from mg/km, is 206 mg/mile. Meanwhile, in EPA’s proposal, Tier
4 emission limit for all LDVs is 70 mg/mile. More details on the difference in emission limits
between the two standards are show in Figure 11. Even if the SFTP early driveaway standard
has a higher NMOG + NOx emission limit at 82 mg/mile, it is more stringent than the Euro 7
emission limit at 206 mg/mile.

For NMOG + NOx fleet average standards, EPA provides two pathways for automakers: default
and early compliance. The default compliance pathway provides a full 4 years of lead time and
3 years of standards stability for LDT3, LDT4, MDPV, MDV class 2b and 3 which are defined as
heavy-duty vehicles by the Clean Air Act.113 However, automakers can opt for the early
compliance pathway, in which LDT3, LDT3, MDPV meet identical and declining fleet averages
like LDV, LDT1, and LDT2. MDV class 2b and 3 also has gradually declining standards every
model year. ICCT supports the option of the early compliance pathway that will result in higher

112 Based on analysis of column BV of output file


2023_03_14_22_42_30_central_3alts_20230314_Proposal_vehicles.csv as well as inspection of input file
offcycle_credits_20230206.csv
113 Emission standards for new motor vehicles or new motor vehicle engines, 42 U.S. Code § 7521,

https://www.law.cornell.edu/uscode/text/42/7521
43
reduction in NMOG + NOx emissions and greater environmental and health benefits. ICCT
supports EPA encouraging automakers to adopt the early compliance pathway.

EPA could consider even lower NMOG + NOx fleet average standards for light-duty vehicles.
EPA’s standards are not as stringent as CARB’s. CARB LEV IV phases out ZEV and PHEV in
fleet average calculation, at 100% in MY 2025, 60% in MY 2026, 30% in MY2027, 15% in
MY2028, to 0% by MY 2029. From MY2029+, the ICE vehicle fleet average is 30 mg/mile. The
NMOG + NOx fleet average emissions under Tier 4 in MY 2029 would be higher than 30
mg/mile. For example, the EPA projects BEV penetration rate of light-duty vehicle fleet at 55%
in 2029. Assuming the early compliance path, NMOG + NOx fleet average emission of ICE
vehicles in 2029 could be 40 mg/mile (18 mg/mile MY2029 standards divided by the ICE vehicle
penetration of 45%). Similarly, in 2032, the ICE fleet average emission could be 36 mg/mile (12
mg/mile divided by 33% ICE fleet penetration), still higher than CARB. EPA does not include a
PHEV projection in the proposal. If PHEVs that still have tail-pipe emissions are included and
the BEV penetration rate is lower than projected, the ICE vehicle NMOG + NOx fleet average
could be higher than these calculated values.

However, for MDVs, the ICE fleet average emission projection could be lower than CARB’s LEV
IV standards for later MY 2030+. EPA projects BEV penetration rate at 46% in MY 2032.
Assuming the early compliance path, NMOG + NOx fleet average emission of ICE MDV
vehicles in MY 2032 could be 111 mg/mile (60 mg/mile MY 2032 standard divided by the ICE
vehicle penetration of 54%), lower than CARB’s fleet average of 150 mg/mile for class 2b and
175 mg/mile for class 3.

International comparison on NMOG + NOx emission standards


Figure 11 compares the light-duty vehicle NMOG + NOx emission standards between three
authorities: European Union (Europe), CARB, and U.S. EPA. All purple markers represent
standards up until MY 2025, and green markers represent MY 2025+ standards. The yellow line
represents the Euro emission limits with purple circles illustrating Euro 6 standards that regulate
emissions from MY2014 to 2025 cars and vans and green circles for Euro 7. Euro 6 had three
separate standards for 3 categories of vehicle classes, based on vehicle weights. 114 The blue
line represents CARB LEV standards. LEV III had two separate standards, one for Passenger
cars and Light-duty truck class 1 (LDT1), and one for LDT 2,3,4 and MDPV.115 These LEV III
standards are illustrated by the purple triangles. Green triangles show LEV IV standards that are
the same for all LDV vehicle classes. The red line shows the U.S. EPA Tier standards. The
purple squares illustrate Tier 3 standards and the green squares Tier 4 standards. The figure
also shows that Tier 3 and LEV III share the same emission limit at 0.16 g/mile or 160mg/mile
and is lower than Euro 6 limits for all vehicle categories. Similarly, Tier 4 and LEV IV limits are
lower than Euro 7 limits, however, Tier 4 emission limit at 70 mg/mile is lower than LEV IV limit
at 125 mg/mile.

114 M1 are vehicles that carry passengers and comprise of no more than 8 seats in addition to the driver seats, M2
are vehicles similar to M1 and have a maximum mass not exceeding 5 tons. N1 vehicles are constructed to carry
goods with maximum mass not exceeding 1305 kg (class I), 1760 kg (class Ii) and 3,500 kg (class III)
115 LDV are passenger cars with 8,500 lbs GVWr maximum, LDT1 vehicles are trucks with 3,750 lbs LVW maximum,

LDT2 vehicles are trucks with 3,750 lbs LVW minimum, LDT3 vehicles are trucks with more than 6,000 lbs GVWR
and with an ALVW of 5,757 lbs or less, LDT4 vehicles are trucks with more than 6,000 lbs GVWR and with an
ALVW of more than 5,750 lbs. MDPVs are any passenger vehicles at or below 14,000 lbs GVWR with a work
factor at or below 5,000 lbs. MDPV also includes pickups with GVWR below 9,900 lbs with an interior bed lengths
less than 8 feet regardless of its work factor above 5,000 lbs. Pickups with GVWR between 9,900 and 14,000 lbs
with a work factor above 5,000 lbs can be included if the interior bed length is less than 6 feet.
44
Figure 11. Light-duty vehicle NMOG + NOx emission standard comparison between CARB,
EPA, and Euro standards
Notes:
1. EU Euro standards are max-per-vehicle or vehicle limit values that apply both in laboratory testing and real-world emission, while CARB
LEV and EPA Tier programs have both limit values and fleet average values that are only laboratory test.
2. CARB LEV III, LEV IV and EPA Tier 3, Tier 4 set NMOG + NOx standards; EU Euro standards are the summation of independent NOx
standards and NMHC/NMOG standards.
3. Starting MY 2029, LEV IV phase out ZEV+PHEV in the fleet average calculation, while Tier 4 still includes them.
4. Tier 4 LDT3,4 and MDPV standards are based on the early compliance pathway that is similar to Tier 4 PC + LDT1,2

As mentioned in the previous section, Tier 4 standards include EV emissions in the NMOG +
NOx fleet average calculation, while LEV IV standards gradually phase out EVs in the
calculation by MY2029. As a result, even though the figure shows Tier 4 with a lower fleet
average value than LEV IV, the projected ICE vehicle fleet average under Tier 4, illustrated as
the dash red line, could be higher than LEV IV in MY 2029-2032.

Figure 12 compares the medium-duty vehicle NMOG + NOx emission standards between two
authorities: CARB and U.S. EPA. All purple markers represent standards up until MY 2025, and
green markers represent MY 2025+ standards. The blue lines represent CARB LEV standards,
and the red lines represent EPA Tier standards. Both Tier 3 and LEV III had the same emission
limit and fleet average for Class 2b and 3. However, starting with MY 2026, the Tier 4 emission
limit is the same for both Class 2b and 3 at 160 mg/mile and lower than LEV IV at 250 mg/mile
for Class 2b and 400 mg/mile for Class 3. Tier 4 MDV standards include EV emissions in the
NMOG + NOx fleet average calculation, while LEV IV standards do not. The figure also shows
that for MY 2030+, the projected ICE fleet average emissions shown as the dash red line could
be lower than for the LEV IV Class 2b standards.

45
Figure 12. Medium-duty vehicle NMOG + NOx emission standard comparison between CARB
and EPA

Evaporative emission
ICCT recommends EPA to set stronger running loss standards. The Tier 4 proposal allows
vehicles to continue to be subject to Tier 3 evaporative emission standards, which states
running loss may not exceed 0.05 g/mile. The running loss is reduced from 0.05 in LEV III to
0.01g/mile in LEV IV. According to CARB’s Initial statement of reasons, since 1990, when the
LEV emission standards were established, the 0.05 g/mile has not changed. The 2019 models’
certification data showed that 87% of vehicles were already emitting at or below 0.01 g/mile.
Thus, it is feasible for EPA to set more stringent running loss standards in alignment with
CARB.116

Particulate matter emission


Recent trends in gasoline light-duty PM emissions

ICCT strongly support’s EPA’s proposed PM limit. Fine particulate matter (PM2.5) emissions
from vehicles are a major environmental health hazard, and setting a more stringent PM
emissions limit is critical for protecting public health. EPA’s proposed multipollutant rule sets the
PM limit at 0.5 mg/mile on all test cycles with the addition of the cold temperature FTP test
cycle, which would help to greatly reduce tailpipe PM emissions from new gasoline vehicles. In
this section, we review recent evidence of increases in PM emissions from recent model year
gasoline light-duty vehicles and trucks, highlighting the urgent need to implement a stronger PM
standard.

116 CARB. (2022). Initial Statement of Reasons. https://ww2.arb.ca.gov/sites/default/


files/barcu/regact/2022/accii/isor.pdf and https://ww2.arb.ca.gov/sites/default/ files/barcu/regact/2022/accii/isor.pdf
46
A recent ICCT analysis looking at remote sensing data shows that, for recent model years, light-
duty vehicles and trucks show an increase in UV smoke, a proxy for PM. 117 While gasoline light-
duty vehicles’ and trucks’ tailpipe carbon monoxide (CO), hydrocarbons (HC), and nitrogen
oxides (NOx) measured by remote sensing show clear and consistent downwards trends, UV
smoke averages increase starting from model years 2015–2020 (Figure 13). For 2020 model
year vehicles, the UV smoke fleet average is similar to that of 2005 model year vehicles,
compared to the other three pollutants, which show a 66%–86% decrease from 2005 to 2020
model year vehicles.
Percent chang e in d ist ance−sp ecific em issio n rat e since MY 20 0 5

0%

−25%
Po llut ant
CO
HC
NOx
−50 % UV sm oke (PM)

−75%

20 0 5 20 10 20 15 20 20
Mo d el year

Figure 13. Percent change in distance-specific emissions for various pollutants by model year
for LDV and LDT combined, compared to model year 2005. Shaded region shows the 95%
confidence interval.

The observed increase in model year average UV smoke levels is connected to a shift in fuel
injection technologies. Gasoline Direct Injection (GDI) technology was first introduced to the
U.S. LDV market in 2008 and gained momentum relatively quickly, largely due to the higher fuel
efficiency and power compared to conventional port fuel injection (PFI) vehicles.118 As of 2021,
GDI vehicles represent approximately 53% of the US LDV market.

117 Particulate matter mass emissions are not measured directly by remote sensing systems; instead, a measurement
of UV smoke is recorded as a ratio of exhaust plume opacity measured at a wavelength of 230nm to fuel burned.
UV smoke measurements are then used as a proxy for PM emissions to assess long-term relative trends. More
detail is available in Meyer, M., Khan, T., Dallmann, T., Yang, Z. Particulate matter emissions from U.S. gasoline
light-duty vehicles and trucks: TRUE Initiative U.S. remote sensing database case study. ICCT, June 2023.
https://theicct.org/publication/true-pm-emissions-jun23/
118 U.S. EPA. (2022, December). The 2022 EPA Automotive Trends Report.

https://www.epa.gov/system/files/documents/2022-12/420r22029.pdf
47
Early GDI vehicles (model years 2008-2014) showed at least 2 times higher PM emissions
compared to PFI vehicles.119 This issue of elevated PM emissions from GDI vehicles and the
need for more stringent regulation of PM emissions has been identified since before EPA
finalized its Tier 3 standards in 2014.120 Recent analysis of remote sensing data adds to this
evidence, showing that for 2015–2020 model year vehicles, the models showing UV smoke
levels above the fleet average are predominantly GDI vehicles.121

Additionally, chassis dynamometer measurements based on the U.S. EPA’s certification tests
database122 combined with information from the U.S. Department of Energy fuel economy test
database123 demonstrate that PM emissions from GDI vehicles are higher than the PFIs on FTP
cycle. An analysis by manufacturer for model years 2019–2023, shows that for most of the
manufacturers that have sales for both GDI and PFI vehicles, GDIs have significantly higher PM
emissions than PFIs on average (Table 9). On the FTP cycle, GDI vehicles emitted 62%–490%
higher average emissions than PFI vehicles of the same manufacturer; on the US06 cycle, GDI
vehicles emitted 16% lower to 130% higher average emissions than PFI vehicles of the same
manufacturer.124

Table 9. Average PM emissions (mg/mile) for GDI and non-GDI vehicles by manufacturer for
2019-2023 model year vehicles
FTP cycle US06 cycle
Phased-in Tier 3 limit = Phased-in Tier 3 limit =
3 mg/mile 6 mg/mile

Manufacturer GDI PFI GDI PFI

A 1.53 0.26 1.53 1.32

B 1.14 0.41 1.49 0.64

C 1 0.23 1.63 1.92

D 0.99 0.18 2.26 1.55

E 0.76 0.47 2.19 0.94

F 0.72 0.26 1.47 1.39

119 Georges Saliba et al Comparison of Gasoline Direct-Injection (GDI) and Port Fuel Injection (PFI) Vehicle
Emissions: Emission Certification Standards, Cold-Start, Secondary Organic Aerosol Formation Potential, and
Potential Climate Impacts. (2017).Environmental Science & Technology 51, no. 11: 6542–52,
https://pubs.acs.org/doi/10.1021/acs.est.6b06509
120 Gladstein, Neandross & Associates. (2013). Ultrafine Particulate Matter and the Benefits of Reducing Particle

Numbers in the United States. https://cdn.gladstein.org/pdfs/MECA_UFP_White_Paper_0713_Final.pdf.


121 Meyer, M., Khan, T., Dallmann, T., Yang, Z. (2023) Particulate matter emissions from U.S. gasoline light-duty

vehicles and trucks: TRUE Initiative U.S. remote sensing database case study. International Council on Clean
Transportation. https://www.trueinitiative.org/data/publications/particulate-matter-emissions-from-us-gasoline-light-
duty-vehicles-and-trucks
122 U.S. EPA. Annual Certification Data for Vehicles, Engines, and Equipment Data and Tools.

https://www.epa.gov/compliance-and-fuel-economy-data/annual-certification-data-vehicles-engines-and-
equipment
123 U.S. Department of Energy, Fuel Economy Guide Datafile. (2022)

https://www.fueleconomy.gov/feg/download.shtml
124 Meyer, M., Khan, T., Dallmann, T., Yang, Z. (2023). Particulate matter emissions from U.S. gasoline light-duty

vehicles and trucks: TRUE Initiative U.S. remote sensing database case study. International Council on Clean
Transportation. https://www.trueinitiative.org/data/publications/particulate-matter-emissions-from-us-gasoline-light-
duty-vehicles-and-trucks
48
G 1.5   1.05  

H 0.94 1.63

I 0.81   1.7  

J 0.8   1.31  

K 0.51   1.51  

L   0.29   1.47

Additionally, an analysis of certification test data across all light-duty vehicles shows that PM
emissions from GDI vehicles have been reducing over the years; however, they still emit higher
PM than PFI vehicles on FTP cycle (Figure 14). For 2023 model year vehicles, GDI vehicles
emit on average about 3 times more PM emissions than the PFI vehicles on FTP cycle. For both
cycles, the trajectory of average PM emissions for GDI vehicles flattens out over the last few
years, suggesting that PM emissions from GDIs will not likely reduce substantially any further
without the adoption of more stringent emission standards.

Figure 14. Trend of average PM emissions for US LDV based on EPA certification database for
GDI and PFI vehicles under the FTP and US06 regulatory cycles.

Evaluation of proposed PM standard

49
The real-world and laboratory test data analyses presented above show the issue of increased
PM emissions from GDI vehicles, highlighting the need for a more stringent standard than the
current Tier 3 standard to prevent a continued increase in PM emissions. EPA’s proposal for PM
standards aligns with this finding; we strongly support the proposed stringency level of 0.5
mg/mile on all test cycles and the addition of the cold temperature FTP test cycle for regulating
PM emissions.

Table 10 shows that EPA’s proposed limit for PM emissions is more stringent than the
California’s most recently enacted LEV IV standards and, also, numerically lower than the
proposed Euro 7 standards in EU and the existing standards in China and India. However, EU,
China, and India have the additional particle number (PN) standards to regulate the ultrafine
particulates.

Table 10. Existing or proposed particulate emissions regulations across regions for gasoline
vehicles.
a Particle number (PN)
Regulation Phase-in timeframe Test cycle PM limit (mg/mile)
limit (#/mile) b
FTP, US06, cold
U.S. proposed >=2027 0.5 ---
FTP
California LEV IV FTP 1 ---
>=2026
enacted125 US06 3 ---
126
EU Euro 7 proposed >=2025 WLTP, RDE 7.2 9.7 x 1011
127
China 6b enacted >=2021 WLTP 4.8 9.7 x 1011
India Bharat VI
>=2020 NEDC 7.2c 9.7 x 1011 c
enacted128
a
All converted to mg/mile
b
All converted to number per mile
c
applicable only for GDI vehicles

The PN limits in the EU, China, and India regulations, combined with the particle mass limit,
effectively regulate both fine particulates (PM2.5) as well as ultrafine particulates (PM0.1).129
These particles, less than 100 nm in size, have been found to be even more dangerous to
human health than PM2.5, as they can be inhaled deeper into the lungs, increasing the chances
of particles to entering the bloodstream.130 Part of the motivation for the stringent particulate
regulations in the EU and China has been the uptake of GDI vehicles, as GDIs have been found

125
Bui, A., Hall, D., and Searle, S. (2022). Advanced Clean Cars II: The next phase of California’s Zero-Emission
Vehicle and Low-Emission Vehicle regulations. International Council on Clean Transportation.
https://theicct.org/publication/accii-zev-lez-reg-update-nov22/
126 Dornoff, J. (2023). How to make Euro 7 more effective: an analysis of the European Commission’s proposal for

light- and heavy-duty vehicles. International Council on Clean Transportation.


https://theicct.org/publication/euro7-analysis-recommendations-jan23/
127 TransportPolicy.net. China: Light-duty: Emissions. https://www.transportpolicy.net/standard/china-light-duty-

emissions/
128 TransportPolicy.net. India: Light-duty: Emissions. https://www.transportpolicy.net/standard/india-light-duty-

emissions/
129 Felix Leach et al. (2021). A Review and Perspective on Particulate Matter Indices Linking Fuel Composition to

Particulate Emissions from Gasoline Engines. SAE International Journal of Fuels and Lubricants 15, no. 1: 3–28
https://doi.org/10.4271/04-15-01-0001
130 . https://doi.org/10.1038/s12276-020-0403-3

50
to emit more ultrafine particulates down to a size of 23 nm and can emit even smaller particles
(sub 23 nm).131

While EPA’s proposed standard does not include a particle number limit, the stringency of the
PM emission standard will help address both PM2.5 and PM0.1 through the implementation of
gasoline particulate filters (GPFs). EPA suggests that the proposed PM emissions limit,
particularly for the cold FTP test, will likely force the installation of gasoline particulate filters
(GPF).132 Studies have demonstrated the effectiveness of GPFs for reducing both PM and PN
emissions, showing that GPFs can reduce PM emissions by 97% to 100% and PN emissions by
80% to 99% compared to vehicles without GPFs.133 Other major vehicle markets, including the
EU and China, have already begun adopting GPFs to meet stringent particulate emission
standards, and the proposed standard would help the US vehicle market follow suit. The
European Commission (2022) reported that the Euro 6d standards, which has led to the near-
universal adoption of GPFs to meet the PN limits for GDI vehicles, reduced the real-world
emission factors of exhaust particles by about 86% compared to Euro 5 compliant vehicles.134
The assessment also reported a typical reduction level of PN emissions by 70% to 80% for a
Euro 6 GDI installed with a GPF and tested on the Real-Driving Emissions (RDE) test route.

GPFs are cost-effective devices, with per vehicle manufacturing cost of $51–$166 based on
EPA estimates and $50–$184 based on other independent studies.135 Based on the European
Commission’s Euro 6/VI assessment, the incremental increase in technology cost per vehicle
for Euro 6 standards compared to Euro 5 was estimated at $92–$113 for the addition of GPF
along with lambda and pressure sensors (currency converted to U.S. dollars, 1 Euro = 1.10 US$
as of June 21, 2023).136 Thus, EPA’s proposed PM standard would provide large health benefits
through reduced fine and ultrafine particulate matter emissions with a relatively low added
manufacturing cost. This is in line with the European Commission’s finding that the Euro 6
benefits from PM and PN emissions reduction were significantly higher than the added cost for
GPF per vehicle.137

Testing of recent European petrol cars provides evidence that the proposed PM limit is
achievable. Cars certified to recent Euro standards have showed PM levels close to the
proposed US 0.5 mg/mile limit. ADAC, a German automobile association, conducted laboratory

131 Barouch Giechaskiel, Urbano Manfredi, and Giorgio Martini. (2014). Engine Exhaust Solid Sub-23 Nm Particles: I.
Literature Survey,” SAE International Journal of Fuels and Lubricants 7, no. 3: 950–64,
https://saemobilus.sae.org/content/2014-01-2834/
132 Felix Leach et al. (2021). A Review and Perspective on Particulate Matter Indices Linking Fuel Composition to

Particulate Emissions from Gasoline EnginesSAE International Journal of Fuels and Lubricants 15, no. 1: 3–28,
https://doi.org/10.4271/04-15-01-0001
133 Felix Leach et al. (2021). A Review and Perspective on Particulate Matter Indices Linking Fuel Composition to

Particulate Emissions from Gasoline Engines,” SAE International Journal of Fuels and Lubricants 15, no. 1: 3–28,
https://doi.org/10.4271/04-15-01-0001.; Jiacheng Yang et al. (2018). Gasoline Particulate Filters as an Effective
Tool to Reduce Particulate and Polycyclic Aromatic Hydrocarbon Emissions from Gasoline Direct Injection (GDI)
Vehicles: A Case Study with Two GDI Vehicles. Environmental Science & Technology 52, no. 5: 3275–84.
https://doi.org/10.1021/acs.est.7b05641
134 European Commission. (2022, October). Euro 6/VI evaluation study: Annexes 1-6.

https://op.europa.eu/en/publication-detail/-/publication/6fd483af-5f1a-11ed-92ed-01aa75ed71a1/language-en
135 Minjares, R and Posada Sanchez, F. (2011). Estimated cost of gasoline particulate filters. International Council on

Clean Transportation. https://theicct.org/publication/estimated-cost-of-gasoline-particulate-filters/; Steininger.


(2011). Particle number emission limits for Euro 6 positive ignition vehicles.
https://www.nanoparticles.ch/archive/2011_Steininger_PR.pdf
136 European Commission. (2022, October). Euro 6/VI evaluation study: Annexes 1-6.

https://op.europa.eu/en/publication-detail/-/publication/a9a2eadb-5f1d-11ed-92ed-01aa75ed71a1/language-en
137 ibid

51
tests, which showed that out of 177 gasoline vehicles certified to Euro 6d-TEMP and Euro 6d
standards, 74 vehicles, or 42%, showed PM emissions below 0.5 mg/mile.138 Additionally, the
UK Department for Transport conducted laboratory testing of five 2019 model year petrol
vehicles.139 Under the hot start Worldwide Harmonized Light Vehicles Test Cycle (WLTC) test,
the cars emitted 0.09–0.28 mg/km (0.14–0.45 mg/mile), all below 0.5 mg/mile. Under the cold
start WLTC test, the cars emitted 0.16–0.446 mg/km (0.26–0.72 mg/mile); two cars emitted
under 0.5 mg/mile, and the other three showed PM emissions of no more than 44% above 0.5
mg/mile. Although these results are not directly comparable to the US standard due to
differences in test cycles (WLTC instead of FTP and US06), this evidence demonstrates that
there already exist vehicles with PM emissions levels similar to the proposed US PM limit.

We support the EPA’s accelerated phase-in option for PM standards relative to other pollutants,
particularly considering evidence of rising gasoline light-duty vehicle and truck PM levels.
Analysis of remote sensing UV smoke measurements as a proxy for PM shows that reductions
since 2005 model year vehicles have been virtually eliminated due to an increase from 2015 to
2020 model year vehicles. Therefore, it is critical that a more stringent PM standard is
implemented as soon as possible to drive the uptake of GPFs. GPFs are at a mature phase and
in large-scale production in several vehicle markets and thus would be feasible to implement
according to EPA’s outlined accelerated timeline, which would phase-in the proposed PM
standards for 50% or 80% of the manufacturer’s fleet by 2027 and 100% by 2028. Due to the
urgent need to address the rise in PM emissions with recent gasoline vehicles, we recommend
EPA to adopt the accelerated phase-in pathway for the PM standard.

Health benefits and environmental justice modeling recommendations

In evaluating the health benefits of the proposed rule, it is critical to consider the evidence of
increases in PM from recent model year vehicles. The health benefits assessment included in
EPA’s proposal uses emission factors from MOtor Vehicle Emission Simulator (MOVES), which
likely underestimates PM emissions from recent model year vehicles. Figure 15 shows a
comparison between EPA MOVES emission factors and remote sensing UV smoke averages
by model year, shown as a percent change from each sources respective model year 2005
averages.140 From model years 2006–2015, year-to-year changes, shown in the slopes of the
lines, are relatively consistent between EPA MOVES and remote sensing data. However, from
model year 2015 on, the trends diverge. For light-duty passenger trucks, the EPA MOVES PM
emission factors sharply decline while the remote sensing data show an increase in UV smoke
averages.141 Though less significant, a similar trend is observed for passenger cars, with the
EPA MOVES PM emission factor continuing to decline after 2015 and remote sensing UV
smoke averages showing a slight increase. These findings suggest that the modeled air quality
and health benefits of the proposed rule may be underestimated when considering the large

138 These results are based on averages of three tests: hot start Worldwide Harmonized Light Vehicles Test Cycle
(WLTC), cold start WLTC, and BAB 130, which is ADAC’s highway cycle. Data was kindly provided by ADAC via
email correspondence (2023).
139 Department for Transport. (2023). Vehicle Market Surveillance Unit: Results of the 2020 programme. (London,

UK). https://www.gov.uk/government/publications/vehicle-market-surveillance-unit-programme-results-
2020/vehicle-market-surveillance-unit-results-of-the-2020-programme#results-petrol-cars
140 EPA MOVES emission factors are converted from fuel-specific to distance-specific emission factors for this

comparison. This is done using real-world fuel economy data from EPA Automotive Trends report, the same
source used to convert the remote sensing emissions from fuel-specific to distance-specific.
141 Light-duty passenger truck emission trends from EPA MOVES are compared to light-duty truck (LDT) emission

trends from remote sensing data. Similarly, passenger car emission trends from EPA MOVES are compared to
light-duty vehicle (LDV) emission trends from remote sensing data.
52
benefits of replacing 2015–2020 model year vehicles with future vehicles meeting the proposed
PM emission limits.

Figure 15. Percent change in distance-specific PM emission factor from EPA MOVES and
remote sensing (RS) UV smoke measurements from Colorado (CO) and Virginia (VA) sources
by model year, compared to model year 2005.

Additionally, a full assessment of relative air quality and health benefits across demographics,
as done for the heavy-duty multi-pollutant rule, is recommended for this light-duty and medium-
duty multi-pollutant rule. It is widely understood that heavy-duty vehicles contribute to racial
disparities in exposure to air pollution, and EPA reported the projected air quality impacts across
demographics in their finalized heavy-duty multi-pollutant rule. Inequities in air pollution
exposure exist for light-duty vehicle emissions as well. One study finds that on a national level,
people of color are exposed to 46% more ambient PM 2.5 from light-duty gasoline vehicles than
White people.142 This is a greater disparity than that of HDVs; the same study finds that people
of color exposed to 35% higher ambient PM 2.5 levels from heavy-duty diesel vehicles compared
to White people. Additionally, the health benefits from the heavy-duty multipollutant rule, as
projected by the EPA, are on similar scales to that of the proposed light-duty and medium-duty
vehicle rule.143 Thus, as the environmental justice and health implications of the light-duty and
medium-duty vehicle rule are significant, a full assessment of the projected distribution of
changes in PM2.5 and ozone concentrations by geography, race/ethnicity, and income is
recommended for the finalized rule.

142 Tessum, C et al. (2021). PM2.5 Polluters Disproportionately and Systemically Affect People of Color in the United
States,” Science Advances 7, no. 18: eabf4491, https://doi.org/10.1126/sciadv.abf4491.
143 EPA’s proposed light-duty and medium-duty rule is estimated to result in $63 billion and $280 billion (for 7% and

3% discount rates, respectively) in total monetized health benefits from 2027 to 2055. EPA’s finalized heavy-duty
multipollutant rule is projected to result in $53–150 billion and $91–260 billion (for 7% and 3% discount rates,
respectively) in total monetized health benefits from 2027 to 2045.
53
Benefits and cost-effectiveness of the proposal
The LDV and MDV proposal is projected to reduce 15,000 tons of PM 2.5, 66,000 tons of NOx,
and 220,000 tons of hydrocarbons from 2027 to 2055.144 These benefits are compared to 2055
levels without the proposal. Based on the previous section on Particulate Matter emission
standards, health benefits associated with PM reductions could be greater than EPA’s
projections. The proposed standards would reduce air pollution near-road where affected
populations are often low-income or communities of color. Reducing these emissions will
provide cleaner air and are critical to improving public health.

The total benefits of the proposal exceed the total cost. Per EPA’s estimate, the overall
technology cost of the proposal ranges from $180 billion to $280 billion for automakers (in 2020
dollars) through 2055. However, the consumers will benefit from the pre-tax fuel savings that
range from $450 to $890 billion, and repair and maintenance savings from $280 to $580 billion.
In addition, the social benefits include $330 billion in climate benefits and between $63 and
$280 billion in criteria pollutant reduction benefits.

The proposal states that the estimated average costs for automakers to meet the proposed
standards would range from $200 to $1,600 (in 2020 dollars) per vehicle in MY 2032, across the
range of sensitivities. Per EPA’s analysis, this estimated MY 2032 average costs of $1,200
represent under three percent of the average cost of a new vehicle today (about $46,000 in
2022). This cost resonates with previous 2012 and 2021 rules which estimated that the cost to
meet standards were about $1,800 (2010 dollars) and $1,000 (2018 dollars) per vehicle.
Similarly, previous ICCT analysis shows that the cost for an ICE vehicle to meet the 2012 rules
or Tier 2 standards would range from $405 to $690.145

In addition to the feasible cost, the continuing technology improvement will bring greater climate
and health benefits. As discussed in the DRIA Chapter 3.2.4, the introduction of more electric
vehicles will help make the declining FTP NMOG + NOX fleet average fully feasible. The
MY2021 test data shows 19 vehicles with emissions performance currently below 15 mg/mile,
and two below 10 mg/mile from a range of automakers. Similarly, MDV 2022 and 2023 also
show similar low emission performance compared to the proposal limit.

PROPOSED STRINGENCY
ICCT supports the proposed standards and recommends EPA finalize Alternative 1. Our
research shows that the proposed standards are likely less costly than estimated and can be
met with a variety of technological approaches and pathways, evidenced by the BEV and ICE
technology updates discussed above. Alternative 1 is even more cost-effective and delivers
greater environmental and health benefits than the proposed standards.

Comparing Tables 6, 17-19 of the Preamble, Alternative 1 is projected to result in higher net
benefits than any other standards contained in the proposal. Moreover, fuel savings alone offset
increased technology costs several times over. As discussed above, BEV and ICE technology

144 US EPA. (2023). Multi-Pollutant emission standards for Model years 2027 and later Light-duty and Medium-duty
vehicles Program Announcement. https://www.epa.gov/system/files/documents/2023-04/420f23009.pdf
145 Sanchez, F., Bandivadekar, A., and German, J. (2012). Estimated cost of emission reduction technologies for

Light-Duty vehicles. International Council on Clean Transportation. https://theicct.org/wp-


content/uploads/2021/06/ICCT_LDVcostsreport_2012.pdf
54
cost, effectiveness, and availability that has yet to be incorporated by EPA’s modeling will make
complying with Alternative 1 standards easier and less costly than currently modeled.
Furthermore, with widespread adoption of California’s ACC II program in other states, the cost
of Alternative 1 decreases substantially. Comparing DRIA Tables 13-45 and 13-49, by MY2032
the projected manufacturing cost under Alternative 1 (versus the ACC II adoption sensitivity
baseline) are nearly identical to the costs of the proposed standards (versus the central analysis
baseline without ACCII adoption).

We believe that initial front-loading of both the proposed standards and Alternative 1 standards
(faster improvement in early years) is reasonable and that both the Proposal and Alternative 1
can be met cost-effectively because of ICE and BEV technology availability (discussed above).
Additionally, announced ZEV product launches, finalization of ACC II, passing of the IRA, ICE
phaseouts worldwide, and recent growth in BEVs in the U.S. suggest automakers may over
comply in MY2023-2026 (Preamble I.A.2.ii, 29187). Such overcompliance could lead to credit
banking which could be used to comply with MY2027 as proposed (Preamble 29240).

OBSERVATIONS ABOUT PROPOSED FOOTPRINT CURVES


ICCT supports EPA’s proposed standards’ stringency, as well as the shape of the footprint
curves. Nevertheless, data suggests that the footprint curves can be adjusted both to increase
stringency (within the limits of the proposal) as well as to slow or reverse the trend of upsizing
and the shift from cars to trucks.

ICCT commends EPA for its proposed standards becoming increasingly flat. The proposal to
flatten the curves over time and move the truck cutpoints inwards are a welcome shift that can
help prevent upsizing to some extent.

ICCT also commends EPA for reducing the gap between the truck and the car curves, and for
its transparency and request for comments on the development of the truck curve (Preamble
III.B.2.ii; DRIA 1.1.3.2). By EPA’s own acknowledgement, and supported by recent data, there is
clear evidence that EPA could flatten both the car and the truck curves even further. This is
discussed below.

EPA assumed in the OMEGA model inputs that the willingness-to-pay (WTP) of vehicle footprint
is $200/sq ft, which is “on the low end of the range suggested in the literature (e. a. Greene
2018)” and recognizes “a higher WTP would create a stronger upsizing tendency, which would
suggest an even flatter "size-neutral" slope”. The Greene, (2018) paper cited by EPA refers to
the summary in Whitefoot and Skerlos, (2012) that the average marginal WTP ranges from
$340 to $2,000 for an additional square foot of vehicle size (i.e., the overall length of a vehicle
multiplied by the width). Considering the potentially wide difference in the average consumer
valuation of vehicle attributes in the literature, EPA could assume the mid-point, rather than the
low end, of the average values of WTP from literature as the basis to determine the slope of the
footprint-based standard curve. As a result, the slope of the passenger cars curve, which is also
used as the baseline to develop the truck curve, would be flatter than the proposed standard
curves.

EPA recognizes that the majority of light-duty vehicles serve to “move passengers and their
nominal cargo” which implies light-duty vehicles “could be represented by a single curve”
(Preamble 29234). ICCT supports the application of a single curve for all light-duty vehicles, due
to the similar purpose and utility across all light vehicles. However, in the absence of a single

55
curve for all light-duty vehicles, the proposed approach of deriving the truck curve from the car
curve is generally sound. As discussed further below, ICCT recommends flattening the truck
curve further by reducing the offset allocated to trucks.

To justify the truck curve offset, EPA states that it “identified a subset of light trucks…which are
more appropriately controlled with a modified set of standards” (Preamble 29234). EPA
identifies these vehicles as those “larger trucks which are designed for more towing and hauling
capability [and] require design changes to allow for handling of these larger loads and this is
reflected in increased engine capability, body-on-frame design, and greater structural mass”
[emphasis added] (DRIA 1.1.3.2, page 1-8). EPA’s analysis of fleet tailpipe CO2 emissions data
from these vehicles establishes the “utility offset” which is built into the truck curve. (EPA also
analyzed an AWD offset, but this offset is a significantly lower g/mi adjustment than the utility
offset.) Importantly, EPA states: “many crossover vehicles and SUVs exhibit similar towing
capability between their 2WD and AWD versions” (Preamble 29235).

Since many crossovers and SUVs classified as trucks differ from their car variants only due to
the presence of AWD, these truck crossovers and SUVs benefit from the utility offset of the
truck curve, without demonstrating any purported need for such a utility. This leads to targets
that are comparatively easy to meet for these vehicles. EPA further states, “The purpose of
maintaining a unique truck curve is centered around accounting for the utility of [full size
pickups] in particular” (Preamble 29235).

In summary, the utility offset is calculated by and added to the truck curve to specifically account
for the higher emissions of full-size pickups, or body-on-frame vehicles, as these vehicles are
distinguished by “[use] for more work-like activity” (Preamble 29235). Because the utility offset is
based on the vehicles most likely to be used for their higher utility, the offset ideally would be
calculated based on the number of vehicles that meet those criteria. Bearing these criteria in
mind, there is evidence that the utility offset could be adjusted lower.

Although the utility offset is based on pickups, since this offset is added to the truck curve, it
applies to all vehicles classified as trucks. EPA projects that pickups will represent 24% of all
trucks in MY2027-2032.146 Thus, only 24% of trucks meet the criteria EPA used to determine the
utility offset. Furthermore, by MY2030, EPA projects the MY2030 pickup share to be 45% (EPA
chooses MY2030 as a reference point, DRIA 1-12). The initial estimated utility offset of 63 g/mi
(DRIA 1-11) could be reduced accordingly to account for the projected share of vehicles that
actually have higher utility and non-zero tailpipe emissions:
Initial utility offset: 63 g/mi
MY2030 pickup share: 24%
MY2030 ICE pickup share: 55%
ICE pickup share-adjusted utility offset: 8.3 g/mi

Alternatively, the utility offset could be recalculated according to the projected MY2030 sales
share of body-on-frame vehicles, which is roughly one third. Of these, roughly half are ICE,
leading to a body-on-frame-based utility offset of about 11 g/mi.

In addition to the characteristic of having greater structural mass (mainly due to body-on-frame
construction), EPA stipulates that increased engine capability also supports the addition of the
utility offset. However, as discussed previously (see Atkinson cycle and MHEV subsections),

146 central analysis output file, 2023_03_14_22_42_30_central_3alts_20230314_Proposal_vehicles.csv


56
most pickups and large SUVs spend most driving time under low load. Contrary to EPA’s
assertion that these vehicles are used for more work-like activity (Preamble 29235), data
suggests truck drivers rarely use the extra utility for which their light-duty trucks are built. One
2019 study found 75% of truck owners tow once or never each year, 35% haul once or never
each year, and 70% go off-road once or less per year.147 In a 2023 report, 63% of F150 owners
were found to tow rarely or never, 32% rarely or never hauled, over half frequently commuted in
their trucks, and 87% frequently used their trucks for shopping/errands.148 That is, despite
having more capable engines, data supports the conclusion that pickups rarely use this extra
capability. Consequently, when outfitted with advanced engines, such vehicles are likely to
spend more time than cars in highly efficient combustion modes, which are better suited for
lower load driving. Pickups and SUVs also stand to benefit the most from advances in MHEV
technologies. As current MY2023-2026 and proposed MY2027-2032 standards engender
efficient pickups, the real-world utility impact will diminish, due to both greater efficiency and
high share of low load driving. This conclusion further supports reducing the truck utility offset.

PLUG-IN HYBRID ELECTRIC VEHICLES

Utility factor
ICCT applauds EPA for adjusting the PHEV utility factor (UF) curve to better fit real-world PHEV
usage data. As the PHEV models available for purchase today differ in operation and all-electric
utility from when the original UF curve was developed, this update is direly need (ICCT 2022
PHEV).149 Due to the growing repository of public, real-world PHEV data, ICCT recommends
EPA adjust the proposed fleet utility factor (FUF) curve to reflect the best-available data, while
allowing for potential updates to the curve based on more real-world data as it accumulates.

In DRIA figure 3-29, EPA’s analysis of the California Bureau of Automotive Repair (BAR) PHEV
data clearly shows that a reduced UF curve is a better fit of the real-world data. This fit matches
ICCT’s non-linear regression fit of the BAR data. Utilizing the same methodology described in
ICCT 2022 PHEV but with EPA’s proposed FUF coefficients (Preamble 29442), ICCT finds the
appropriate normalized distance to be 802 miles (vs proposed 583 miles). This curve more
accurately reflects the current state of PHEV usage. Thus, ICCT recommends that EPA adopt a
normalized distance of 802 miles, instead of 583 miles as proposed.

Within the timeframe of the proposed rule EPA expects PHEVs with longer all-electric range to
become available (in part due to California’s ACCII PHEV performance requirements), and EPA
seeks to avoid disincentivizing PHEVs with a too low FUF curve (Preamble 29254). For these
reasons, EPA proposed a FUF curve higher than what the BAR data supports. This argument
puts the expectation of future PHEV performance ahead of the data. However, because there is
no guarantee that future PHEV models will achieve the assumed higher electric driving shares,
the best practice is to let the data dictate the shape of the UF curve. EPA can establish a
provision by which it can adjust the UF curve to better account for PHEVs with varying all-

147 Berk, B. (2019, March 13). You Don’t Need a Full-Size Pickup Truck, You Need a Cowboy Costume.
Thedrive.com.
https://www.thedrive.com/news/26907/you-dont-need-a-full-size-pickup-truck-you-need-a-cowboy-costume
148 Chase, W., Whalen, J., Muller, J. (2023, Jan 23) Pickup Trucks: from workhorse to joyride. Axios.

https://www.axios.com/ford-pickup-trucks-history
149 Isenstadt, A., Yang, Z., Searle, S., German, J. (2022). Real world usage of plug-in hybrid vehicles in the United

States. International Council on Clean Transportation. https://theicct.org/publication/real-world-phev-us-dec22/


(ICCT 2022 PHEV)
57
electric range, as additional data is collected. To simultaneously ensure PHEVs are not given
too-high UF without disincentivizing longer all-electric range PHEVs, EPA can allow
manufacturers to provide publicly available, real-world data with accurate UF measurements
that support higher FUF for specific PHEV models. Alternatively, EPA could automatically apply
the higher FUF curve (e.g., the current proposed curve, as opposed to the ICCT-BAR curve) for
vehicles that meet minimum all-electric performance requirements, such as 70 miles all-electric
2-cycle range and 40 miles all-electric US06 range. This latter option would incentivize greater
all-electric capability among PHEVs.

Incorporating PHEVs in the proposal analysis


EPA did not include PHEVs in its OMEGA modeling for this proposal; ICCT recommends EPA
include PHEVs in its updated modeling for the final rule, if possible. Since PHEVs are eligible to
be used for compliance with the standards, incorporating them into the modeling would more
accurately project compliance pathways. If PHEVs are cost effective in some years and
segments, including them as an option in the modeling would more accurately reflect the lower
compliance costs that could be achieved by PHEVs in those uses compared to modeling that
excludes them.

For reference, ICCT previously analyzed the costs of light-duty150 and medium-duty PHEVs.151
For light-duty vehicles, ICCT projected PHEVs to have higher prices and 6-year ownership
costs than 300-mile BEVs of all body styles before MY2027. Light-duty PHEVs are also
expected to remain costlier than non-plugin vehicles for the foreseeable future.

For medium-duty pickups, ICCT projected gasoline PHEVs to reach purchase price parity with
300-mile range BEV pickups around MY2027, but likely to remain costlier than both diesel and
gasoline non-plugin pickups. Diesel PHEVs are not expected to reach price parity with either
BEVs or non-plugin pickups. For medium-duty vans, gasoline PHEVs are expected to reach
price parity with diesel non-plugin vehicles by MY2032, but 300-mi and 150-mile range BEVs
are expected to cost less than PHEV vans throughout MY2027-2032. When it comes to 5-year
total cost of ownership, medium-duty gasoline PHEV pickups and vans are expected to cost
less than diesel non-plugin vehicles before MY2027 (diesel PHEVs are still expected to cost
more throughout the timeframe of the proposed rule).

As discussed in the combustion vehicle technology section previously, developments in


dedicated hybrid engines suitable for (primarily gasoline) serial/range-extended PHEV
applications may simultaneously improve the efficiency of the engine while also reducing its
cost. Thus, EPA may consider modeling PHEVs in the largest light-duty segments, as well as
both medium-duty pickups and vans. Doing so could lead to reduced overall compliance costs
as compared to those currently modeled in the proposal.

ALPHA validations of P2 and PS hybrids were based on PHEVs (DRIA Table 2-6). With PHEV-
specific data, response surface equations (RSEs) for PHEVs could be developed and
incorporated into OMEGA. Absent full simulation and RSEs for PHEVs, EPA can approximate

150 Slowik. P., Isenstadt, A., Pierce, L, Searle, S. (2022). Assessment of light-duty electric vehicle costs and
consumer benefits in the United States in the 2022-2035 time frame. International Council on Clean
Transportation. https://theicct.org/publication/ev-cost-benefits-2035-oct22/
151 Mulholland, E. (2022). Cost of electric commercial vans and pickup trucks in the United States through 2040.

ICCT. International Council on Clean Transportation. https://theicct.org/publication/cost-ev-vans-pickups-us-2040-


jan22/
58
PHEVs by combining BEV RSE with P2 or PS RSEs (utilizing UF for electric/ICE split). Such a
combination would need to be validated against the existing PHEV test data.

MEDIUM-DUTY VEHICLES
ICCT supports EPA's proposed standards for MDVs but believes the available evidence could
justify greater stringency than proposed. ICCT also recommends the MDV standards be fuel
neutral and require public availability of MDV-related data (sales, emissions, fuel consumption,
vehicle attributes/technologies).152 These recommendations form the basis for ICCT’s
comments on the proposed MDV standards, discussed further below.

Stringency
ICCT supports the proposed MDV standards but believes they could be strengthened. The
proposed standards would not require as high GHG reductions for MDVs as for light-duty trucks.
The proposed MDV standards, if finalized, would require a net 36.5% improvement in tailpipe
CO2 emissions, for a given work factor over MY2027-2032. This total improvement is much
higher than that required by the current Phase 2 standards. However, the proposed LD truck
standards require a total 46% improvement for a given footprint over MY2027-2032. This level
of improvement comes after years of increased efficiency required by MY2017-2026
regulations. Figure 16 compares the 2-cycle CO2-e emissions of ICE LD pickups, MD vans, and
MD pickups versus footprint. Each datapoint represents a single ICE model as calculated by
OMEGA in MY2022 (baseline MY2021 emissions values are not available for all models, and
OMEGA projects virtually 0% BEV share of these segments in MY2022). Since MD vehicles are
tested at higher weight than LD vehicles, to compare MD and LD data, the emissions data of
MDV were adjusted downwards assuming a test weight equal to curb weight plus 300 lbs
(equivalent to the LD test weight). It was assumed that for every 10% reduction in test weight,
CO2 emissions decreased by 4%.153 Solid-filled points represent a single model, and pattern-
filled points represent the sales-weighted average emissions in MY2022, and the projected
emissions in MY2032 to comply with the standard.154 Projected emissions were calculated
based on the proposed standard curves for MY2032 assuming constant footprint (for LD
pickups) or constant work factor with average test weight adjustment (for MD vehicles). As the
figure shows, LD pickups are more efficient than similarly sized MDV, and the required
improvement over the course of MY2022-MY2032 is much higher for LD than for MD. The
different utility of LD vs MD notwithstanding, due to years of lagging behind the stringency levels
of LDV, there are opportunities for increasing the stringency of the MD standards.

152 Lutsey. N. (2015). Regulatory considerations for advancing commercial pickup and van efficiency technology in
the United States. International Council on Clean Transportation.
https://theicct.org/sites/default/files/publications/ICCT_pickup-van-efficiency_20150417.pdf
153 Isenstadt, A. and German, J (ICCT); Piyush Bubna and Marc Wiseman (Ricardo Strategic Consulting);

Umamaheswaran Venkatakrishnan and Lenar Abbasov (SABIC); Pedro Guillen and Nick Moroz (Detroit
Materials); Doug Richman (Aluminum Association), Greg Kolwich (FEV). (2016). Lightweighting technology
development and trends in U.S. passenger vehicles. http://www.theicct.org/lightweighting-technology-
development-and-trends-us-passenger-vehicles
154 Preamble Tables 27 & 31

59
700

600

500
Oncycle CO2e (g/mi)

2022 MD pickup models


400
2022 MD van models
2022 LD pickup models
300 -38% MD a verage pickup improvement
MD a verage van improvement
200 -45% LD average pickup improvement

100 -70%

0
50 60 70 80 90
Footprint (sq ft)

Figure 16. Comparison of on-cycle CO2 emissions of light- and medium-duty trucks, by model
(solid fill) and sales-weighted average (pattern fill) calculated in OMEGA in MY2022. Average
MY2022 and projected required MY2032 emissions are shown connected by arrows, indicating
the percent reduction in emissions over MY2022-2032.

As acknowledged by EPA (DRIA 1-19), many gasoline and diesel efficiency-improving


technologies have yet to be broadly implemented among medium-duty vehicles. In particular,
many LD-related technologies like those discussed in previous sections in these comments can
be applied to MD vans and some pickups. For instance, strong and mild hybrid systems that are
well-suited for full-size LD SUVs and pickups could be scaled up for MD implementation. Light
heavy-duty diesel powertrains can also benefit from hybridization and heavy-duty versions of LD
efficiency technologies.155,156,157 Hybrid versions of LD vans and pickups today suggest that
mild, strong, or plugin hybridization may be viable options for certain MD applications.158

155
Isenstadt, A., German, J. (2017). Diesel Engines. International Council on Clean Transportation.
https://theicct.org/publication/diesel-engines/
156 Buysse, C., Sharpe, B., Delgado, O. (2021). Efficiency technology potential for heavy-duty diesel vehicles in the

United States through 2035. International Council on Clean Transportation.


https://theicct.org/publication/efficiency-technology-potential-for-heavy-duty-diesel-vehicles-in-the-united-states-
through-2035/
157 Posada, F., Isenstadt, A., Badshah, H. (2020). Estimated cost of diesel emission-control technology to meet the

future California low NOx standards in 2024 and 2027. International Council on Clean Transportation.
https://theicct.org/publication/estimated-cost-of-diesel-emissions-control-technology-to-meet-the-future-california-
low-nox-standards-in-2024-and-2027/
158 According to the MY2023 Fuel Economy Guide (https://fueleconomy.gov/feg/download.shtml), the Ford F150

hybrid is 17%-20% more efficient in combined, unadjusted (2-cycle) fuel consumption than its non-hybrid

60
If EPA incorporated into its analysis additional cost curve classes (incorporating MD versions of
advanced gasoline engines and hybrid powertrains), as well as additional options for
lightweighting (high strength steel body and frame and/or percentage-based mass reduction),
and PHEVs (see PHEV discussion earlier), then its MD analysis would be much more robust.

Just as many LD ICE technology improvements extend to MD ICE vehicles, all of the
innovations and developments in BEV technologies and battery packs discussed earlier also
apply to both LD and MD classes. As with LDV BEV modeling, incorporating lower battery costs,
rightsized motors, and improved EV efficiency within OMEGA for MDV would reduce the cost of
MD BEVs. In a 2022 study on relative costs of EV MDV, ICCT projected that MD pickups and
vans of 300-mile range or less would reach price parity with their diesel counterparts within the
timeframe of this proposal.159 Over the first five years of ownership, MD BEVs of 300-mile range
or less reach total cost parity with both their diesel and gasoline counterparts before 2030. In
other words, well within the timeframe of the proposal battery-electric MDVs are cheaper to
purchase and own than non-plugin MDVs.

Discussed further below, it is difficult to assess specific areas to improve the analysis of the MD
proposed standards due to minimal explanation of the technical inputs into the MD analysis, and
how they differ from or are similar to the LD analysis.

Regarding some finer details of the proposed MD standards, ICCT supports EPA’s proposal to
phaseout ZEV multipliers in MY2026 instead of MY2027. The expected growth in MD BEV van
share by MY2027 could result in overcompliance credits being available in MY2027, which can
be used to offset undercompliance by vans and pickups which are still ICE-powered. Generally,
ending multipliers earlier would improve the real-world climate benefits of the proposed rule.
Additionally, recent ICCT research on the costs of BEV and PHEV MD vans and pickups
suggests that within the timeframe of the proposed rule, plug-in MDVs will be cost-competitive
with their gasoline or diesel counterparts either upfront or both upfront and during ownership.160
Consequently, such advanced technology vehicles will not need multipliers as a production
incentive. Lastly, ICCT also supports the proposed 22,000 lbs limit to the GCWR input in the
work factor equation, for the reasons EPA provided in its proposal (Preamble 29242).

Fuel neutrality
ICCT strongly supports EPA’s switch to a single set of standards curves for all MDVs,
regardless of fuel type.

counterpart with the same 3.5L engine; the Pacifica PHEV in charge sustaining mode is 31% more efficient than
its equivalent non-hybrid counterpart (3.6L engine); Tundra HEV is 5.3%-10.3% more efficient than non-HEV
equipped with Atkinson cycle engine; Ram 1500 mild hybrid is 13.3%-16% more efficient than its non-hybrid
version
159 Mulholland, E. (2022). Cost of electric commercial vans and pickup trucks in the United States through 2040.

ICCT. International Council on Clean Transportation. https://theicct.org/publication/cost-ev-vans-pickups-us-2040-


jan22/
160 Mulholland, E. (2022). Cost of electric commercial vans and pickup trucks in the United States through 2040.

International Council on Clean Transportation. https://theicct.org/publication/cost-ev-vans-pickups-us-2040-jan22/


61
MDV data and transparency
EPA’s fuel economy guides, annual certification data, and automotive trends reports are
invaluable resources for detailed information on light-duty sales, emissions, fuel consumption,
vehicle attributes, and applied technologies. This level of detail is absent for medium-duty
vehicles. Public access to this data is necessary for many reasons, including the ability to
identify trends in pickup and van sales, possible shifts from light-duty to medium-duty, and
technology penetration and effectiveness. ICCT recommends EPA collect and make public
these data for MDV as it does for LDV.

Paralleling the limited public MDV data availability, the discussion of MDV technology costs and
effectiveness within the Preamble and DRIA is not as robust as that discussion for LDV. While
certain MD-specific technologies are mentioned (e.g. the MD Ford 7.3L engine which was not
incorporated in the analysis, DRIA 3.5.1.2), it is unclear what engine maps apply to MDV, if any
technology costs/effectiveness were borrowed from LDV, how LDV costs/effectiveness were
modified for MDV (if at all), and if there are any other modeling parameters specific to MDV.

By examining several OMEGA MD input and output files, there are several areas ICCT finds
worthy of comment.

The central “vehicles” output file161 shows that all diesel vehicles throughout MY2021-MY2035
are considered to have Miller cycle engines (MIL) and all gasoline vehicles are GDI only. Only
these 2 ICE technology packages are available for MD, and both are without any level of
electrification (micro through strong hybrid).162 This lack of additional ICE technology packages
suggest EPA is significantly underestimating the possibility for improved ICE-based MD pickups
and vans. As there are many cost-effective ICE technologies available, the lack of detail in the
analysis likely overstates the cost of compliance. Relatedly, it is unclear which engine (and
electric motor) maps were used to develop the technology packages’ RSEs, as well as the
appropriateness of approximating diesel engines’ CO2e emissions with those of Miller cycle
engines. The MD cost equations appear identical to the LD costs.163 While this may be
appropriate for many technologies that are shared between LD and MD, there are certain
performance and utility differences between LD and MD that may lead to differing costs
(consider, in particular, costs related to diesel powertrain improvement, which are unassessed
in the LD analysis).

The notion that MDV have different use cases than LDV also suggests that consumer and
manufacturer purchase and production decisions may differ between MD and LD. As MDV are
generally used in commercial applications, consumers likely consider the purchase of MDV as a
business decision. Consequently, MDV purchasers are more likely than LD purchasers to
consider the full cost of ownership when making such a decision. In particular, MDV buyers may
fully value the fuel savings from efficiency technologies or BEVs. Because of this full valuation
of fuel savings, ICCT recommends EPA model MDV consumers as valuing at least 5 years of

161 2023_03_17_13_57_07_md_central_v3_Proposal_vehicles.csv
162 simulated_vehicles_rse_ice_MD_input_rse_only.csv
163 powertrain_cost_20230314.csv

62
fuel savings (double the 2.5 years of fuel savings currently used for both LD and MD164).
According to ICCT’s 2022 MD EV cost study, over the first 5 years of ownership, before 2030
both MD BEV pickups and vans are expected to cost less to own and operate than their diesel
and gasoline counterparts.165

164 According to producer_generalized_cost-body_style_20220613.csv—used in both the LD and MD central


analyses
165 Mulholland, E. (2022). Cost of electric commercial vans and pickup trucks in the United States through 2040.

International Council on Clean Transportation. https://theicct.org/publication/cost-ev-vans-pickups-us-2040-jan22/


63

You might also like