Jump to main content.


Control of Emissions of Air Pollution From Nonroad Diesel Engines and Fuel [[pp. 28427-28476]]

 [Federal Register: May 23, 2003 (Volume 68, Number 100)]
[Proposed Rules]
[Page 28427-28476]
From the Federal Register Online via GPO Access [wais.access.gpo.gov]
[DOCID:fr23my03-38]
 
[[pp. 28427-28476]]
Control of Emissions of Air Pollution From Nonroad Diesel Engines 
and Fuel

[[Continued from page 28426]]
[[Page 28427]]

than 10 ppm sulfur. Refiners producing only high sulfur distillate 
today should have an added advantage in meeting a 15 ppm sulfur cap for 
nonroad fuel over that for highway fuel. They would be able to design 
their hydrotreater from the ground up, while most refiners producing 15 
ppm diesel fuel for highway use will be trying to utilize their 
existing 500 ppm hydrotreaters, which may not be designed to be 
revamped to produce 15 ppm fuel in the most efficient manner.
---------------------------------------------------------------------------

    \260\ ``Highway Diesel Progress Review,'' EPA, June 2002, 
EPA420-R-02-016.
---------------------------------------------------------------------------

    Based on our review of the limited catalyst performance data in the 
published literature and the one set of confidential data submitted, we 
believe that the projections of the more optimistic vendors are the 
most accurate for the 2010 timeframe given this additional leadtime. 
For example, the confidential commercial data indicated that five ppm 
sulfur levels could be achieved with two-stage hydrotreating at 
moderate hydrogen pressure despite the presence of a significant amount 
of light cycle oil (LCO). The key factor was the inclusion of a 
hydrogenation catalyst in the second stage, which saturated many of the 
poly-nuclear, aromatic rings in the diesel fuel, allowing the removal 
of sulfur from the most sterically hindered compounds. In addition, 
refiners that are able to defer production of 15 ppm highway diesel 
fuel through the purchase of credits, as well as refiners producing 15 
ppm nonroad in 2010, would have the added benefit of being able to 
observe the operation of those hydrotreating units starting up in 2006. 
This should allow these refiners to be able to select from the best 
technologies which are employed in the highway program.
    In addition, a number of alternative technologies are presently 
being developed which could produce 15 ppm fuel at lower cost. 
ConocoPhillips, for example, has developed a version of their S-Zorb 
technology for diesel fuel desulfurization. This technology utilizes a 
catalytic adsorbent to remove the sulfur atom from hydrocarbon 
molecules. It then sends the sulfur-laden catalyst to a separate 
reactor, where the sulfur is removed and the catalyst is restored. 
Unipure is developing a process which selectively oxidizes the sulfur 
contained in diesel fuel. This process have the advantage that the 
sulfur containing compounds which are most difficult to desulfurize via 
hydrotreating are quite easily desulfurized via oxidation. Finally, 
Linde has developed a method which greatly improves the concentration 
of hydrogen on hydrotreating catalysts. This process promises to 
greatly reduce the reactor volume necessary to produce 15 ppm diesel 
fuel.
    These three new technologies are at various stages of development. 
This is discussed in more detail in the next section. Due to the 
projected ability of these technologies to reduce the cost of meeting a 
15 ppm sulfur cap and the leadtime available between now and 2010, we 
project that 80% of the new volume of 15 ppm nonroad diesel fuel would 
be produced using advanced technologies.
7. Has Technology to Meet a 15 ppm Cap Been Commercially Demonstrated?
    EPA just completed a review of refiners' progress in preparing to 
produce 15 ppm highway diesel fuel.\261\ The information we obtained 
during that review confirm the projections we made in the HD 2007 
program--refiners are technically capable of producing 15 ppm sulfur 
diesel fuel using extensions of conventional technology and, in fact, 
they are moving forward with their plans to comply with the program. 
Thus, we believe there are no technological hurdles to producing 15 ppm 
diesel fuel.
---------------------------------------------------------------------------

    \261\ Ibid.
---------------------------------------------------------------------------

    The European Union has also determined that diesel fuel can be 
desulfurized to meet a sulfur cap in the range of 10-15 ppm. Europe has 
established a 10 ppm sulfur cap on highway diesel fuel, effective in 
2009, with plans underway for a 10 ppm sulfur cap for nonroad diesel 
fuel soon thereafter. As with our standards, Europe's 10 ppm cap 
applies throughout the distribution system. However, fuel tends to be 
transported much shorter distances in Europe. Therefore, we believe 
that both the 10 and 15 ppm sulfur caps will require refiners to meet 
the same 7-8 ppm sulfur target at the refinery gate. Given this, the 
European standard will require the same technology as that required in 
the U.S. Most European diesel fuel must meet a higher cetane number 
specification than U.S. diesel fuel, which causes it to be 
predominantly comprised of straight run material. This material is 
easier to desulfurize to sub-15 ppm levels using conventional 
hydtrotreating technology. In some European countries, nonroad diesel 
fuel is the same as heating oil and contains significant amounts of 
cracked material. Thus, on average, it should be easier for European 
refiners to meet a 10 ppm sulfur cap with their highway diesel fuel 
than in the U.S. As the 10 ppm cap is extended to nonroad diesel fuel, 
the stringency of the European standard will be much closer to that of 
a 15 ppm cap here in the U.S.
    We have met with a number of diesel fuel refiners to learn about 
their plans to produce 15 ppm highway diesel fuel by the June 2006 
program compliance date. Since the 15 ppm diesel fuel sulfur standard 
was established based on the use of extensions of conventional diesel 
desulfurization technologies, diesel fuel refineries are well 
positioned to make firm plans for implementation by 2006. Our review 
has found that this is exactly what refiners are doing. We are very 
encouraged by the actions some refiners have already taken in terms of 
announcing specific plans for low sulfur diesel fuel production. It may 
still be early in the process, but virtually all refiners are already 
in the stage of planning their approach for compliance. Thus, the 
refining industry is where we anticipated it would be at this point in 
time. Moreover, some refining companies are ahead of schedule and will 
be capable of producing significant quantities of 15 ppm sulfur diesel 
fuel as early as next year. Thus, we expect that the capability of 
conventional hydrotreating to produce 15 ppm diesel fuel in refinery-
scale quantities will be demonstrated in the U.S. by the end of 2003.
    Phillips Petroleum is currently in the process of designing and 
constructing a commercial sized S-Zorb unit to produce sub-15 ppm 
diesel fuel at their Sweeney, Texas refinery. This plant is scheduled 
to begin commercial operation in 2004. This would provide refiners with 
roughly 3 years of operating data before they would have to decide 
which technology to use to meet the 15 ppm nonroad sulfur cap in 2010. 
This should be enough operating experience for most refiners to have 
sufficient confidence in this advanced process to include it in their 
options for 2010 compliance. Based on information received from 
Phillips Petroleum, we estimate that this technology could reduce the 
cost of meeting the 15 ppm cap for many refiners by 25 percent.
    Linde has also developed a new approach for improving the contact 
between hydrogen, diesel fuel and conventional desulfurization 
catalysts. Linde projects that their Iso-Therming process could reduce 
the hydrotreater volume required to achieve sub-15 ppm sulfur levels by 
roughly a factor of 2. Linde has already built a commercial-sized 
demonstration unit at a refinery in New Mexico and has been operating 
the equipment since September 2002. Thus, refiners would have 4-5 years 
of operating data available on this process before they would have to 
decide which technology to use to meet the 15 ppm nonroad sulfur cap in 
2010. This should be ample operating experience for

[[Page 28428]]

essentially all refiners to include this process in their options for 
2010. Based on information received from Linde, we estimate that this 
technology could reduce the cost of meeting the 15 ppm cap for many 
refiners by 40 percent.
    Finally, Unipure Corporation is developing a desulfurization 
process which oxidizes the sulfur atom in diesel fuel molecules, 
facilitating its removal. This process operates at low temperatures and 
ambient pressure, so it avoids the need for costly, thick walled, 
pressure vessels and compressors. It also consumes no hydrogen. Thus, 
it could be particularly advantageous for refiners who lack an 
inexpensive supply of hydrogen (e.g., isolated or smaller refineries 
who cannot construct a world scale hydrogen plant based on inexpensive 
natural gas). However, the oxidant is very powerful, so specialized, 
oxidation resistant materials are needed. Unipure has demonstrated its 
process at the pilot plant level, but has yet to build a commercial 
sized demonstration unit. However, time still remains for this to be 
done before refiners need to make final decisions for their 2010 
compliance plans. Thus, while more uncertain than the other two 
advanced processes, the Unipure oxidation process could be selected by 
a number of refiners to meet the 2010 15 ppm cap. Based on inputs from 
Unipure, we estimate that their process could reduce the cost of 
meeting the 15 ppm cap for roughly one-fourth of all refineries by 25-
35 percent.
    The savings associated with each technology varies with the size, 
location and complexity of the refinery. However, on average the Linde 
process appears to have the potential reduce the cost of desulfurizing 
500 ppm diesel fuel to 15 ppm by 35-40 percent. The savings associated 
with the Phillips and Unipure processes appear to be more refinery 
specific. For about 25 refineries, the Phillips process appears to have 
the potential to reduce these desulfurization costs by 20-40 percent. 
The primary advantage of the Unipure process is its lower capital 
costs. For about 30 refineries, the Unipure process appears to have the 
potential to reduce the capital investment related to produce 15 ppm 
fuel from 500 ppm diesel fuel by an average of 40 percent.
8. Availability of Leadtime To Meet the 2010 15 ppm Sulfur Cap
    If we promulgate this proposal one year from today, this would 
provide refiners and importers with more than six years before they 
would have to begin complying with the 15 ppm cap for nonroad diesel 
fuel on June 1, 2010. Our leadtime analysis, which is presented in the 
draft RIA, projects that 30-39 months are typically needed to design 
and construct a diesel fuel hydrotreater.\262\ Thus, refiners would 
have about 3 years before they would have to begin detailed design and 
construction. This would allow them time to observe the performance of 
the hydrotreaters being used to produce 15 ppm highway diesel fuel for 
at least one year. While not a full catalyst cycle, any unusual 
degradation in catalyst performance over time should be apparent within 
the first year. Thus, we project that the 2010 start date would allow 
refiners to be quite certain that the designs they select in mid-2007 
will perform adequately in 2010.
---------------------------------------------------------------------------

    \262\ ``Highway Diesel Progress Review,'' USEPA, EPA420-R-02-
016, June 2002.
---------------------------------------------------------------------------

    In addition, we expect that most of the advanced technologies will 
be demonstrated on a commercial scale by the end of 2004. Thus, 
refiners would have at least two and a half years to observe the 
performance of these technologies before having to select a technology 
to meet the 2010 15 ppm cap. This should be more than adequate to fully 
access the costs and capabilities of these technologies for all but the 
most cautious refiners.
9. Feasibility of Distributing Nonroad, Locomotive and Marine Diesel 
Fuels That Meet the Proposed Sulfur Standards
    There are two considerations with respect to the feasibility of 
distributing non-highway diesel fuels meeting the proposed sulfur 
standards. The first pertains to whether sulfur contamination can be 
adequately managed throughout the distribution system so that fuel 
delivered to the end-user does not exceed the specified maximum sulfur 
concentration. The second pertains to the physical limitations of the 
system to accommodate any additional segregation of product grades.
a. Limiting Sulfur Contamination
    With respect to limiting sulfur contamination during distribution, 
the physical hardware and distribution practices for non-highway diesel 
fuel do not differ significantly from those for highway diesel fuel. 
Therefore, we do not anticipate any new issues with respect to limiting 
sulfur contamination during the distribution of non-highway fuel that 
would not have already been accounted for in distributing highway 
diesel fuel. Highway diesel fuel has been required to meet a 500 ppm 
sulfur standard since 1993. Thus, we expect that limiting contamination 
during the distribution of 500 ppm non-highway diesel engine fuel can 
be readily accomplished by industry.
    In the highway diesel rule, EPA acknowledged that meeting a 15 ppm 
sulfur specification would pose a substantial new challenge to the 
distribution system. Refiners, pipelines and terminals would have to 
pay careful attention to and eliminate any potential sources of 
contamination in the system (e.g., tank bottoms, deal legs in 
pipelines, leaking valves, interface cuts, etc.) In addition, bulk 
plant operators and delivery truck operators would have to carefully 
observe recommended industry practices to limit contamination, 
including practices as simple as cleaning out transfer hoses, proper 
sequencing of fuel deliveries, and parking on a level surface. Due to 
the need to prepare for compliance with the highway diesel program, we 
anticipate that issues related to limiting sulfur contamination during 
the distribution of 15 ppm nonroad diesel fuel will be resolved well in 
advance of the proposed 2010 implementation date for nonroad fuel. We 
are not aware of any additional issues that might be raised unique to 
nonroad fuel. If anything we anticipate limiting contamination will 
become easier as batch sizes are allowed to increase and potential 
sources of contamination decrease. We request comment on whether there 
are unique considerations regarding the transition to a 15 ppm standard 
for nonroad diesel fuel and what actions we should take beyond those 
that are already underway in preparation for the 15 ppm highway diesel 
program.
b. Potential Need for Additional Product Segregation
    As discussed in sub-section B, we have designed the proposed 
program to minimize the need for additional product segregation and the 
associated feasibility and cost issues associated with it. This 
proposal would allow for the fungible distribution of 500 ppm highway 
and 500 ppm NRLM diesel fuel in 2007, and 15 ppm highway and 15 ppm 
nonroad diesel fuel in 2010, up until the point where NRLM or nonroad 
fuel must be dyed for IRS excise tax purposes. Heating oil would be 
required to be segregated as a separate pool beginning in 2007 through 
the use of a new marker, and locomotive and marine fuel by use of the 
same marker beginning in 2010. With this program design, we believe we 
have eliminated any potential feasibility issues associated with the 
need for product segregation. This is not to say that steps will not 
have to be taken. We have

[[Page 28429]]

identified only a single instance where it seems likely that the 
adoption of this proposal would result in entities in the distribution 
system choosing to add new tankage due to new product segregation. Bulk 
plants in areas of the country where heating oil is expected to remain 
in the market will have to decide whether to add tankage to distribute 
both heating oil and 500 ppm NRLM fuel. In all other cases we 
anticipate segments of the distribution system will choose to avoid any 
fuel segregation costs by limiting the range of sulfur grades they 
choose to carry, just as they do today. Regardless, however, the costs 
and impacts of these choices are small. We request comment on this 
assessment. A more detailed explanation of this assessment can be found 
in Chapter 5.6 of the draft RIA.

G. What Are the Potential Impacts of the 15 ppm Sulfur Diesel Program 
on Lubricity and Other Fuel Properties?

1. What Is Lubricity and Why Might it Be a Concern?
    Engine manufacturers and owner/operators depend on diesel fuel 
lubricity properties to lubricate and protect moving parts within fuel 
pumps and injection systems for reliable performance. Unit injector 
systems and in-line pumps, commonly used in diesel engines, are 
actuated by cams lubricated with crankcase oil, and have minimal 
sensitivity to fuel lubricity. However, rotary and distributor type 
pumps, commonly used in light and medium-duty diesel engines, are 
completely fuel lubricated, resulting in high sensitivity to fuel 
lubricity. The types of fuel pumps and injection systems used in 
nonroad diesel engines are the same as those used in highway diesel 
vehicles. Consequently, nonroad and highway diesel engines share the 
same need for adequate fuel lubricity to maintain fuel pump and 
injection system durability.
    Diesel fuel lubricity concerns were first highlighted for private 
and commercial vehicles during the initial implementation of the 
Federal 500 ppm sulfur highway diesel program and the state of 
California's diesel program. The Department of Defense (DoD) also has a 
longstanding concern regarding the lubricity of distillate fuels used 
in its equipment as evidenced by the implementation of its own fuel 
lubricity improver performance specification in 1989.\263\ The diesel 
fuel requirements in the state of California differed from the federal 
requirements by substantially restricting the content of diesel fuel 
requires more severe hydrotreating than reducing the sulfur content to 
meet a 500 ppm standard.\264\ Consequently, concerns regarding diesel 
fuel lubricity have primarily been associated with California diesel 
fuel and some California refiners treat their diesel fuel with a 
lubricity additive as needed. Outside of California, hydrotreating to 
meet the current 500 ppm sulfur specification does not typically result 
in a substantial reduction of lubricity. Diesel fuels outside of 
California seldom require the use of a lubricity additive. Therefore, 
we anticipate only a marginal increase in the use of lubricity 
additives in NRLM diesel fuel meeting the proposed 500 ppm sulfur 
standard for 2007.\265\ This proposal would require diesel fuel used in 
nonroad engines to meet a 15 ppm sulfur standard in 2010. Based on the 
following discussion, we believe that the increase in the use of 
lubricity additives in 15 ppm nonroad diesel fuel would be the same as 
that estimated for 15 ppm highway diesel fuel.
---------------------------------------------------------------------------

    \263\ DoD Performance Specification, Inhibitor, Corrosion/
Lubricity Improver, Fuel Soluble, , MIL-PRF-25017F, 10 November 
1997, Superseding MIL-I-25017E, 15 June 1989.
    \264\ Chevron Products Diesel Fuel Technical Review provides a 
discussion of the impacts on fuel lubricity of current diesel fuel 
compositional requirements in California versus the rest of the 
nation. http://www.chevron.com/prodserv/fuels/bulletin/diesel/l2%5F7%5F2%5Frf.htm.
Exit Disclaimer
    \265\ The cost from the increased use of lubricity additives in 
500 ppm NRLM diesel fuel in 2007 and in 15 ppm nonroad diesel fuel 
in 2010 is discussed in section V of today's preamble.
---------------------------------------------------------------------------

    The state of California currently requires the same standards for 
diesel fuel used in nonroad equipment as in highway equipment. Outside 
of California, highway diesel fuel is often used in nonroad equipment 
when logistical constraints or market influences in the fuel 
distribution system limit the availability of high sulfur fuel. Thus, 
for nearly a decade nonroad equipment has been using federal 500 ppm 
sulfur diesel fuel and California diesel fuel, some of which may have 
been treated with lubricity additives. During this time, there has been 
no indication that the level of diesel lubricity needed for fuel used 
in nonroad engines differs substantially from the level needed for fuel 
used in highway diesel engines.
    Blending small amounts of lubricity-enhancing additives increases 
the lubricity of poor-lubricity fuels to acceptable levels. These 
additives are available in today's market, are effective, and are in 
widespread use around the world. Among the available additives, 
biodiesel has been suggested as one potential means for increasing the 
lubricity of conventional diesel fuel. Indications are that low 
concentrations of biodiesel would be sufficient to raise the lubricity 
to acceptable levels.
    Considerable research remains to be performed to better understand 
which fuel components are most responsible for lubricity. Consequently, 
it is unclear whether and to what degree the proposed sulfur standards 
for non-highway diesel engine fuel will impact fuel lubricity. 
Nevertheless, there is evidence that the typical process used to remove 
sulfur from diesel fuel--hydrotreating--can impact lubricity depending 
on the severity of the treatment process and characteristics of the 
crude. We expect that hydrotreating will be the predominant process 
used to reduce the sulfur content of non-highway diesel engine fuel to 
meet the 500 ppm sulfur standard during the first step of the proposed 
program. The highway diesel program projected that hydrotreating would 
be the process most frequently used to meet the 15 ppm sulfur standard 
for highway diesel fuel. The 2010 implementation date for the proposed 
15 ppm standard for nonroad diesel fuel would allow the use of new 
technologies to remove sulfur from fuel.\266\ These new technologies 
have less of a tendency to affect other fuel properties than does 
hydrotreating.
---------------------------------------------------------------------------

    \266\ See section IV.F for a discussion of which desulfurization 
processes we expect will be used to meet the 15 ppm standard for 
nonroad diesel fuel.
---------------------------------------------------------------------------

    Based on our comparison of the blendstocks and processes used to 
manufacture non-highway diesel fuels, we believe that the potential 
decrease in the lubricity of these fuels from hydrotreating that might 
result from the proposed sulfur standards should be approximately the 
same as that experienced in desulfurizing highway diesel fuel.\267\ To 
provide a conservative, high cost estimate, we assumed that the 
potential impact on fuel lubricity from the use of the new 
desulfurization processes would be the same as that experienced when 
hydrotreating diesel fuel to meet a 15 ppm sulfur standard. We request 
comment on the potential impact of these new desulfurization 
technologies on lubricity (as well as other fuel properties) that might 
help us to improve our estimate of the potential impacts of this 
proposal on fuel properties other than sulfur. Given that the 
requirements for fuel lubricity in highway and non-highway engines are 
the same, and the potential decrease in lubricity from desulfurization 
of non-highway diesel engine would be no greater than that experienced 
in desulfurizing highway diesel fuel, we

[[Page 28430]]

estimate that the potential need for lubricity additives in non-highway 
diesel engine fuel under this proposal would be the same as that for 
highway diesel fuel meeting the same sulfur standard.
---------------------------------------------------------------------------

    \267\ See chapter 5 of the RIA for a discussion of the potential 
impacts on fuel lubricity of this proposal.
---------------------------------------------------------------------------

2. A Voluntary Approach on Lubricity
    In the United States, there is no government or industry standard 
for diesel fuel lubricity. Therefore, specifications for lubricity are 
determined by the market. Since the beginning of the 500 ppm sulfur 
highway diesel program in 1993, refiners, engine manufacturers, engine 
component manufacturers, and the military have been working with the 
American Society for Testing and Materials (ASTM) to develop protocols 
and standards for diesel fuel lubricity in its D-975 specifications for 
diesel fuel. ASTM is working towards a single lubricity specification 
that would be applicable to all diesel fuel used in any type of engine. 
Although ASTM has not yet adopted specific protocols and standards, 
refiners that supply the U.S. market have been treating diesel fuel 
with lubricity additives on a batch to batch basis, when poor lubricity 
fuel is expected. Other examples include the U.S. military, Sweden, and 
Canada. The U.S. military has found that the traditional corrosion 
inhibitor additives used in its fuels have been highly effective in 
reducing fuel system component wear. Since 1991, the use of lubricity 
additives in Sweden's 10 ppm sulfur Class I fuel and 50 ppm sulfur 
Class II fuel has resulted in acceptable equipment durability.\268\ 
Since 1997, Canada has required that its 500 ppm sulfur diesel fuel not 
meeting a minimum lubricity be treated with lubricity additives.
---------------------------------------------------------------------------

    \268\ Letter from L. Erlandsson, MTC AB, to Michael P. Walsh, 
dated October 16, 2000. EPA air docket A-99-06, docket item IV-G-42.
---------------------------------------------------------------------------

    The potential need for lubricity additives in diesel fuel meeting a 
15 ppm sulfur specification was evaluated during the development of 
EPA's highway diesel rule. In response to the proposed highway diesel 
rule, all comments submitted regarding lubricity either stated or 
implied that the proposed sulfur standard of 15 ppm would likely cause 
the refined fuel to have lubricity characteristics that would be 
inadequate to protect fuel injection equipment, and that mitigation 
measures such as lubricity additives would be necessary. However, the 
commenters suggested varied approaches for addressing lubricity. For 
example, some suggested that we need to establish a lubricity 
requirement by regulation while others suggested that the current 
voluntary, market based system would be adequate. The Department of 
Defense recommended that we encourage the industry (ASTM) to adopt 
lubricity protocols and standards before the 2006 implementation date 
of the 15 ppm sulfur standard for highway diesel fuel.
    The final highway diesel rule did not establish a lubricity 
standard for highway diesel fuel. We believe the issues related to the 
need for diesel lubricity in fuel used in non-highway diesel engines 
are substantially the same as those related to the need for diesel 
lubricity for highway engines. Consequently, we expect the same 
industry-based voluntary approach to ensuring adequate lubricity in 
non-highway diesel fuels that we recognized for highway diesel fuel. We 
believe the best approach is to allow the market to address the 
lubricity issue in the most economical manner, while avoiding an 
additional regulatory scheme. A voluntary approach should provide 
adequate customer protection from engine failures due to low lubricity, 
while providing the maximum flexibility for the industry. This approach 
would be a continuation of current industry practices for diesel fuel 
produced to meet the current federal and California 500 ppm sulfur 
highway diesel fuel specifications, and benefits from the considerable 
experience gained since 1993. It would also include any new 
specifications and test procedures that we expect would be adopted by 
the American Society for Testing and Materials (ASTM) regarding 
lubricity of NRLM diesel fuel quality.
    Regardless, this is an issue that will be resolved to meet the 
demands of the highway diesel market, and whatever resolution is 
reached for highway diesel fuel could be applied to non-highway diesel 
engine fuel with sufficient advance notice. We are continuing to 
participate in the ASTM Diesel Fuel Lubricity Task Force \269\ and will 
assist their efforts to finalize a lubricity standard in whatever means 
possible. We are hopeful that ASTM can reach a consensus early this 
summer at the next meeting of the ASTM's Lubricity Task Force. We 
request comment on what actio ns EPA should take to ensure adequate 
lubricity of non-highway diesel engine fuel beyond those already 
underway for highway diesel fuel.
---------------------------------------------------------------------------

    \269\ ASTM sub committee D02.E0.
---------------------------------------------------------------------------

3. What Other Impact Would Today's Actions Have on the Performance of 
Diesel and Other Fuels?
    We do not expect that the proposed fuel program would have any 
negative impacts on the performance of diesel engines in the existing 
fleet which would use the fuels regulated today. In the early 1990's, 
California lowered the maximum allowable level of sulfur content of 
highway and nonroad diesel fuel to 500 ppm, and at the same time 
California significantly lowered the aromatic content of diesel fuel. 
California required a cap on total aromatics of 10 percent by volume, 
while the in-use average at the time was on the order of 35 percent. 
The lowering of the total aromatic content resulted in some problems 
with leaks from the fuel pump O-ring seals in some diesel engines due 
to a change specifically in the polynuclear aromatics content (PNA). In 
the process of meeting California's 10 percent total aromatic content 
requirement, the end result typically lowered PNA's from approximately 
10-15 percent by volume to near-zero. In the early 1990's, some diesel 
engine manufacturers used a certain material (Nitrile) for O-rings in 
diesel fuel pumps. The Nitrile seals were found to be susceptible to 
leakage with the use of diesel fuel with very low PNA content. 
Normally, the PNA in the fuel penetrated the Nitrile material and cause 
it to swell, thereby providing a seal with the throttle shaft. When 
very low PNA fuel is used after conventional fuel has been used, the 
PNA already in the swelled O-ring would leach out into the very low PNA 
fuel. Subsequently, the Nitrile O-ring would shrink and pull away, thus 
causing leaks, or the stress on the O-ring during the leaching process 
would cause it to crack and leak. Not all 500 ppm sulfur fuels caused 
this problem, because the amount and type of aromatics varied, and the 
in-use seal problems were focused in California due to the 10 percent 
aromatic requirements and the resulting very low PNA content. This was 
not a wide-spread issue for the rest of the U.S. where highway diesel 
fuel also had a 500ppm sulfur cap because the federal requirements did 
not include a lower aromatic cap. While the process of lowering sulfur 
levels to 500ppm does lower PNA, it does not achieve the near-zero 
levels seen in California. Since the 1990's, diesel engine 
manufacturers have switched to alternative materials (such as Viton), 
which do not experience leakage. We believe that no issues with leaking 
fuel pump O-rings would occur with the changes in diesel fuel sulfur 
levels

[[Page 28431]]

contained in this proposal (both the 500 ppm requirement in 2008 and 
the 15 ppm requirement in 2010) because while we do believe PNA content 
will be reduced, we are not predicting it will achieve the near-zero 
level experienced in California.
    We expect that this proposal would have no negative impacts on 
other fuels, such as jet fuel or heating oil. We do expect that the 
sulfur levels of heating oil would decrease because of this proposal. 
Beginning in mid-2007, we expect that controlling NRLM diesel fuel to 
500 ppm would lead many pipelines to discontinue carrying high sulfur 
heating oil as a separate grade. In areas served by these pipelines, 
heating oil users would likely switch to 500 ppm diesel fuel. This 
would reduce emissions of sulfur dioxide and sulfate PM from furnaces 
and boilers fueled with heating oil. The primary exception to this 
would likely be the Northeast and some areas of the Pacific Northwest, 
where a distinct higher sulfur heating oil would still be distributed 
as a separate fuel. Also, we expect that a small volume of high sulfur 
distillate fuel would be created during distribution from the mixing of 
low sulfur diesel fuels and higher sulfur fuels, such as jet fuel in 
the pipeline interface. Such high sulfur distillate would likely be 
sold by the terminal as high sulfur heating oil or reprocessed by 
transmix processors.

H. Refinery Air Permitting

    Prior to making diesel desulfurization changes, some refineries may 
be required to obtain a preconstruction permit, under the New Source 
Review (NSR) program, from the applicable state/local air pollution 
control agency.\270\ We believe that the proposed program provides 
sufficient lead time for refiners to obtain any necessary NSR permits 
well in advance of the compliance date.
---------------------------------------------------------------------------

    \270\ Hydrotreating diesel fuel involves the use of process 
heaters, which have the potential to emit pollutants associated with 
combustion, such as NOX, PM, CO and SO3. In 
addition, reconfiguring refinery processes to add desulfurization 
equipment could increase fugitive VOC emissions. The emissions 
increases associated with diesel desulfurization would vary widely 
from refinery to refinery, depending on many source-specific 
factors, such as crude oil supply, refinery configuration, type of 
desulfurization technology, amount of diesel fuel produced, and type 
of fuel used to fire the process heaters.
---------------------------------------------------------------------------

    Given that today's diesel sulfur program would provide roughly 
three years of lead time before the 500 ppm standard would take effect, 
we believe refiners would have time to obtain any necessary 
preconstruction permits. Nevertheless, we believe it is reasonable to 
continue our efforts under the Tier 2 and highway diesel fuel programs, 
to help states in facilitating the issuance of permits under the NRLM 
diesel sulfur program. For example, the guidance on Best Available 
Control Technology (BACT) and Lowest Achievable Emission Rate (LAER) 
control technology that was developed for the gasoline sulfur program 
should have application for diesel desulfurization (highway and NRLM) 
projects as well. Similarly, we believe the concept of EPA permit teams 
for gasoline sulfur projects could readily be extended to permits 
related to diesel projects as well. These teams, as needed, would track 
the overall progress of permit issuance and would be available to 
assist state/local permitting authorities, refineries and the public 
upon request to resolve site-specific permitting questions. In 
addition, these teams would be available, as necessary, to assist in 
resolving case specific issues to ensure timely issuance of permits. 
Finally, to facilitate the processing of permits, we encourage 
refineries to begin discussions with permitting agencies and to submit 
permit applications as early as possible.

V. Program Costs and Benefits

    In this section, we present the projected cost impacts and cost 
effectiveness of the proposed nonroad Tier 4 emission standards and 
low-sulfur fuel requirement. We also present a benefit-cost analysis 
and an economic impact analysis. The benefit-cost analysis explores the 
net yearly economic benefits to society of the reduction in mobile 
source emissions likely to be achieved by this rulemaking. The economic 
impact analysis explores how the costs of the rule will likely be 
shared across the manufacturers and users of the engines, equipment and 
fuel that would be affected by the standards.
    The results detailed below show that this rule would be highly 
beneficial to society, with net present value benefits through 2030 of 
$550 billion, compared to a net present value of social cost of only 
about $16.5 billion (net present values in the year 2004). The impact 
of these costs on society should be minimal, with the prices of goods 
and services produced using equipment and fuel affected by the proposal 
being expected to increase about 0.02 percent.
    Further information on these and other aspects of the economic 
impacts of our proposal are summarized in the following sections and 
are presented in more detail in the Draft RIA for this rulemaking. We 
invite the reader to comment on all aspects of these analyses, 
including our methodology and the assumptions and data that underlie 
our analysis.

A. Refining and Distribution Costs

    As described above, the fuel-related requirements associated with 
this proposed rule would be implemented in two steps. Nonroad, 
locomotive and marine diesel fuel would be subject to a 500 ppm sulfur 
cap beginning June 1, 2007, while nonroad diesel fuel would be subject 
to a 15 ppm sulfur cap beginning June 1, 2010. Meeting these standards 
would generally require refiners adding hydrotreating equipment and 
possibly new or expanded hydrogen and sulfur plants in their refineries 
for desulfurizing their nonroad diesel fuel and dispensing of the 
removed sulfur. Using information provided by vendors of 
desulfurization equipment and through discussions with distributors of 
nonroad diesel fuel, we estimated the desulfurization and associated 
distribution and additive cost for complying with this two step 
desulfurization program. Except for the costs presented at the end of 
this section, the costs below reflect a fully phased in fuels program 
without the proposed small refiner exemption. Costs are in 2002 
dollars. We request comment on the cost estimates presented below and 
the methodologies used to develop them. You can refer to the Draft RIA 
for details.
    The cost to provide nonroad, locomotive and marine diesel fuel 
under the proposed fuel program is summarized in Table V-A-1 below. The 
costs shown (and all of the costs described in the rest of this 
section) only apply to the roughly 65 percent of current nonroad, 
locomotive and marine diesel fuel that contains more than 500 ppm 
sulfur (hereafter referred to as the affected volume). We estimate that 
the other 35 percent of this fuel is actually fuel certified to the 
highway diesel fuel standards and project that this will continue. 
Thus, the proposed fuel program would not affect this fuel and no 
additional costs would be incurred by its refiners or distributors. The 
costs and benefits of desulfurizing this highway fuel which spills over 
into the non-highway markets was already included in EPA's 2007 highway 
diesel fuel rule.

[[Page 28432]]

              Table V-A-1.--Increased Cost of Providing Nonroad, Locomotive and Marine Diesel Fuel
----------------------------------------------------------------------------------------------------------------
                                                         Cents per gallon of affected fuel         Affected fuel
                                                 ------------------------------------------------     volume
                                                                                                     (million
                                                     Refining      Lubricity and       Total       gallons/year)
                                                                   distribution                          a
----------------------------------------------------------------------------------------------------------------
Step One--500 ppm NRLM diesel fuel..............             2.2             0.3             2.5           9,504
Step Two--5 ppm Nonroad diesel fuel.............             4.4             0.4             4.8           7,803
Step Two--500 ppm Locomotive and Marine diesel               2.2           b 0.2             2.4          4,093
 fuel...........................................
----------------------------------------------------------------------------------------------------------------
Notes:
a 2008 for Step One (without consideration of small refiner provisions), 2015 for Step Two.
b 0.4 cent per gallon from mid-2010 to mid-2014 due to need for marker.

    The majority of the fuel-related cost of the proposal is refining-
related. These costs include required capital investments amortized at 
7 percent per annum before taxes. The derivation of these costs is 
discussed in more detail below and in the Draft RIA. We request comment 
on the estimated cost of meeting the 15 ppm and 500 ppm sulfur caps.
    We also project that the increased cost of refining and 
distributing 15 ppm and 500 ppm fuel would be substantially offset by 
reductions in maintenance costs. These savings would apply to all 
diesel engines in the field, not just new engines. Refer to section V. 
B for a more complete discussion on the projected maintenance savings 
associated with lower sulfur fuels.
1. Refining Costs
    Our process for estimating the refining costs associated with the 
proposed fuel program consisted of four steps. One, we estimated the 
volume of 500 and 15 ppm nonroad, locomotive and marine diesel fuel 
which had to be produced in each PADD \271\ in each phase of the 
program. This step utilized diesel fuel and heating oil use estimates 
from the Energy Information Administration's (EIA) Fuel Oil and 
Kerosene Survey for 2000, shipments of diesel fuel between PADDs, 
projected loss of 15 and 500 ppm volume due to contamination during 
distribution and small refiner provisions. This nonroad diesel fuel 
consumption in 2000 is lower than that inherent in the emission 
estimates described above, which are based directly on the results of 
EPA's NONROAD emission model. We are investigating ways to make the two 
estimates more consistent.
---------------------------------------------------------------------------

    \271\ Petroleum Administrative for Defense Districts.
---------------------------------------------------------------------------

    Growth in distillate fuel use off this year 2000 base was estimated 
using projections from EIA's Annual Energy Outlook, with one exception. 
This exception was that the growth in nonroad diesel fuel use was taken 
from EPA's NONROAD emission model (roughly three percent per year), as 
opposed to EIA's projected growth of roughly one percent per year. The 
higher growth rate is consistent with that inherent in the emission 
estimates described above.
    Refinery production of low and high sulfur distillate fuel in the 
year 2000 was based on actual reports provided to EIA by all U.S. 
refiners and importers. Refinery production of low and high sulfur 
distillate fuel was assumed to grow at the same rate as consumption of 
the two types of fuel, respectively. These rates were roughly three 
percent and one and a half percent for low and high sulfur distillate 
fuel production, respectively. The specific volumes of highway, 
nonroad, locomotive, and marine diesel fuel by calendar year are 
presented in chapter 7 of the Draft RIA.
    Two, we estimated the cost for each refinery to desulfurize its 
high sulfur fuel to 500 and 15 ppm. This was based on their historical 
production volume of high sulfur diesel fuel and estimates of the 
composition of this fuel (straight run, light cycle oil, etc.).\272\ We 
also considered whether these refineries would be modifying or building 
hydrotreating capacity in order to meet the 15 ppm highway cap.
---------------------------------------------------------------------------

    \272\ The composition of nonroad diesel fuel in each PADD was 
based on a survey conducted by API and NPRA in 1996. Crude oils 
processed by domestic refiners have been becoming heavier over time, 
necessitating greater use of coking and hydrocracking to convert the 
heavy material into lighter, saleable products. Thus, the 
contributions of coker and hydrocracked distillate to the overall 
distillate pool are rising. Coker distillate is somewhat more 
difficult to desulfurize than average distillate, but hydrocracked 
distillate is much easier to desulfurize. Overall, this trend could 
increase projected desulfurization costs slightly. We plan to update 
these compositions to reflect trends in crude oil quality and 
refinery configuration in our analysis for the final rule to the 
extent that more recent data allow.
---------------------------------------------------------------------------

    Three, we estimated which refineries would find it difficult to 
market all of their current high sulfur diesel fuel as heating oil, due 
to their location relative to major pipelines and the size of the 
heating oil market in their area. Those not located in major heating 
oil markets and not connected to pipelines serving these areas were 
projected to have to meet the 500 ppm cap in 2007.
    Four, we determined the additional refineries which would produce 
500 ppm and 15 ppm fuel to satisfy demand during each phase of the fuel 
program. Refineries projected to have the lowest compliance costs in 
each PADD were projected to produce the lower sulfur fuels until demand 
was met. PADD 3 refineries were allowed to ship low sulfur fuel to the 
Northeast, but no other inter-PADD transfers were assumed. Imports of 
500 ppm highway diesel fuel were assumed to increase at the rate of 
highway diesel fuel consumption and be converted to 15 ppm diesel fuel, 
80 percent in 2006 and 100 percent in 2010. Imports of high sulfur 
distillate fuel were assumed to increase at the rate of high sulfur 
distillate fuel consumption, but were assumed to remain entirely high 
sulfur heating oil even after today's NRLM fuel proposal. In other 
words, all 15 ppm and 500 ppm NRLM fuel produced under this proposal 
was assumed to be produced by domestic refineries. This assumption 
increased the projected costs of the proposal described above more than 
would have been the case had we assumed that domestic production and 
imports of high sulfur distillate fuel would each keep their respective 
shares of the NRLM diesel fuel and heating oil markets in response to 
this proposal. The relative costs of producing 15 ppm nonroad diesel 
fuel by domestic and overseas refiners is discussed further in section 
V.A.6. below.
    With the onset of a 2007 500 ppm sulfur cap for nonroad, locomotive 
and marine diesel fuel, we project that the market for high sulfur 
diesel fuel and heating oil would become so small that high sulfur fuel 
would no longer be shipped through common carrier pipelines in most 
areas. The prime exception to this would be the Northeast, where the 
heating oil market is very large. Thus, refiners located in the 
Northeast and those along the major pipelines serving the Northeast, 
namely the Colonial and Plantation pipelines, could continue to produce 
high sulfur

[[Page 28433]]

heating oil. Other refineries would shift the production of high sulfur 
diesel fuel and heating oil to the 500 ppm NRLM market. The second 
exception would be refiners granted special provisions due to the small 
size of their business (i.e., SBREFA refiners) or economic hardship, as 
discussed in section IV above. The high sulfur distillate production 
levels of these refineries is small enough that they can sell into more 
local nonroad, locomotive and marine markets or the heating oil market 
without using pipelines and so they could continue to produce high 
sulfur distillate.
    Based on refinery distillate production data from the Energy 
Information Administration (EIA), there are 122 refineries currently 
producing highway diesel fuel and 105 refineries producing high sulfur 
diesel fuel or heating oil. Using the methodology described above, 
absent this proposal, we project that roughly 114 refineries will 
invest in additional desulfurization equipment to produce 15 ppm 
highway diesel fuel; 74 refineries in 2006 and 40 in 2010.\273\ These 
114 refineries include 109 of the 122 refineries which currently 
produce highway diesel fuel, plus 5 refineries which currently only 
produce high sulfur distillate fuel today. Again absent the proposed 
NRLM diesel fuel program, we project that roughly 13 refineries 
currently producing highway diesel fuel will shift to producing high 
sulfur distillate fuel. This would leave a total of 113 refineries 
still producing high sulfur distillate after full implementation of the 
2007 highway diesel fuel program.
---------------------------------------------------------------------------

    \273\ These (and the subsequent) estimates of the number of 
refineries investing in new equipment to produce diesel fuels of 
various sulfur levels should be understood as rough estimates which 
assist us in projecting costs and other impacts related to this 
proposal. They are most reasonable when evaluating the total number 
of refineries investing in a particular year or region. We are not 
indicating that we believe that we can predict which specific 
refineries would invest in desulfurization equipment in response to 
this proposal.
---------------------------------------------------------------------------

    The number of these 113 domestic refineries expected to produce 
either 500 ppm of 15 ppm NRLM diesel fuel in response to this proposal 
is summarized in Table V-A-2.

                Table V-A-2 Refineries Projected to Produce NRLM Diesel Fuel Under This Proposal
----------------------------------------------------------------------------------------------------------------
                                                        500 ppm diesel fuel             15 ppm diesel fuel
                                                 ---------------------------------------------------------------
                 Year of Program                                       Small                           Small
                                                  All refineries    refineries    All refineries    refineries
----------------------------------------------------------------------------------------------------------------
2007-2010.......................................              42               0               0               0
2010-2014.......................................              37              19              25               0
2014+...........................................              25              12              37               7
----------------------------------------------------------------------------------------------------------------

    As shown in this table, we project that 42 of the 113 refineries 
currently producing some high sulfur distillate would desulfurize their 
high sulfur diesel fuel in response to the proposed 500 ppm standard in 
2007. The remainder would continue producing either high sulfur NRLM 
diesel fuel under the proposed small refiner provisions, or high sulfur 
heating oil. As explained in section IV.F, we project that these 
refiners would use conventional hydrotreating technology to meet this 
standard. Of these 42 refineries, we project that 32 would build new 
hydrotreaters to meet the 500 ppm sulfur cap. We project that three of 
the remaining ten refineries would be able to meet the 500 ppm cap with 
their existing hydrotreater which is currently being used to produce 
highway diesel fuel. These three refineries are projected to build a 
new hydrotreater to produce 15 ppm highway diesel fuel in 2006, so 
their existing highway fuel hydrotreater could process their current 
high sulfur diesel fuel. The remaining seven refineries currently 
produce relatively small amounts of high sulfur diesel fuel compared to 
their highway diesel fuel production. We project that these refiners 
would be able to economically revamp their existing highway 
hydrotreater to process their non-highway diesel fuel.
    We project that the capital cost involved to meet the 2007 500 ppm 
sulfur cap would be $600 million, or $9.7 million per refinery building 
a new hydrotreater. The bulk of this capital would be invested in 2007 
($500 million), with the remainder being invested in 2010.\274\ 
Operating costs would be about $3 million per year for the average 
refinery. We request comment on the number of refiners who would need 
to build new equipment to meet the 500 ppm sulfur cap, the capital cost 
for this new equipment and the cost of operating this equipment.
---------------------------------------------------------------------------

    \274\ Some refineries would be able to delay production of 500 
ppm NRLM fuel until 2010 due to the proposed small refiner 
provisions. Likewise, some refineries would be able to delay 
production of 15 ppm nonroad diesel fuel until 2014.
---------------------------------------------------------------------------

    Starting in mid-2010, we project that 25 refineries would add or 
revamp equipment to meet the 15 ppm cap on nonroad diesel fuel, while 
20 refineries (nearly all of them small refiners) would add or revamp 
equipment to produce 500 ppm nonroad or locomotive and marine diesel 
fuel. Finally, an additional 12 refineries (again nearly all of them 
small refiners) would begin producing 15 ppm nonroad diesel fuel in 
2014.
    We project that 80 percent of the 15 ppm nonroad diesel fuel volume 
would be desulfurized by advanced technologies, while the remaining 20 
percent would be desulfurized by conventional hydrotreaters. Since the 
bulk of the hydrotreating capacity being used to meet the 2007 500 ppm 
standard for NRLM diesel fuel would have just been built in 2007 or 
2010, we expect that it would have been designed to facilitate further 
processing to 15 ppm sulfur and the added 15 ppm facilities would be 
revamps. However, those refiners who used their existing highway diesel 
fuel hydrotreaters to meet the proposed 500 ppm cap in 2007 would 
likely have to construct new equipment in 2010 or 2014 to meet the 15 
ppm cap on nonroad diesel fuel, since these hydrotreaters could not be 
revamped in 2006 to produce 15 ppm highway diesel fuel. When the 
proposed NRLM diesel fuel program would be fully implemented in 2014, 
roughly 51 refineries are still projected to produce high sulfur 
heating oil and thus, would not face any refining costs related to this 
proposal.
    Our projection that 80 percent of refineries would utilize some 
form of advanced technology to meet the proposed 15 ppm nonroad fuel 
sulfur cap is based on the fact that this 15 ppm cap would follow the 
production of 15 ppm highway diesel fuel by four years. Several firms 
are expending significant research and development resources to bring 
such advanced technologies to the market for the highway diesel fuel

[[Page 28434]]

program. We developed cost estimates for two such technologies: Linde 
Iso-Therming and Phillips S-Zorb. The development of cost estimates for 
these two advanced technologies, as well as conventional hydrotreating, 
is described in detail in Chapter 7 of the Draft RIA. We request 
comment on the potential viability and cost savings associated with 
advanced desulfurization technologies, particularly in the 2010 
timeframe.
    The total capital cost of new equipment and revamps related to the 
proposed 2010 sulfur standard would be $640 million, or $17 million per 
refinery adding or revamping equipment. Total operating costs would be 
about $5 million per year for the average refinery. The total refining 
cost, including the amortized cost of capital, would be 4.4 cents per 
gallon of new 15 ppm nonroad fuel. This cost is relative to the cost of 
producing high sulfur fuel today, and includes the cost of meeting the 
500 ppm standard beginning in 2007. We request comment on the number of 
refiners who would need to build new equipment to meet the 15 ppm 
sulfur cap, the capital cost for this new equipment and the cost of 
operating this equipment. The average cost of continuing to meet the 
500 ppm standard for locomotive and marine fuel would continue at 2.2 
cents per gallon.
    The above costs reflect national averages for the fully phased in 
program for each control step. Some refiners would face lower costs 
while others would face higher costs. Excluding small refiners because 
they are able to take advantage of the proposed small refiner 
provisions, the average refining costs by refining region are shown in 
the table below. Combined costs are shown for PADDs 1 and 3 because of 
the large volume of diesel fuel which is shipped from PADD 3 to PADD 1.

    Table V-A-3.--Average Refining Costs by Region (cents per gallon)
------------------------------------------------------------------------
                                2007 500 ppm Cap       2010 15 ppm Cap
------------------------------------------------------------------------
PADDs 1 and 3...............                   1.4                   2.6
PADD 2......................                   2.9                   5.7
PADD 4......................                   4.0                   8.5
PADD 5......................                   2.6                   5.4
Nationwide..................                   2.2                   4.4
------------------------------------------------------------------------

    We request comment on the range of estimated refining costs for the 
various regions for both the proposed 500 and 15 ppm sulfur caps.
2. Cost of Lubricity Additives
    Hydrotreating diesel fuel tends to reduce the natural lubricating 
quality of diesel fuel, which is necessary for the proper functioning 
of certain fuel system components. There are a variety of fuel 
additives which can be used to restore diesel fuel's lubricating 
quality. These additives are currently used to some extent in highway 
diesel fuel. We expect that the need for lubricity additives that would 
result from the proposed 500 ppm sulfur standard for off-highway diesel 
engine fuel would be similar to that for highway diesel fuel meeting 
the current 500 ppm sulfur cap standard.\275\ Industry experience 
indicates that the vast majority of highway diesel fuel meeting the 
current 500 ppm sulfur cap does not need lubricity additives. 
Therefore, we expect that the great majority of off-highway diesel 
engine fuel meeting the proposed 500 ppm sulfur standard would also not 
need lubricity additives. In estimating lubricity additive costs for 
500 ppm diesel fuel, we assumed that fuel suppliers would use the same 
additives at the same concentration as we projected would be used in 15 
ppm highway diesel fuel. Based on our analysis of this issue for the 
2007 highway diesel fuel program, the cost per gallon of the lubricity 
additive is about 0.2 cent. This level of use is likely conservative, 
as the amount of lubricity additive needed increases substantially as 
diesel fuel is desulfurized to lower levels. We also project that only 
5 percent of all 500 ppm NRLM diesel fuel would require the use of a 
lubricity additive. Thus, we project that the cost of additional 
lubricity additives for the affected 500 ppm NRLM diesel fuel would be 
0.01 cent per gallon. See the Draft RIA for more details on the issue 
of lubricity additives.
---------------------------------------------------------------------------

    \275\ Please refer to section IV in today's preamble for 
additional discussion regarding our projections of the potential 
impact on fuel lubricity of this proposed rule.
---------------------------------------------------------------------------

    We project that all nonroad diesel fuel meeting a 15 ppm cap would 
require treatment with lubricity additives. Thus, the projected cost 
would be 0.2 cent per affected gallon of 15 ppm nonroad diesel fuel.
3. Distribution Costs
    The proposed fuel program is projected to impact distribution costs 
in three ways. One, we project that more diesel fuel would have to be 
distributed under the proposal than without it. This is due to the fact 
that some of the desulfurization processes reduce the fuel's volumetric 
energy density during processing. Total energy is not lost during 
processing, as the total volume of fuel is increased. However, a 
greater volume of fuel must be consumed in the engine to produce the 
same amount of power. We assumed that the current cost of distributing 
diesel fuel of 10 cents per gallon (see Draft RIA for further details) 
would stay constant (i.e., a 1 percent increase in the amount of fuel 
distributed would increase total distribution costs by 1 percent).
    We project that desulfurizing diesel fuel to 500 ppm would reduce 
volumetric energy content by 0.7 percent. This would increase the cost 
of distributing fuel by 0.07 cent per gallon. We project that 
desulfurizing diesel fuel to 15 ppm would reduce volumetric energy 
content by an additional 0.35 percent. This would increase the cost of 
distributing fuel by an additional 0.04 cent per gallon, or a total 
cost of 0.11 cent per gallon of affected 15 ppm nonroad diesel fuel.
    Two, while this proposal minimizes the segregation of similar 
fuels, some additional segregation of products in the distribution 
system would still be required. The proposed allowance that highway and 
off-highway diesel engine fuel meeting the same sulfur specification 
can be shipped fungibly until it leaves the terminal obviates the need 
for additional storage tankage in this segment of the distribution 
system.\276\ This proposal would also allow 500 ppm NRLM diesel fuel to 
be mixed with high-sulfur NRLM diesel fuel once the fuels are dyed to 
meet IRS requirements. This provision would ease the last part of the 
distribution of high-sulfur NRLM diesel fuel.
---------------------------------------------------------------------------

    \276\ Including the refinery, pipeline, marine tanker, and barge 
segments of the distribution system.
---------------------------------------------------------------------------

    However, we expect that the implementation of the proposed 500 ppm 
standard for NRLM diesel fuel in 2007 would compel some bulk plants in 
those parts of the country still

[[Page 28435]]

distributing heating oil as a separate fuel grade to install a second 
diesel storage tank to handle this 500 ppm nonroad fuel. These bulk 
plants currently handle only high-sulfur fuel and hence would need a 
second tank to continue their current practice of selling fuel into the 
heating oil market in the winter and into the nonroad market in the 
summer.\277\ We believe that some of these bulk plants would convert 
their existing diesel tank to 500 ppm fuel in order to avoid the 
expense of installing an additional tank. However, to provide a 
conservatively high estimate we assumed that 10 percent of the 
approximately 10,000 bulk plants in the U.S. (1,000) would install a 
second tank in order to handle both 500 ppm NRLM diesel fuel and 
heating oil. The cost of an additional storage tank at a bulk plant is 
estimated at $90,000 and the cost of de-manifolding their delivery 
truck at $10,000.\278\ If all 1,000 bulk plants were to install a new 
tank, the total one-time capitol cost would be $100,000,000. Amortizing 
the capital costs over 20 years, results in a estimated cost for 
tankage at such bulk plants of 0.1 cent per gallon of affected NRLM 
diesel fuel supplied. Although the impact on the overall cost of the 
proposed program is small, the cost to those bulk plant operators who 
need to put in a separate storage tank may represent a substantial 
investment. Thus, as discussed in section IV.F., we believe many of 
these bulk plants could make other arrangements to continue servicing 
both heating oil and NRLM markets.
---------------------------------------------------------------------------

    \277\ See section IV.E.9. of this proposal and chapter 5 of the 
RIA for additional discussion of the potential impacts of the 
proposed sulfur standards on the distribution system.
    \278\ This estimated cost includes the addition of a separate 
delivery system on the tank truck.
---------------------------------------------------------------------------

    Due to the end of the highway program temporary compliance option 
(TCO) in 2010 and the disappearance of high-sulfur diesel fuel from 
much of the fuel distribution system due to the implementation of this 
proposed rule, we expect that storage tanks at many bulk plants which 
were previously devoted to 500 ppm TCO highway fuel and high-sulfur 
fuel would become available for dyed 15 ppm nonroad diesel service. 
Based on this assessment, we do not expect that a significant number of 
bulk plants would need to install an additional storage tank in order 
to provide dyed and undyed 15 ppm diesel fuel to their customers 
beginning in 2010 (the proposed implementation date for the 15 ppm 
nonroad standard).\279\ There could potentially be some additional 
costs related to the need for new tankage in some areas not already 
carrying 500 ppm fuel under the temporary compliance option of the 
highway diesel program and which continue to carry high sulfur fuel. 
However, we expect them to minimal relative to the above 0.1 cent per 
gallon cost. Thus, we estimate that the total cost of additional 
storage tanks that would result from the adoption of this proposal 
would be 0.1 cent per gallon of affected off-highway diesel engine fuel 
supplied.
---------------------------------------------------------------------------

    \279\ See section IV of today's preamble for additional 
discussion of our rational for this conclusion.
---------------------------------------------------------------------------

    Three, the proposed requirement that high sulfur heating oil be 
marked between 2007 and 2010 and that locomotive and marine diesel fuel 
be marked from 2010 until 2014 would increase the cost of distributing 
these fuels slightly. Based on input from marker manufacturers, we 
estimate that marking these fuels would cost no more than 0.2 cent per 
gallon and could cost considerably less. There should be no capital 
cost associated with this requirement, as we are proposing to remove 
the current requirement that refiners dye all high sulfur distillate at 
the refinery. The current dyeing equipment should work equally well for 
the marker. Because heating oil is being marked to prevent its use in 
NRLM engines, we have spread the cost for this marker over NRLM diesel 
fuel. Thus, from a regulatory point of view, the heating oil marker 
would increase the cost of NRLM diesel fuel between 2007 and 2010 by 
0.16 cent per gallon. We attribute the cost of marking 500 ppm 
locomotive and marine diesel fuel directly to this fuel, so the marker 
cost is simply 0.2 cent per gallon of locomotive and marine diesel fuel 
between 2010 and 2014.
    We do not project any additional downgrade of 15 ppm diesel fuel 
would result from the proposed fuel program. In our analysis of the 15 
ppm highway fuel program, we also projected additional distribution 
costs due to the need to downgrade more volume of highway diesel fuel 
to a lower value product. This is a consequence of the large difference 
between the sulfur content of 15 ppm fuel and other distillate 
products, like high sulfur diesel fuel, heating oil and jet fuel.\280\ 
We do not project that these costs would increase with this proposed 
rule. Highway diesel fuel meeting a 15 ppm cap will already be being 
distributed in all major pipeline and terminal networks. Thus, we 
expect that 15 ppm nonroad fuel would be added to batches of 15 ppm 
already being distributed. In this situation, the total interface 
volume needing to be downgraded would not increase. At the same time, 
we are not projecting that interface volume would decrease, as high 
sulfur fuels, such as jet fuel, would still be in the system.
---------------------------------------------------------------------------

    \280\ Off-highway diesel fuel sulfur content is currently 
unregulated and is approximately 3,400 ppm on average. The maximum 
allowed sulfur content of heating oil is 5,000 ppm. The maximum 
allowed sulfur content of kerosene (and jet fuel) is 3,000 ppm.
---------------------------------------------------------------------------

    Thus, overall, we estimate that the total additional distribution 
would be 0.3 cent per gallon of nonroad, locomotive and marine fuel 
during the first step of the proposed program (from 2007 through 2010). 
We project that distribution costs would increase to 0.4 cent gallon 
for 500 ppm locomotive and marine diesel fuel from 2010 to 2014, but 
decrease to 0.2 cent per gallon thereafter. Finally, we project that 
distribution costs for 15 ppm nonroad diesel fuel would be 0.2 cent 
gallon.
4. How EPA's Projected Costs Compare to Other Available Estimates
    We used two different methods for evaluating how well our cost 
estimates reflect the true costs for complying with the two step 
nonroad fuel program. The first method compared our costs with the 
incremental market price of diesel fuel meeting a 15 or 500 ppm 
standard. The second method compared our cost estimate to that from an 
engineering analysis analogous to the one we performed.
    Beginning with market prices, highway diesel fuel meeting a 500 ppm 
sulfur cap has been marketed in the U.S. for almost ten years. Over the 
five year period from 1995-1999, its national average price has 
exceeded that of high sulfur diesel fuel by about 2.4 cent per gallon 
(see chapter 7 of the Draft RIA). While fuel prices are a often a 
function of market forces which might not reflect the cost of producing 
the fuel, the comparison of the price difference over a fairly long 
period such as 5 years would tend to reduce the effect of the market on 
the prices and more closely reflect the cost of complying with the 500 
ppm cap standard. Thus, we feel that this is a sound basis for 
evaluating our cost estimate. This price difference is essentially the 
same as our estimated cost for refining and distributing 500 ppm non-
highway diesel fuel, thus the price difference for producing and 
distributing 500 ppm highway fuel corroborates our cost analysis.
    Some 15 ppm diesel fuel is marketed today. However, it is either 
being produced in very limited quantities using equipment designed to 
meet less

[[Page 28436]]

stringent sulfur standards or with other properties which make it 
unrepresentative of typical U.S. NRLM diesel fuel. Thus, current market 
prices are not a good indication of the long term price impact of the 
proposed 15 ppm cap.
    Regarding engineering studies, the Engine Manufactures Association 
(EMA) commissioned a study by Mathpro to estimate the cost of 
controlling the sulfur content of highway and nonroad diesel fuel to 
levels consistent with both 500 ppm and 15 ppm cap standards.\281\ 
Mathpro used a higher rate of return on new capital so we adjusted 
their per-gallon costs to reflect our own amortization methodology. 
Also, the Mathpro study was completed in 1999 so we adjusted their 
costs for inflation to year 2002 dollars. After these two adjustments, 
Mathpro's cost to desulfurize the high sulfur non-highway pool to 500 
ppm is 2.5 cents per gallon, while that for a 15 ppm cap is 5.8 cents 
per gallon.\282\ The 500 ppm cost estimate compares quite favorably 
with our own estimate of 2.2 cents per gallon cost. One reason for our 
somewhat lower estimate for complying with the 500 ppm standard is that 
our refinery-specific analysis has only the lowest cost refineries 
complying as many more expensive refineries can continue to produce 
heating oil. It is likely that the refineries which our analysis show 
would comply are more optimized for desulfurizating diesel fuel than 
the average refinery used by Mathpro. This reason applies even more for 
15 ppm cap standard as fewer, more optimized refineries need to comply 
to produce nonroad diesel fuel which complies with a 15 ppm sulfur cap 
standard. Furthermore, we considered the use of advanced 
desulfurization technologies for complying with the 15 ppm standard, 
while Mathpro did not. Since the Mathpro study was performed in 1999, 
cost estimates were not available for either of the two technologies 
which we included. The adjustment of the Mathpro costs and the 
comparison with our own cost estimates are discussed in detail in the 
Draft RIA. We request comment on the degree that the results of the 
Mathpro study for EMA and the comparison with real-world prices support 
our own cost estimates.
---------------------------------------------------------------------------

    \281\ Hirshfeld, David, MathPro, Inc., ``Refining economics of 
diesel fuel sulfur standards,'' performed for the Engine 
Manufactuers Association, October 5, 1999.
    \282\ The Mathpro costs cited reflect their case where current 
diesel fuel hydrotreaters are revamped with a new reactor in series, 
which is the most consistent with our technology projection.
---------------------------------------------------------------------------

5. Supply of Nonroad, Locomotive and Marine Diesel Fuel
    EPA has developed the proposed fuel program to minimize its impact 
on the supply of distillate fuel. For example: we have proposed to 
transition the fuel sulfur level down to 15 ppm in two steps, providing 
an estimated 6 years of leadtime for the final step; we are proposing 
to provide flexibility to refiners through the availability of banking 
and trading provisions; and we have provided relief for small refiners 
and hardship relief for any qualifying refiner. In order to evaluate 
the effect of this proposal on supply, EPA evaluated four possible 
cases: (1) whether the proposed standards could cause refiners to 
remove certain blendstocks from the fuel pool, (2) whether the proposed 
standards could require chemical processing which loses fuel in the 
process, (3) whether the cost of meeting the proposed standards could 
lead some refiners to leave that market, and (4) whether the cost of 
meeting the proposed standards could lead some refiners to stop 
operations altogether (i.e., shut down). In all cases, as discussed 
below, we have concluded that the answer is no. Therefore, consistent 
with our findings made during the 2007 highway diesel rule, we do not 
expect this proposed rule to cause any supply shortages of nonroad, 
locomotive and marine diesel fuel. The reader is referred to the draft 
RIA for a more detailed discussion of the potential supply impact of 
this proposed rule.
    Blendstock Shift: There should be no long term reduction in the 
amount of material derived from crude oil available for blending into 
diesel fuel or heating oil as a result of this proposal. Technology 
exists to desulfurize any commercial diesel fuel to less than 10 ppm 
sulfur. This technology is just now being proven on a commercial scale 
with a range of no. 2 diesel fuel blendstocks, as a number of refiners 
are producing 15 ppm fuel for diesel fleets which have been retro-
fitted with PM traps or for pipeline testing. Therefore, there is no 
technical necessity to remove certain blendstocks from the diesel fuel 
pool. It costs more to process certain blendstocks, such as light cycle 
oil, than others. Therefore, there may be economic incentives to move 
certain blendstocks out of the diesel fuel market to reduce compliance 
costs. However, that is an economic issue, not a technical issue and 
will be addressed below when we consider whether refiners might choose 
to exit the NRLM diesel fuel market.
    Processing Losses: The impact of the proposed rule on the total 
output of liquid fuel from refineries would be negligible. Conventional 
desulfurization processes do not reduce the energy content of the input 
material. However, the form of the material is affected slightly. With 
conventional hydrotreating, about 98 percent of the diesel fuel fed to 
a hydrotreater producing 15 ppm sulfur product leaves as diesel fuel. 
Of the 2 percent loss, three-fourths, or about 1.5 percent leaves the 
unit as naphtha (i.e., gasoline feedstock). The remainder is split 
evenly between liquified petroleum gas (LPG) and refinery fuel gas. 
Both naphtha and LPG have higher valuable uses as liquid fuels. Naphtha 
can be used to produce gasoline. Refiners can adjust the relative 
amounts of gasoline and diesel fuel which they produce, especially to 
this small degree. This additional naphtha can displace other gasoline 
blendstocks, which can then be shifted to the diesel fuel pool. LPG, on 
the other hand, is primarily used in heating, where it competes with 
heating oil. Thus, additional LPG can be used to displace gasoline and 
heating oil, which in turn can be shifted to the diesel fuel pool. 
Thus, there should be little or no direct impact of desulfurization on 
refinery fuel production. The shift from diesel fuel to fuel gas is 
very small (0.25 percent) and this fuel gas can be used to reduce 
consumption of natural gas within the refinery. These figures apply to 
the full effect of the proposed standards (i.e., the reduction in 
sulfur content from 3400 ppm to 15 ppm). For the first step of the 
proposed fuel program and that portion of the diesel fuel pool which 
would remain at the 500 ppm level indefinitely, the impacts would only 
be about 40 percent of those described above.
    The use of advanced desulfurization technologies would further 
reduce these impacts. These technologies are projected to be used in 
the second step of reducing 500 ppm diesel fuel to 15 ppm sulfur. We 
project that the Linde process would reduce the above losses for the 
second step by 55 percent, while the Phillips SZorb process would have 
no loss in diesel fuel production.
    Exit the NRLM Diesel Fuel Market: While the cost of meeting the 
proposed standards might cause some individual refiners to consider 
reducing their production of NRLM fuel or leave the market entirely, we 
do not believe that across the entire industry such a shift is possible 
or likely. As mentioned above, all diesel fuels and heating oil are 
essentially identical both chemically and physically, except for sulfur 
level. Thus, if a refiner could shift his high

[[Page 28437]]

sulfur distillate material from the nonroad, locomotive and marine 
diesel fuel markets to the heating oil market starting in mid-2007, it 
would avoid the need to invest in new desulfurization equipment. 
Likewise, starting in mid-2010, a refiner could focus his 500 ppm 
diesel fuel in the locomotive and marine diesel fuel markets or shift 
this material to the heating oil market. The problem would be a 
potential oversupply of heating oil starting in 2007 and locomotive and 
marine diesel fuel and heating oil starting in 2010. An oversupply 
could lead to a substantial drop in market price, significantly 
increasing the cost of leaving the nonroad, locomotive and marine 
diesel fuel markets. Or, it may be necessary to export the higher 
sulfur fuel in order to sell it. This could entail transportation costs 
and overseas prices no higher than existed in the U.S. before the 
oversupply (and possibly lower due to these imports now entering these 
overseas markets).
    We addressed this same issue during the development of 2007 highway 
diesel fuel program. There, the issue was whether refiners would shift 
some or all of their current highway diesel fuel production to either 
domestic or overseas markets for high sulfur diesel fuel or heating oil 
in order to avoid investing to meet the 15 ppm cap for highway diesel 
fuel. A study by Charles River Associates, et al., sponsored by API 
projected that there could be a near-term shortfall in highway diesel 
fuel supply of as much as 12 percent.\283\ However, supported by a 
study by Muse, Stancil, we concluded that refiners would incur greater 
economic loss in trying to avoid meeting the 15 ppm highway diesel fuel 
cap than they would by complying at current production levels even if 
the market did not allow them to recover their capital investment. A 
study by Mathpro, Inc. for AAM and EMA also criticized the conclusions 
of the Charles River study, particularly their assumption that 
compliance costs alone would drive investment decisions and that there 
was essentially a single highway diesel fuel market nationwide.\284\ 
Mathpro demonstrated that smaller refineries located, for example, in 
the Rocky Mountain region, likely faced higher per gallon compliance 
costs, but also had been more profitable over the past 15 years than 
larger refiners in other areas with lower overall costs. This was due 
to their market niches and the inability for lower cost refiners to 
ship large volumes of fuel economically to their market.
---------------------------------------------------------------------------

    \283\ ``An Assessment of the Potential Impacts of Proposed 
Environmental Regulations on U.S. Refinery Supply of Diesel Fuel,'' 
Charles River Associates and Baker and O'Brien, for API, August 
2000.
    \284\ ``Prospects for Adequate Supply of Ultra Low Sulfur Diesel 
Fuel in the Transition Period (2006-2007), An Analysis of Technical 
and Economic Driving Forces for Investment in ULSD Capacity in the 
U.S. Refining Sector,'' MathPro, Inc., for AAM and EMA, December 7, 
2001.
---------------------------------------------------------------------------

    We believe that the same conclusions apply to the proposed fuel 
program for six reasons. One, the alternative markets for high sulfur 
diesel fuel and heating oil would be even more limited after the 
proposed sulfur caps on nonroad, locomotive and marine diesel fuel than 
they will be in 2006, as half of the current U.S. market for high 
sulfur, no. 2 distillate would disappear. We expect that high sulfur 
heating oil would not even by carried be common carrier pipelines 
except those serving the Northeast. Therefore, refiners' sale of high 
sulfur distillate may be limited to markets serviceable by truck. Two, 
the desulfurization technology to meet a 500 ppm cap has been 
commercially demonstrated for over a decade. The desulfurization 
technology to meet a 15 ppm cap will have been commercially 
demonstrated in mid-2006, a full four years prior to the implementation 
of the 15 ppm cap on nonroad diesel fuel. Three, the volume of fuel 
affected by the 15 ppm nonroad diesel fuel standard would be only one-
seventh of that affected by the highway diesel fuel program. This 
dramatically reduces the required capital investment. Four, both Europe 
and Japan are implementing sulfur caps for highway and nonroad diesel 
fuel in the range of 10-15 ppm, eliminating these markets as a sink for 
high sulfur diesel fuel. Five, refineries outside of the U.S. and 
Europe are operating at a lower percentage of their capacity than U.S. 
refineries. Thus, U.S. refineries would not be able to obtain 
attractive prices for high sulfur diesel fuel overseas. Finally, 
refinery profit margins were much higher during the last part of 2000 
and most of 2001 than over the past ten years, indicating a potential 
long-term improvement in profitability. Margins decreased again during 
most in 2002, but recovered during the last few months of that year and 
in early 2003.
    Once refiners have made their investments to meet the proposed NRLM 
diesel fuel standards, or have decided to produce high sulfur heating 
oil, we expect that the various distillate markets would operate very 
similar to today's markets. When fully implemented in 2014, there will 
be three distillate fuels in the market, 15 ppm highway and nonroad 
diesel fuel, 500 ppm locomotive and marine diesel fuel and high sulfur 
heating oil. The market for 500 ppm locomotive and marine diesel fuel 
is much smaller than the other two, particularly considering that it is 
nationwide and the heating oil market is geographically concentrated. 
Therefore, the vast majority of refiners are expected to focus on 
producing either 15 ppm or high sulfur distillate, which is similar to 
today, where there are two fuels, 500 ppm and high sulfur distillate. 
In this case, refiners with the capability of producing 15 ppm diesel 
fuel have the most flexibility, since they can sell their fuel to any 
of the three markets. Refiners with only 500 ppm desulfurization 
capability can supply two markets. Those refiners only capable of 
producing high sulfur distillate would not be able to participate in 
either the 15 or 500 ppm markets. However, this is not different from 
today. Generally, we do not expect one market to provide vastly 
different profit margins than the others, as high profit margins in one 
market will attract refiners from another via investment in 
desulfurization equipment.
    Refinery Closure: There are a number of reasons why we do not 
believe that refineries would completely close down under this proposed 
rule. One reason is that we have included provisions to provide relief 
for small refiners, as well as any refiner facing unusual financial 
hardship. Another reason is that nonroad, locomotive and marine diesel 
fuel is usually the third or fourth most important product produced by 
the refinery from a financial perspective. A total shutdown would mean 
losing all the revenue and profit from these other products. Gasoline 
is usually the most important product, followed by highway diesel fuel 
and jet fuel. A few refineries do not produce either gasoline or 
highway diesel fuel, so jet fuel and high sulfur diesel fuel and 
heating oil are their most important products. The few refiners in this 
category likely face the biggest financial challenge in meeting the 
proposed requirements. However, those refiners would also presumably be 
in the best position to apply for special hardship provisions, 
presuming that they do not have readily available source of investment 
capital. The additional time afforded by these provisions should allow 
the refiner to generate sufficient cash flow to invest in the required 
desulfurization equipment. Investment here could also provide them the 
opportunity to expand into more profitable (e.g., highway diesel) 
markets.
    A quantitative evaluation of whether the cost of the proposed fuel 
program could cause some refineries to cease operations completely 
would be very difficult, if not impossible to perform. A

[[Page 28438]]

major factor in any decision to shut down is the refiner's current 
financial situation. It is very difficult to assess an individual 
refinery's current financial situation. This includes a refiner's debt, 
as well as its profitability in producing fuels other than those 
affected by a particular regulation. It can also include the 
profitability of other operations and businesses owned by the refiner.
    Such an intensive analysis can be done to some degree in the 
context of an application for special hardship provisions, as discussed 
above. However, in this case, EPA can request detailed financial 
documents not normally available. Prior to such application, as is the 
case now, this financial information is usually confidential. Even when 
it is published, the data usually apply to more than just the operation 
of a single refinery.
    Another factor is the need for capital investments other than for 
this proposed rule. EPA can roughly project the capital needed to meet 
other new fuel quality specifications, such as the Tier 2 or highway 
diesel sulfur standards. However, we cannot predict investments to meet 
local environmental and safety regulations, nor other investments 
needed to compete economically with other refiners.
    Finally, any decision to close in the future must be based on some 
assumption of future fuel prices. Fuel prices are very difficult to 
project in absolute terms. The response of prices to changes in fuel 
quality specifications, such as sulfur content, as is discussed in the 
next section, are also very difficult to predict. Thus, even if we had 
complete knowledge of a refiner's financial status and its need for 
future investments, the decision to stay in business or close would 
still depend on future earnings, which are highly dependent on the 
prices of all products produced by that refinery.
    Some studies in this area point to fuel pricing over the past 15 
years or so and conclude that prices will only increase to reflect 
increased operating costs and will not reflect the cost of capital. In 
fact, the rate of return on refining assets has been poor over the past 
15 years and until recently, there has been a steady decline in the 
number of refineries operating in the U.S. However, this may have been 
due to a couple of circumstances specific to that time period. One, 
refinery capacity utilization was less than 80 percent in 1985. Two, at 
least regarding gasoline, the oxygen mandate for reformulated gasoline 
caused an increase in gasoline supply despite low refinery utilization 
rates. While this led to healthy financial returns for oxygenate 
production, it did not help refining profit margins.
    Today, refinery capacity utilization in the U.S. is generally 
considered to be at its maximum sustainable rate. There are no 
regulatory mandates on the horizon which will increase production 
capacity significantly, even if ethanol use in gasoline increases 
substantially.\285\ Consistent with this, refining margins have been 
much better over the past two and a half years than during the previous 
15 years and the refining industry itself is projecting good returns 
for the foreseeable future.
---------------------------------------------------------------------------

    \285\ Both houses of the U.S. Congress are considering bills 
which would require the increased use of renewables, like ethanol, 
in gasoline and diesel fuel. While the amount of renewables could be 
considerable, it is well below the annual growth in transportation 
fuel use.
---------------------------------------------------------------------------

6. Fuel Prices
    It is well known that it is difficult to predict fuel prices in 
absolute terms with any accuracy. The price of crude oil dominates the 
cost of producing gasoline and diesel fuel. Crude oil prices have 
varied by more than a factor of two in the past year. In addition, 
unexpectedly warm or cold winters can significantly affect heating oil 
consumption, which affects the amount of gasoline produced and the 
amount of distillate material available for diesel fuel production. 
Economic growth, or its lack, affects fuel demand, particularly for 
diesel fuel. Finally, both planned and unplanned shutdowns of 
refineries for maintenance and repairs can significantly affect total 
fuel production, inventory levels and resulting fuel prices.
    Predicting the impact of any individual factor on fuel price is 
also difficult. The overall volatility in fuel prices limits the 
ability to determine the effect of a factor which changed at a specific 
point in time which might have led to the price change, as other 
factors continue to change over time. Occasionally, a fuel quality 
change, such as reformulated gasoline or a 500 ppm cap on diesel fuel 
sulfur content, only affects a portion of the fuel pool. In this case, 
an indication of the impact on price can be inferred by comparing the 
prices of the two fuels at the same general location over time. 
However, this is still only possible after the fact, and cannot be done 
before the fuel quality change takes place.
    Because of these difficulties, EPA has generally not attempted to 
project the impact of its rules on fuel prices. However, in response to 
Executive Order 13211, we are doing so for this proposed rule. To 
reflect the inherent uncertainty in making such projections, we 
developed three projections for the potential impact of the proposed 
fuel program on fuel prices. The range of potential long-term price 
increases are shown in Table V-A-4. Short-term price impacts are highly 
volatile, as are short-term swings in absolute fuel prices, and much 
too dependent on individual refiners' decisions, unexpected shutdowns, 
etc. to be predicted even with broad ranges.

             Table V-A-4.--Range of Possible Total Diesel Fuel Price Increases (cents per gallon) a
----------------------------------------------------------------------------------------------------------------
                                                                    Lower Limit      Mid-Point        Maximum
----------------------------------------------------------------------------------------------------------------
               2007 500 ppm Sulfur Cap: Nonroad, Locomotive and Marine Diesel Fuel
-------------------------------------------------------------------------------------------------
PADDs 1 and 3...................................................             0.9             1.5             3.4
PADD 2..........................................................             2.3             3.0             4.8
PADD 4..........................................................             1.7             4.1             5.8
PADD 5..........................................................             1.0             2.8             4.3
-----------------------------------------------------------------
                           2010 15 ppm Sulfur Cap: Nonroad Diesel Fuel
-------------------------------------------------------------------------------------------------
PADDs 1 and 3...................................................             1.8             3.0             5.4
PADD 2..........................................................             2.9             6.1             7.4
PADD 4..........................................................             3.0             8.9             9.3
PADD 5..........................................................             1.7             5.9            8.4
----------------------------------------------------------------------------------------------------------------
Notes:
a At the current wholesale price of approximately $1.00 per gallon, these values also represent the percentage
  increase in diesel fuel price.

[[Page 28439]]

    The lower end of the range assumes that prices within a PADD 
increased to reflect the highest operating cost increase faced by any 
refiner in that PADD. In this case, this refiner with the highest 
operating cost would not recover any of his invested capital, but all 
other refiners would recover some or all of their investment. In this 
case, the price of nonroad, locomotive and marine diesel fuel would 
increase in 2007 by 1-2 cents per gallon, depending on the area of the 
country. In 2010, the price of nonroad diesel fuel would increase a 
total of 2-3 cents per gallon. Locomotive and marine diesel fuel prices 
would continue to increase by 1-2 cents per gallon.
    The mid-range estimate of price impacts assumes that prices within 
a PADD increase by the average refining and distribution cost within 
that PADD, including full recovery of capital (at 7 percent per annum 
before taxes). Lower cost refiners would recover more than their 
capital investment, while those with higher than average costs recover 
less. Under this assumption, the price of nonroad, locomotive and 
marine diesel fuel would increase in 2007 by 2-4 cents per gallon, 
depending on the area of the country. In 2010, the price of nonroad 
diesel fuel would increase a total of 3-9 cents per gallon. Locomotive 
and marine diesel fuel prices would continue to increase by 2-4 cents 
per gallon.
    The upper end estimate of price impacts assumes that prices within 
a PADD increase by the maximum total refining and distribution cost of 
any refinery within that PADD, including full recovery of capital (at 7 
percent per annum before taxes). All other refiners would recover more 
than their capital investment. Under this assumption, the price of 
nonroad, locomotive and marine diesel fuel would increase in 2007 by 3-
6 cents per gallon, depending on the area of the country. In 2010, the 
price of nonroad diesel fuel would increase a total of 5-9 cents per 
gallon. Locomotive and marine diesel fuel prices would continue to 
increase by 3-6 cents per gallon.
    In addition to the differences noted above, there are a number of 
assumptions inherent in all three of the above price projections. 
First, both the lower and upper limits of the projected price impacts 
described above assume that the refinery facing the highest compliance 
costs is currently the price setter in their market. This is a worse 
case assumption which is impossible to validate. Many factors affect a 
refinery's total costs of fuel production. Most of these factors, such 
as crude oil cost, labor costs, age of equipment, etc., are not 
considered in projecting the incremental costs associated with lower 
NRLM diesel fuel sulfur levels. Thus, current prices may very well be 
set in any specific market by a refinery facing lower incremental 
compliance costs than other refineries. This point was highlighted in a 
study by the National Economic Research Associates (NERA) for AAM of 
the potential price impacts of EPA's 2007 highway diesel fuel 
program.\286\ In that study, NERA criticized the above referenced study 
performed by Charles River Associates, et al. for API, which projected 
that prices would increase nationwide to reflect the total cost faced 
by the U.S. refinery with the maximum total compliance cost of all the 
refineries in the U.S. producing highway diesel fuel. To reflect the 
potential that the refinery with the highest projected compliance costs 
under the maximum price scenario is not the current price setter, we 
included the mid-point price impacts above. It is possible that even 
the lower limit price impacts are too high, if the conditions exist 
where prices are set based on operating costs alone. However, these 
price impacts are sufficiently low that considering even lower price 
impacts was not considered critical to estimating the potential 
economic impact of this rule.
---------------------------------------------------------------------------

    \286\ ``Potential Impacts of Environmental Regulations on Diesel 
Fuel Prices,'' NERA, for AAM, December 2000.
---------------------------------------------------------------------------

    Second, we assumed that a single refinery's costs could affect fuel 
prices throughout an entire PADD. While this is a definite improvement 
over analyses which assume that a single refinery's costs could affect 
fuel prices throughout the entire nation, it is still conservative. 
High cost refineries are more likely to have a more limited 
geographical impact on market pricing than an entire PADD.
    Third, by focusing solely on the cost of desulfurizing NRLM diesel 
fuel, we assume that the production of NRLM diesel fuel is independent 
of the production of other refining products, such as gasoline, jet 
fuel and highway diesel fuel. However, this is clearly not the case. 
Refiners have some flexibility to increase the production of one 
product without significantly affecting the others, but this 
flexibility is quite limited. It is possible that the relative 
economics of producing other products could influence a refiner's 
decision to increase or decrease the production of NRLM diesel fuel 
under the proposed standards. This in turn could increase or decrease 
the price impact relative to those projected above.
    Fourth, all three of the above price projections are based on the 
projected cost for U.S. refineries of meeting the proposed NRLM diesel 
fuel sulfur caps. Thus, these price projections assume that imports of 
NRLM fuel, which are currently significant in the Northeast, are 
available at roughly the same cost as those for U.S. refineries in 
PADDs 1 and 3. We have not performed any analysis of the cost of lower 
sulfur caps on diesel fuel produced by foreign refiners. However, there 
are reasons to believe that imports of 500 and 15 ppm NRLM diesel fuel 
would be available at prices in the ranges of those projected for U.S. 
refiners.
    One recent study analyzed the relative cost of lower sulfur caps 
for Asian refiners relative to those in the U.S., Europe and 
Japan.\287\ It concluded that costs for Asian refiners would be 
comparatively higher, due to the lack of current hydrotreating capacity 
at Asian refineries. This conclusion is certainly valid when evaluating 
lower sulfur levels for highway diesel fuels which are already at low 
levels in the U.S., Europe and Japan and for which refineries in these 
areas have already invested in hydrotreating capacity. It would appear 
to be less valid when assessing the relative cost of meeting lower 
sulfur standards for nonroad diesel fuels and heating oils which are 
currently at much higher sulfur levels in the U.S., Europe and Japan. 
All refineries face additional investments to remove sulfur from these 
fuels and so face roughly comparable control costs on a per gallon 
basis.
---------------------------------------------------------------------------

    \287\ ``Cost of Diesel Fuel Desulfurization In Asian 
Refineries,'' Estrada International Ltd., for the Asian Development 
Bank, December 17, 2002.
---------------------------------------------------------------------------

    One factor arguing for competitively priced imports is the fact 
that refinery utilization rates are currently higher in the U.S. and 
Europe than in the rest of the world. The primary issue is whether 
overseas refiners will invest to meet tight sulfur standards for U.S., 
European and Japanese markets. Many overseas refiners will not invest, 
instead focusing on local, higher sulfur markets. However, many 
overseas refiners focus on exports. Both Europe and the U.S. are moving 
towards highway and nonroad diesel fuel sulfur caps in the 10-15 ppm 
range. Europe is currently and projected to continue to need to import 
large volumes of highway diesel fuel. Thus, it seems reasonable to 
expect that a number of overseas refiners would invest in the capacity 
to produce some or all of their diesel fuel at these levels. Overseas 
refiners also have the flexibility to produce 10-15 ppm diesel fuel 
from their cleanest blendstocks, as

[[Page 28440]]

most of their available markets have less stringent sulfur standards. 
Thus, there are reasons to believe that some capacity to produce 10-15 
ppm diesel fuel would be available overseas at competitive prices. If 
these refineries were operating well below capacity, they might be 
willing to supply complying product at prices which only reflect 
incremental operating costs. This could hold prices down in areas where 
importing fuel is economical. However, it is unlikely that these 
refiners could supply sufficient volumes to hold prices down 
nationwide. Despite this expectation, to be conservative, in the 
refining cost analysis conducted earlier in this chapter, we assumed no 
imports of 500 ppm or 15 ppm NRLM diesel fuel. All 500 ppm and 15 ppm 
nonroad diesel fuel was produced by domestic refineries. This raised 
the average and maximum costs of 500 ppm and 15 ppm NRLM diesel fuel 
and increased the potential price impacts projected above beyond what 
would have been projected had we projected that 5-10 percent of NRLM 
diesel fuel would be imported at competitive prices.

B. Cost Savings to the Existing Fleet from the Use of Low Sulfur Fuel

    We estimate that reducing fuel sulfur to 500 ppm would reduce 
engine wear and oil degradation to the existing nonroad diesel 
equipment fleet and that a further reduction to 15 ppm sulfur would 
result in even greater reductions. This reduction in wear and oil 
degradation would provide a dollar savings to users of nonroad 
equipment. The cost savings would also be realized by the owners of 
future nonroad engines that are subject to the standards in this 
proposal. As discussed below, these maintenance savings have been 
conservatively estimated to be greater than 3 cents per gallon for the 
use of 15 ppm sulfur fuel when compared to the use of today's 
unregulated nonroad diesel fuel. A summary of the benefits of low-
sulfur fuel is presented in Table V.B-1.\288\
---------------------------------------------------------------------------

    \288\ See Heavy-duty 2007 Highway Final RIA, Chapter V.C.5, and 
``Study of the Effects of Reduced Diesel Fuel Sulfur Content on 
Engine Wear'', EPA report # 460/3-87-002, June 1987.

   Table V.B-1--Engine Components Potentially Affected by Lower Sulfur
                          Levels in Diesel Fuel
------------------------------------------------------------------------
                                 Effect of Lower    Potential Impact on
     1Affected Components             Sulfur           Engine System
------------------------------------------------------------------------
Piston Rings..................  Reduced corrosion  Extended engine life
                                 wear.              and less frequent
                                                    rebuilds.
Cylinder Liners...............  Reduced corrosion  Extended engine life
                                 wear.              and less frequent
                                                    rebuilds.
Oil Quality...................  Reduced deposits,  Reduce wear on piston
                                 reduced acid       ring and cylinder
                                 build-up, and      liner and less
                                 less need for      frequent oil
                                 alkaline           changes.
                                 additives.
Exhaust System (tailpipe).....  Reduced corrosion  Less frequent part
                                 wear.              replacement.
Exhaust Gas Recirculation       Reduced corrosion  Less frequent part
 System.                         wear.              replacement.
------------------------------------------------------------------------

    The monetary value of these benefits over the life of the equipment 
will depend upon the length of time that the equipment operates on low-
sulfur diesel fuel and the degree to which engine and equipment 
manufacturers specify new maintenance practices and the degree to which 
equipment operators change engine maintenance patterns to take 
advantage of these benefits. For equipment near the end of its life in 
the 2008 time frame, the benefits will be quite small. However, for 
equipment produced in the years immediately preceding the introduction 
of 500 ppm sulfur fuel, the savings would be substantial. Additional 
savings would be realized in 2010 when the 15 ppm sulfur fuel would be 
introduced.
    We estimate the single largest savings would be the impact of lower 
sulfur fuel on oil change intervals. The draft RIA presents our 
analysis for the oil change interval extension which would be realized 
by the introduction of 500 ppm sulfur fuel in 2007, as well as the 
additional oil extension which would be realized with the introduction 
of 15 ppm sulfur nonroad diesel fuel in 2010. As explained in the draft 
RIA, these estimates are based on our analysis of publically available 
information from nonroad engine manufacturers. Due to the wide range of 
diesel fuel sulfur which today's nonroad engines may see around the 
world, engine manufacturers specify different oil change intervals as a 
function of diesel sulfur levels. We have used this data as the basis 
for our analysis. Taken together, when compared to today's relatively 
high nonroad diesel fuel sulfur levels, we estimate the use of 15 ppm 
sulfur fuel will enable an oil change interval extension of 35 percent 
from today's products.
    We present here a fuel cost savings attributed to the oil change 
interval extension in terms of a cents per gallon operating cost. We 
estimate that an oil change interval extension of 31 percent, as would 
be enabled by the use of 500 ppm sulfur fuel in 2007, results in a fuel 
operating costs savings of 3.0 cents per gallon for the nonroad fleet. 
We project an additional cost savings of 0.3 cents per gallon for the 
oil change interval extension which would be enabled by the use of 15 
ppm sulfur beginning in 2010. Thus, for the nonroad fleet as a whole, 
beginning in 2010 nonroad equipment users can realize an operating cost 
savings of 3.3 cents per gallon compared to today's engine. This means 
that the end cost to the typical user for 15ppm sulfur fuel is 
approximately 1.5 cents per gallon (4.8 cent per gallon cost for fuel 
minus 3.3 cent per gallon maintenance savings). For a typical 100 
horsepower nonroad engine this represents a net present value lifetime 
savings of more than $500.
    These savings will occur without additional new cost to the 
equipment owner beyond the incremental cost of the low-sulfur diesel 
fuel, although these savings are dependent on changes to existing 
maintenance schedules. Such changes seem likely given the magnitude of 
the savings. We have not estimated the value of the savings from the 
other benefits listed in Table V.B-1, and therefore we believe the 3.3 
cents per gallon savings is conservative as it only accounts for the 
impact of low sulfur fuel on oil change intervals.

C. Engine and Equipment Cost Impacts

    The following sections briefly discuss the various engine and 
equipment cost elements considered for this proposal and present the 
total costs we have estimated; the reader is referred to the draft RIA 
for a complete discussion. Estimated engine and equipment costs depend 
largely on both the size of the piece of equipment and its engine, and 
on the technology package being added to the engine to ensure 
compliance with the proposed standards. The wide size variation (e.g., 
<4 horsepower engines through £2500 horsepower engines) and

[[Page 28441]]

the broad application variation (e.g., lawn equipment through large 
mining trucks) that exists in the nonroad industry makes it difficult 
to present here an estimated cost for every possible engine and/or 
piece of equipment. Nonetheless, for illustrative purposes, we present 
some example per engine/equipment cost impacts throughout this 
discussion. This analysis is presented in detail in Chapter 6 of the 
draft RIA. We are also considering doing a sensitivity analysis on 
cost/engine data, which would be put into the docket for comment.
    It is important to note that the costs presented here do not 
reflect any savings that are expected to occur because of the engine 
ABT program and the equipment manufacturer transition program, both of 
which are discussed in Section VII. As discussed in the draft RIA, 
these optional programs have the potential to provide significant 
savings for both engine and equipment manufacturers. We request comment 
with supporting data and/or analysis on the cost estimates presented 
here and the underlying analysis presented in chapter 6 of the draft 
RIA.
1. Engine Cost Impacts
    Estimated engine costs are broken into fixed costs (for research 
and development, retooling, and certification), variable costs (for new 
hardware and assembly time), and life-cycle operating costs. Total 
operating costs include the estimated incremental cost for low-sulfur 
diesel fuel, any expected increases in maintenance costs associated 
with new emission control devices, any costs associated with increased 
fuel consumption, and any decreases in operating cost (i.e., 
maintenance savings) expected due to low-sulfur fuel. Cost estimates 
presented here represent an expected incremental cost of engines in the 
model year of their introduction. Costs in subsequent years would be 
reduced by several factors, as described below. All engine and 
equipment costs are presented in 2001 dollars.
a. Engine Fixed Costs
i. Engine and Emission Control Device R&D
    The technologies described in section III represent those 
technologies we believe will be used to comply with the proposed Tier 4 
emission standards. These technologies are part of an ongoing research 
and development effort geared toward compliance with the 2007 heavy-
duty diesel highway emission standards. The engine manufacturers making 
R&D expenditures toward compliance with highway emission standards will 
have to undergo some additional R&D effort to transfer emission control 
technologies to engines they wish to sell into the nonroad market. 
These R&D efforts will allow engine manufacturers to develop and 
optimize these new technologies for maximum emission-control 
effectiveness with minimum negative impacts on engine performance, 
durability, and fuel consumption. Many nonroad engine manufacturers are 
not part of the ongoing R&D effort toward compliance with highway 
emissions standards because they do not sell engines into the highway 
market. These manufacturers are expected to benefit from the R&D work 
that has already occurred and will continue through the coming years 
through their contact with highway manufacturers, emission control 
device manufacturers, and the independent engine research laboratories 
conducting relevant R&D.
    Several technologies are projected for complying with the proposed 
Tier 4 emission standards. We are projecting that NOX 
adsorbers and catalyzed diesel particulate filters (CDPFs) would be the 
most likely technologies applied by industry to meet our proposed 
emissions standards for £75 horsepower engines. The fact that 
these technologies are being developed for implementation in the 
highway market prior to the implementation dates in this proposal, and 
the fact that engine manufacturers would have several years before 
implementation of the proposed Tier 4 standards, ensures that the 
technologies used to comply with the nonroad standards would undergo 
significant development before reaching production. This ongoing 
development could lead to reduced costs in three ways. First, we expect 
research will lead to enhanced effectiveness for individual 
technologies, allowing manufacturers to use simpler packages of 
emission control technologies than we would predict given the current 
state of development. Similarly, we anticipate that the continuing 
effort to improve the emission control technologies will include 
innovations that allow lower-cost production. Finally, we believe that 
manufacturers would focus research efforts on any drawbacks, such as 
fuel economy impacts or maintenance costs, in an effort to minimize or 
overcome any potential negative effects.
    We anticipate that, in order to meet the proposed standards, 
industry would introduce a combination of primary technology upgrades. 
Achieving very low NOX emissions would require basic 
research on NOX emission control technologies and 
improvements in engine management to take advantage of the exhaust 
emission control system capabilities. The manufacturers are expected to 
take a systems approach to the problem of optimizing the engine and 
exhaust emission control system to realize the best overall 
performance. Since most research to date with exhaust emission control 
technologies for nonroad applications has focused on retrofit programs, 
there remains room for significant improvements by taking such a 
systems approach. The NOX adsorber technology in particular 
is expected to benefit from re-optimization of the engine management 
system to better match the NOX adsorber's performance 
characteristics. The majority of the dollars we have estimated for 
research is expected to be spent on developing this synergy between the 
engine and NOX exhaust emission control systems. Therefore, 
for engines requiring both a CDPF and a NOX adsorber (i.e., 
£75 horsepower), we have attributed two-thirds of the R&D 
expenditures to NOX control, and one-third to PM control.
    In the 2007 HD highway rule, we estimated that each engine 
manufacturer would expend $35 million for R&D to redesign their engines 
and apply catalyzed diesel particulate filters (CDPF) and 
NOX adsorbers. For their nonroad R&D efforts on engines 
requiring CDPFs and NOX adsorbers (i.e., £75 
horsepower), engine manufacturers selling into the highway market would 
incur some level of R&D effort but not at the level incurred for the 
highway rule. In many cases, the engines used by highway manufacturers 
in nonroad products are based on the same engine platform as those used 
in highway products. However, horsepower and torque characteristics are 
often different so some effort will have to be expended to accommodate 
those differences. For these manufacturers, we have estimated that they 
would incur an R&D expense of $3.5 million. This $3.5 million R&D 
expense would allow for the transfer of R&D knowledge from their 
highway experience to their nonroad engine product line. Two-thirds of 
this R&D is attributed to NOX control and one-third to PM 
control.
    For those manufacturers that sell engines only into the nonroad 
market, and where those engines require a CDPF and a NOX 
adsorber, we believe that they will incur an R&D expense nearing that 
incurred by highway manufacturers for the highway rule, although not at 
the level incurred by highway manufacturers for the highway rule. 
Nonroad manufacturers would be able to learn from the R&D efforts 
already

[[Page 28442]]

under way for both the highway rule and for the Tier 2 light-duty 
highway rule (65 FR 6698). This learning could be done via seminars, 
conferences, and contact with highway manufacturers, emission control 
device manufacturers, and the independent engine research laboratories 
conducting relevant R&D. Therefore, for these manufacturers, we have 
estimated an expenditure of $24.5 million. This lower number--$24.5 
million versus $35 million in the highway rule--reflects the transfer 
of knowledge to nonroad manufacturers that would occur from the many 
stakeholders in the diesel industry. Two-thirds of this R&D is 
attributed to NOX control and one-third to PM control.
    Note that the $3.5 million and $24.5 million estimates represent 
our estimate of the average R&D expected by manufacturers. These 
estimates would be different for each manufacturer--some higher, some 
lower--depending on product mix and the ability to transfer knowledge 
from one product to another.
    For those engine manufacturers selling engines that would require 
CDPF-only R&D (i.e., 25 to 75 horsepower engines in 2013), we have 
estimated that the R&D they would incur would be roughly one-third that 
incurred by manufacturers conducting CDPF/NOX adsorber R&D. 
We believe this is a good estimate because CDPF technology is further 
along in its development than is NOX adsorber technology 
and, therefore, a 50/50 split would not be appropriate. Using this 
estimate, the R&D incurred by manufacturers that have already done 
selling any engines into both the highway and the nonroad markets would 
be $1.2 million, and the R&D for manufacturers selling engines into 
only the nonroad market would be roughly $8 million. All of this R&D is 
attributed to PM control.
    For those engine manufacturers selling engines that would require 
DOC-only or some engine-out modification R&D (i.e., <75 horsepower 
engines in 2008), we have estimated that the R&D they would incur would 
be roughly one-half the amount estimated for their CDPF-only R&D. Using 
this estimate, the R&D incurred by manufacturers selling any engines 
into both the highway and nonroad markets would be roughly $600,000, 
and the R&D for manufacturers selling engines into only the nonroad 
market would be roughly $4 million. All of this R&D is attributed to PM 
control.
    Some manufacturers of engines produce engines to specifications 
developed by other manufacturers. Such joint venture manufacturers do 
not conduct engine-related R&D but simply manufacture an engine 
designed and developed by another manufacturer. For such manufacturers, 
we have assumed no R&D expenditures given that we believe they will 
conduct no R&D themselves and will rely on their joint venture partner. 
This is true unless the parent company has no engine sales in the 
horsepower categories covered by the partner company. Under such a 
situation, we have accounted for the necessary R&D by attributing it to 
the parent company. We have also estimated that some manufacturers will 
choose not to invest in R&D for the U.S. nonroad market due to low 
volume sales that probably cannot justify the expense. More detail on 
these assumptions and the number of manufacturers assumed not to expend 
R&D is presented in Chapter 6 of the draft RIA. We welcome comments and 
supporting documentation.
    We have assumed that all R&D expenditures occur over a five year 
span preceding the first year any emission control device is introduced 
into the market. Where a phase-in exists (e.g., for NOX 
standards on £75 horsepower engines), expenditures are 
assumed to occur over the five year span preceding the first year 
NOX adsorbers would be introduced, and then to continue 
during the phase-in years; the expenditures would be incurred in a 
manner consistent with the phase-in of the standard. All R&D 
expenditures are then recovered by the engine manufacturer over an 
identical time span following the introduction of the technology. We 
assume a seven percent rate of return for all R&D. We have apportioned 
these R&D costs across all engines that are expected to use these 
technologies, including those sold in other countries or regions that 
are expected to have similar standards. We have estimated the fraction 
of the U.S. sales to this total sales at 42 percent. Therefore, we have 
attributed this amount to U.S. sales.
    Using this methodology, we have estimated the total R&D 
expenditures attributable to the proposed standards at $199 million.
ii. Engine-Related Tooling Costs
    Once engines are ready for production, new tooling will be required 
to accommodate the assembly of the new engines. In the 2007 highway 
rule, we estimated approximately $1.6 million per engine line for 
tooling costs associated with CDPF/NOX adsorber systems. For 
the proposed nonroad Tier 4 standards, we have estimated that nonroad-
only manufacturers would incur the same $1.6 million per engine line 
requiring a CDPF/NOX adsorber system and that these costs 
would be split evenly between NOX control and PM control. 
For those systems requiring only a CDPF, we have estimated one-half 
that amount, or $800,000 per engine line. For those systems requiring 
only a DOC or some engine-out modifications, we have applied a one-half 
factor again, or $400,000 per engine line. Tooling costs for CDPF-only 
and for DOC engines are attributed solely to PM control.
    For those manufacturers selling into both the highway and nonroad 
markets, we have estimated one-half the baseline tooling cost, or 
$800,000, for those engine lines requiring a CDPF/NOX 
adsorber system. We believe this is reasonable since many nonroad 
engines are produced on the same engine line with their highway 
counterparts. For such lines, we believe very little to no tooling 
costs would be incurred. For engine lines without a highway 
counterpart, something approaching the $1.6 million tooling cost would 
be applicable. For this analysis, we have assumed a 50/50 split of 
engine product lines for highway manufacturers and, therefore, a 50 
percent factor applied to the $1.6 million baseline. These tooling 
costs would be split evenly between NOX control and PM 
control. For engine lines <75 horsepower, we have used the same tooling 
costs as the nonroad-only manufacturers because these engines tend not 
to have a highway counterpart. Therefore, for those engine lines 
requiring only a CDPF (i.e., those between 25 and 75 horsepower), we 
have estimated a tooling cost of $800,000. Similarly, the tooling costs 
for DOC and/or engine-out engine lines has been estimated to be 
$400,000. Tooling costs for CDPF-only and for DOC engines are 
attributed solely to PM control.
    We expect engines in the 25 to 50 horsepower range to apply EGR 
systems to meet the proposed NOX standards for 2013. For 
these engines, we have included an additional tooling cost of $40,000 
per engine line, consistent with the EGR-related tooling cost estimated 
for 50-100 horsepower engines in our Tier 2/3 rulemaking. This tooling 
cost is applied equally to all engine lines in that horsepower range 
regardless of the markets into which the manufacturer sells. We have 
applied this tooling cost equally because engines in this horsepower 
range do not tend to have highway counterparts. Tooling costs for EGR 
systems are attributed solely to NOX control.
    We have applied all the above tooling costs to all manufacturers 
that appear to actually make engines. We have not

[[Page 28443]]

eliminated joint venture manufacturers because these manufacturers 
would still need to invest in tooling to make the engines even if they 
do not conduct any R&D. We have assumed that all tooling costs are 
incurred one year in advance of the new standard and are recovered over 
a five year period following implementation of the new standard; all 
tooling costs are marked up seven percent to reflect the time value of 
money. As done for R&D costs, we have attributed a portion of the 
tooling costs to U.S. sales and a portion to sales in other countries 
expected to have similar levels of emission control. More information 
is contained in Chapter 6 of the draft RIA and we request comment on 
how we have applied our tooling cost estimates and to whom we have 
applied them.
    Using this methodology, we estimate the total tooling expenditures 
attributable to the proposed standards at $67 million.
iii. Engine Certification Costs
    Manufacturers will incur more than the normal level of 
certification costs during the first few years of implementation 
because engines will need to be certified to the new emission 
standards. Consistent with our recent standard setting regulations, we 
have estimated engine certification costs at $60,000 per new engine 
certification to cover testing and administrative costs. To this we 
have added the proposed certification fee of $2,156 per new engine 
family. This cost, $62,156 per engine family was used for <75 
horsepower engines certifying to the 2008 standards. For 25 to 75 
horsepower engines certifying to the 2013 standards, and for 
£75 horsepower engines certifying to their proposed 
standards, we have added costs to cover the proposed test procedures 
for nonroad diesel engines (i.e., the transient test and the NTE); 
these costs were estimated at $10,500 per engine family. These 
certification costs--whether it be the $62,156 or the $72,656 per 
engine family--apply equally to all engine families for all 
manufacturers regardless of into what markets the manufacturer sells. 
We have applied these certification costs to only the US sold engines 
because the certification conducted for US sales is not presumed to 
fulfill the certification requirements of other countries.
    Applying these costs to each of the 665 engine families as they are 
certified to a new emissions standard results in total costs of $72 
million expended during implementation of the proposed standards. These 
costs are attributed to NOX and PM control consistent with 
the phase-in of the new emissions standards--where new NOX 
and PM standards are introduced together, the certification costs are 
split evenly; where only a new PM standard is introduced, the 
certification costs are attributed to PM only; where a NOX 
phase-in becomes 100% in a year after full implementation of a PM 
standard, the certification costs are attributed to NOX 
only. All certification costs are assumed to occur one year prior to 
the new emission standard and are then recovered over a five year 
period following compliance with the new standard; all certification 
costs are marked up seven percent to reflect the time value of money.
b. Engine Variable Costs
    This section summarizes the detailed analysis presented in the 
draft RIA for this proposed rule. We encourage the reader to refer to 
chapter 6 of that draft RIA for the details of what is presented here 
and encourage comments and supporting data and/or analysis regarding 
those details. Of particular interest are comments regarding the costs 
of precious metals, or platinum group metals (PGM). The PGM costs are a 
significant fraction of the total costs for aftertreatment devices. For 
our analysis, we have used the 2002 annual average costs for platinum 
and rhodium (the two PGMs we expect will be used) because we believe 
they represent a better estimate of the cost for PGM than other 
metrics. We request comment on this approach and whether an alternative 
approach would be more appropriate. Specifically, we request comment 
regarding the use of a five year average in place of the one year 
average we have used. Additionally, EPA invites comment on the impacts, 
if any, that this rulemaking would have in the context of a variety of 
rulemakings on the market impacts on precious metals.
i. NOX Adsorber System Costs
    The NOX adsorber system that we are anticipating would 
be applied for Tier 4 would be the same as that used for highway 
applications. In order for the NOX adsorber to function 
properly, a systems approach that includes a reductant metering system 
and control of engine A/F ratio is also necessary. Many of the new air 
handling and electronic system technologies developed in order to meet 
the Tier 2/3 nonroad engine standards can be applied to accomplish the 
NOX adsorber control functions as well. Some additional 
hardware for exhaust NOX or O2 sensing and for 
fuel metering will likely be required. The cost estimates include a DOC 
for clean-up of hydrocarbon emissions that occur during NOX 
adsorber regeneration events. We have also assumed that warranty costs 
would increase due to the application of this new hardware. Chapter 6 
of the draft RIA contains the details for how we estimated costs 
associated with the new NOX control technologies required to 
meet the proposed Tier 4 emission standards. These costs are estimated 
to increase engine costs by roughly $670 in the near-term for a 150 
horsepower engine, and $2,070 in the near-term for a 500 horsepower 
engine. In the long-term, we estimate these costs to be $550 and $1,670 
for the 150 horsepower and 500 horsepower engines, respectively. Note 
that we have estimated costs for all engines in all horsepower ranges, 
and these estimates are presented in detail in the draft RIA. 
Throughout this discussion of engine and equipment costs, we present 
costs for a 150 and a 500 horsepower engine for illustrative purposes.
ii. Catalyzed Diesel Particulate Filter (CDPF) Costs
    CDPFs can be made from a wide range of filter materials including 
wire mesh, sintered metals, fibrous media, or ceramic extrusions. The 
most common material used for CDPFs for heavy-duty diesel engines is 
cordierite. We have based our cost estimates on the use of silicon 
carbide (SiC) even though it is more expensive than other filter 
materials. We request comment on our assumption that SiC will be used 
in favor of cordierite. We estimate that the CDPF systems will add $780 
to engine costs in the near-team for a 150 horsepower engine and $2,770 
in the near-term for a 500 horsepower engine. In the long-term, we 
estimate these CDPF system costs to be $590 and $2,110 for the 150 
horsepower and the 500 horsepower engines, respectively.
iii. CDPF Regeneration System Costs
    Application of CDPFs in nonroad applications is expected to present 
challenges beyond those of highway applications. For this reason, we 
anticipate that some additional hardware beyond the diesel particulate 
filter itself may be required to ensure that CDPF regeneration occurs. 
For some engines this may be new fuel control strategies that force 
regeneration under some circumstances, while in other engines it might 
involve an exhaust system fuel injector to inject fuel upstream of the 
CDPF to provide necessary heat for regeneration under some operating 
conditions. We estimate the near-term costs of a CDPF regeneration 
system to be $190 for a 150

[[Page 28444]]

horsepower engine and $320 for a 500 horsepower engine. In the long-
term, we estimate these costs at $140 and $240, respectively.
iv. Closed-Crankcase Ventilation System (CCV) Costs
    We are proposing to eliminate the exemption that allows turbo-
charged nonroad diesel engines to vent crankcase gases directly to the 
environment. Such engines are said to have an open crankcase system. We 
project that this requirement to close the crankcase on turbo-charged 
engines would force manufacturers to rely on engineered closed 
crankcase ventilation systems that filter oil from the blow-by gases 
prior to routing them into either the engine intake or the exhaust 
system upstream of the CDPF. We have estimated the initial cost of 
these systems to be roughly $40 for low horsepower engines and up to 
$100 for very high horsepower engines. These costs are incurred only by 
turbo-charged engines because today's naturally aspirated engines 
already have CCV systems.
v. Variable Costs for Engines Below 75 Horsepower and Above 750 
Horsepower
    This proposal includes standards for engines <25 horsepower that 
begin in 2008, and two sets of standards for 25 to 75 horsepower 
engines--one set that begins in 2008 and another that begins in 2013. 
The 2008 standards for all engines <75 horsepower are of similar 
stringency and are expected to result in similar technologies (i.e., 
the addition of a DOC). The 2013 standards for 25 to 75 horsepower 
engines are considerably more stringent than the 2008 standards and are 
expected to force the addition of a CDPF along with some other engine 
hardware to enable the proper functioning of that new technology. More 
detail on the mix of technologies expected for all engines <75 
horsepower is presented in section III. As discussed there, if changes 
are needed to comply, we expect manufacturers to comply with the 2008 
standards through either engine improvements or through the addition of 
a DOC. From a cost perspective, we have projected that engines would 
comply by either adding a DOC or by making some engine modifications 
resulting in engine-out emission reductions. Presumably, the 
manufacturer would choose the least costly approach that provided the 
necessary reduction. If engine-out modifications are less costly than a 
DOC, our estimate here is conservative. If the DOC proves to be less 
costly, then our estimate is representative of what most manufacturers 
would do. Therefore, we have assumed that, beginning in 2008, all 
engines below 75 horsepower add a DOC. Note that this is a conservative 
estimate in that we have assume this cost for all engines when, as 
discussed in section IV, some engines <75 horsepower already meet the 
proposed PM standards. We have estimated this added hardware to result 
in an increased engine cost of $150 in the near-term and $140 in the 
long-term for a 30 horsepower engine.
    We have also projected that some engines in the 25 to 75 horsepower 
range would have to upgrade their fuel systems to accommodate the CDPF. 
We have estimated the incremental costs for these fuel systems at 
roughly $740 in the 25-50 horsepower range, and around $430 in the 50-
75 horsepower range. This difference reflects a different base fuel 
system, with the smaller engines assumed to have mechanical fuel 
systems and the larger engines assumed to already be electronic. The 
electronic systems will incur lower costs because they already have the 
control unit and electronic fuel pump. Also, we have assumed these fuel 
changes would occur for only direct injection (DI) engines; indirect 
injection engines (IDI) are assumed to remain IDI but to add more 
hardware as part of their CDPF regeneration system to ensure proper 
regeneration under all operating conditions. Such a regeneration 
system, described above, is expected to cost roughly twice that 
expected for DI engines, or around $320 for a 30 horsepower IDI engine 
versus $160 for a DI engine.
    We have also projected that engines in the 25-50 horsepower range 
would add cooled EGR to comply with their new NOX standard. 
We have estimated that this would add $90 in the near-term and $70 in 
the long-term to the cost of a 30 horsepower engine.
    We believe there are factors that would cause variable hardware 
costs to decrease over time, making it appropriate to distinguish 
between near-term and long-term costs. Research in the costs of 
manufacturing has consistently shown that as manufacturers gain 
experience in production, they are able to apply innovations to 
simplify machining and assembly operations, use lower cost materials, 
and reduce the number or complexity of component parts.\289\ Our 
analysis, as described in more detail in the draft RIA, incorporates 
the effects of this learning curve by projecting that the variable 
costs of producing the low-emitting engines decreases by 20 percent 
starting with the third year of production. For this analysis, we have 
assumed a baseline that represents such learning already having 
occurred once due to the 2007 highway rule (i.e., a 20 percent 
reduction in emission control device costs is reflected in our near-
term costs). We have then applied a single learning step from that 
point in this analysis. We invite comment on this methodology to 
account for the learning curve phenomenon and also request comment on 
whether learning is likely to reduce costs even further in this 
industry (e.g., should a second learning step be applied to our near-
term costs?). Additionally, manufacturers are expected to apply ongoing 
research to make emission controls more effective and to have lower 
operating costs over time. However, because of the uncertainty involved 
in forecasting the results of this research, we conservatively have not 
accounted for it in this analysis.
---------------------------------------------------------------------------

    \289\ ``Learning Curves in Manufacturing,'' Linda Argote and 
Dennis Epple, Science, February 23, 1990, Vol. 247, pp. 920-924.
---------------------------------------------------------------------------

c. Engine Operating Costs
    We are projecting that a variety of new technologies will be 
introduced to enable nonroad engines to meet the proposed Tier 4 
emissions standards. Primary among these are advanced emission control 
technologies and low-sulfur diesel fuel. The technology enabling 
benefits of low-sulfur diesel fuel are described in section III, and 
the incremental cost for low-sulfur fuel is described in section V.A. 
The new emission control technologies are themselves expected to 
introduce additional operating costs in the form of increased fuel 
consumption and increased maintenance demands. Operating costs are 
estimated in the draft RIA over the life of the engine and are 
expressed in terms of cents/gallon of fuel consumed. In section V.C.3, 
we present these lifetime operating costs as a net present value (NPV) 
in 2001 dollars for several example pieces of equipment.
    Total operating cost estimates include the following elements: the 
change in maintenance costs associated with applying new emission 
controls to the engines; the change in maintenance costs associated 
with low sulfur fuel such as extended oil change intervals; the change 
in fuel costs associated with the incrementally higher costs for low 
sulfur fuel, and the change in fuel costs due to any fuel consumption 
impacts associated with applying new emission controls to the engines. 
This latter cost is attributed to the CDPF and its need for periodic 
regeneration which we estimate may result in a one percent fuel 
consumption increase where a NOX

[[Page 28445]]

adsorber is also applied, or a two percent fuel consumption increase 
where no NOX adsorber is applied (refer to chapter 6, 
section 6.2.3.3). Maintenance costs associated with the new emission 
controls on the engines are expected to increase since these devices 
represent new hardware and, therefore, new maintenance demands. For 
CDPF maintenance, we have used a maintenance interval of 3,000 hours 
for smaller engines and 4,500 hours for larger engines and a cost of 
$65 through $260 for each maintenance event. For closed-crankcase 
ventilation (CCV) systems, we have used a maintenance interval of 675 
hours for all engines and a cost per maintenance event of $8 to $48 for 
small to large engines. Offsetting these maintenance cost increases 
would be a savings due to an expected increase in oil change intervals 
because low sulfur fuel would be far less corrosive than is current 
nonroad diesel fuel. Less corrosion would mean a slower acidification 
rate (i.e., less degradation) of the engine lubricating oil and, 
therefore, more operating hours between needed oil changes. As 
discussed in section V.B, the use of 15 ppm sulfur fuel can extend oil 
change intervals by as much as 35 percent for both new and existing 
nonroad engines and equipment. We have used a 35 percent increase in 
oil change interval along with costs per oil change of $70 through $400 
to arrive at estimated savings associated with increased oil change 
intervals.
    These operating costs are expressed as a cent/gallon cost (or 
savings). As a result, operating costs are directly proportional to the 
amount of fuel consumed by the engine. We have estimated these 
operating costs, inclusive of fuel-related costs, to be 3.4 cents/
gallon for a 150 horsepower engine and 4.2 cents/gallon for a 500 
horsepower engine. More detail on operating costs can be found in 
chapter 6 of the draft RIA.
    The existing fleet will also benefit from lower maintenance costs 
due to the use of low sulfur diesel fuel. The operating costs for the 
existing fleet are discussed in Section V.B.
2. Equipment Cost Impacts
    In addition to the costs directly associated with engines that 
incorporate new emission controls to meet new standards, we expect cost 
increases due to the need to redesign the nonroad equipment in which 
these engines are used. Such redesigns would probably be necessary due 
to the expected addition of new emission control systems, but could 
also occur if the engine has a different shape or heat rejection rate, 
or is no longer made available in the configuration previously used. 
Based on their past experiences, equipment manufacturers have told EPA 
that a major concern with a new standard is their ability to redesign a 
large number of applications in a short period of time. Therefore, we 
have provided equipment manufacturers transition flexibility provisions 
to help them avoid business disruptions resulting from the changes 
associated with new emission standards. These flexibility provisions 
are presented in detail in Section III.E.4.
    In assessing the economic impact of the new emission standards, EPA 
has made a best estimate of the modifications to equipment that relate 
to packaging (installing engines in equipment engine compartments). The 
incremental costs for new equipment would be comprised of fixed costs 
(for redesign to accommodate new emission control devices) and variable 
costs (for new equipment hardware and for labor to install new emission 
control devices). Note that the fixed costs do not include 
certification costs, as did the engine fixed costs, because equipment 
is not certified to emission standards. We have attributed all changes 
in operating costs (e.g., additional maintenance) to the cost estimates 
for engines. Included in section V.C.3 is a discussion of several 
example pieces of equipment (e.g., skid/steer loader, dozer, etc.) and 
the costs we have estimated for these specific example pieces of 
equipment. Full details of our equipment cost analysis can be found in 
chapter 6 of the draft RIA. All costs are presented in 2001 dollars.
a. Equipment Fixed Costs
    The most significant changes anticipated for equipment redesign are 
changes to accommodate the physical changes to engines, especially for 
those engines that add PM traps and NOX adsorbers. The costs 
for engine development and the emission control devices are included as 
costs to the engines, as described above. What remains to be quantified 
for equipment manufacturers is the effort to integrate the engine and 
emissions control devices into the overall functioning of the 
equipment. What remains to be quantified for equipment manufacturers is 
the effort to integrate the engine and emissions control devices into 
the overall functioning of the equipment. We have allocated extensive 
engineering time for this effort.
    The costs we have estimated are based on engine power and whether 
an application is non-motive (e.g., a generator set) or motive (e.g., a 
skid steer loader). The designs we have considered to be non-motive are 
those that lack a propulsion system. In addition, the proposed emission 
standards for engines rated under 25 horsepower and the proposed 2008 
standards for 25-75 horsepower engines are projected to require no 
significant equipment redesign beyond that done to accommodate the Tier 
2 standards. We expect that these engines would comply with the 
proposesd Tier 4 standards through either engine modifications to 
reduce engine-out emissions or through the addition of a DOC. We have 
projected that engine modifications would not affect the outer 
dimensions of the engine and that a DOC would replace the existing 
muffler. Therefore, either approach taken by the engine manufacturer 
should have minimal to no impact on the equipment design. Nonetheless, 
we have conservatively estimated their redesign costs at $50,000 per 
model.
    A number of equipment manufacturers have shared detailed 
information with us regarding the investments made for Nonroad Tier 2 
equipment redesign efforts, as well as redesign estimates for 
significant changes such as installing a new engine design. These 
estimates range from approximately $50,000 for some lower powered 
equipment models to well over $1 million dollars for high horsepower 
equipment with very challenging design constraints. Based on that 
input, for the proposed Tier 4 standards, we have estimated that 
equipment redesign costs would range from $50,000 per model for 25 
horsepower equipment up to $750,000 per model for 300 horsepower 
equipment and above. We have attributed only a portion of the equipment 
redesign costs to U.S. sales in a manner consistent with that taken for 
engine R&D costs and engine tooling costs. In addition, we expect 
manufacturers to incur some fixed costs to update service and operation 
manuals to address the maintenance demands of new emission control 
technologies and the new oil service intervals which we estimate to be 
between $2,500 and $10,000 per equipment model.
    These equipment fixed costs (redesign and manual updates) were then 
allocated appropriately to each new model to arrive at a total 
equipment fixed cost of $697 million. We have assumed that these costs 
would be recovered over a ten year period at a seven percent interest 
rate.
b. Equipment Variable Costs
    Equipment variable cost estimates are based on costs for additional 
materials to mount the new hardware (i.e., brackets and bolts required 
to secure the

[[Page 28446]]

aftertreatment devices) and additional sheet metal assuming that the 
body cladding of a piece of equipment (i.e., the hood) might change to 
accommodate the aftertreatment system. Variable costs also include the 
labor required to install these new pieces of hardware. For engines 
£75 horsepower--those expected to incorporate CDPF and 
NOX adsorber technology--the amount of sheet metal is based 
on the size of the aftertreatment devices.
    For equipment of 150 horsepower and 500 horsepower, respectively, 
we have estimated the costs to be roughly $60 to $140. Note that we 
have estimated costs for equipment in all horsepower ranges, and these 
estimates are presented in detail in the draft RIA. Throughout this 
discussion of engine and equipment costs, we present costs for a 150 
and a 500 horsepower engine for illustrative purposes.
3. Overall Engine and Equipment Cost Impacts
    To illustrate the engine and equipment cost impacts we are 
estimating for the proposed standards, we have chosen several example 
pieces of equipment and presented the estimated costs for them. Using 
these examples, we can calculate the costs for a specific piece of 
equipment in several horsepower ranges and better illustrate the cost 
impacts of the proposed standards. These costs along with information 
about each example piece of equipment are shown in Table V.C-1. Costs 
presented are near-term and long-term costs for the final standards to 
which each piece of equipment would comply. Long-term costs are only 
variable costs and, therefore, represent costs after all fixed costs 
have been recovered and all projected learning has taken place. 
Included in the table are estimated prices for each piece of equipment 
to provide some perspective on how our estimated control costs relate 
to existing equipment prices.

                                   Table V.C-1--Near-Term and Long-Term Costs for Several Example Pieces of Equipmenta
                                      ($2001, for the final emission standards to which the equipment must comply)
--------------------------------------------------------------------------------------------------------------------------------------------------------
                                                            Skid/steer                                                                      Off-highway
                                              GenSet          loader          Backhoe          Dozer        Ag tractor         Dozer           truck
--------------------------------------------------------------------------------------------------------------------------------------------------------
Horsepower                                          9 hp           33 hp           76 hp          175 hp          250 hp          503 hp        1,000 hp
Incremental engine & equipment cost
  Long-term                                         $120            $760          $1,210          $2,590          $2,000          $4,210          $6,780
  Near-term                                         $170          $1,100          $1,680           3,710          $2,950          $6,120         $10,100
Estimated equipment price when new b              $3,500         $13,500         $50,000        $235,000        $130,000        $575,000        $700,000
Incremental operating costs c                       -$90             $40            $370          $1,550          $1,320          $4,950         $12,550
Baseline operating costs (fuel & oil                $940          $2,680          $7,960         $77,850         $23,750         $77,850       $179,530
 only) c
--------------------------------------------------------------------------------------------------------------------------------------------------------
Notes:
a Near-term costs include both variable costs and fixed costs; long-term costs include only variable costs and represent those costs that remain
  following recovery of all fixed costs.
b ``Estimated Price of New Nonroad Example Equipment,'' memorandum from Zuimdie Guerra to docket A-2001-28.
c Present value of lifetime costs.

    More detail and discussion regarding what these costs and prices 
mean from an economic impact perspective can be found in section V.E.

D. Annual Costs and Cost Per Ton

    One tool that can be used to assess the value of the proposed 
standards for nonroad fuel and engines is the costs incurred per ton of 
emissions reduced. This analysis involves a comparison of our proposed 
program to other measures that have been or could be implemented.
    We have calculated the cost per ton of our proposed program based 
on the net present value of all costs incurred and all emission 
reductions generated over a 30 year time window following 
implementation of the program. This approach captures all of the costs 
and emissions reductions from our proposed program including those 
costs incurred and emissions reductions generated by the existing 
fleet. The baseline (i.e., the point of comparison) for this evaluation 
is the existing set of fuel and engine standards (i.e., unregulated 
fuel and the Tier 2/Tier 3 program). The 30 year time window chosen is 
meant to capture both the early period of the program when very few new 
engines that meet the proposed standards would be in the fleet, and the 
later period when essentially all engines would meet the proposed 
standards.
    As discussed in section IV, the proposal contains two separate fuel 
programs. We are proposing a 500 ppm sulfur cap on nonroad, locomotive, 
and marine fuels beginning in 2007. This fuel program, the first step 
in our two step fuel program, provides significant air quality benefits 
through reduced SO2 and PM emissions from both new and 
existing nonroad, locomotive, and marine engines. In sections V.D.1 and 
2, we summarize the cost for this program as if it remained in place 
for 30 years, even though it would be supplanted by the second step of 
our fuel program in 2010. We also provide an analysis of the cost per 
ton for the SO2 reductions that would be realized by the 500 
ppm fuel program for the same 30 year time window. In this way, the 
cost per ton of the SO2 reductions realized by the 500 ppm 
fuel program can be compared to other available means to control 
SO2 emissions. The significant PM reductions are not 
accounted for in the relative cost per ton estimate, but are accounted 
for in our inventory analysis presented in section II and in the 
benefits analysis presented later in this section. Additional detail 
regarding all of the estimates presented here are available in the 
draft RIA.
    We are proposing a second step in the fuel program that would cap 
nonroad fuel sulfur levels at 15 ppm beginning in 2010. This fuel 
program enables the introduction of advanced emission control 
technologies including CDPFs and NOX adsorbers. The 
combination of the two-step fuel program and the new diesel engine 
standards represents the total Tier 4 program for nonroad diesel 
engines and fuel proposed today. In sections V.D.3 and 4, we present 
our estimate of the annual and total costs for

[[Page 28447]]

this complete program beginning in 2007 and continuing for 30 years. 
Also included is an estimate of the cost per ton of emissions 
reductions realized by this program for NMHC+NOX, PM, and 
SO2.
1. Annual Costs for the 500 ppm Fuel Program
    Cent per gallon costs for the proposed 500 ppm fuel program (i.e., 
the reduction to a 500 ppm sulfur cap) were presented in section V.A. 
Having this fuel would result in maintenance savings associated with 
increased oil change intervals for both the new and the existing fleet 
of nonroad, locomotive, and marine engines. These maintenance savings 
were discussed in section V.B. There are no engine and equipment costs 
associated with the 500 ppm fuel program because new emission standards 
are not part of that proposed program. Figure V.D-1 shows the annual 
costs associated with the 500 ppm fuel program.
    As can be seen in Figure V.D-1, the costs for refining and 
distributing the 500 ppm fuel range from $250 million in 2008 to nearly 
$400 million in 2036. These control costs are largely offset by the 
maintenance savings that range from $200 million in 2008 to $380 
million in 2036. Despite the fact that the costs of the 500 ppm fuel 
for nonroad diesel fuel is 2.5 cents/gallon and the maintenance savings 
are 3 cents per gallon, the net costs are positive because of the costs 
for the locomotive and marine fuel is not off-set by the maintenance 
savings. As a whole, the net cost of the program in each year is 
essentially zero, ranging from $50 million in the early years to only 
$18 million in 2036. The net present value of the net costs and savings 
associated with the proposed 500 ppm fuel program during the years 2007 
to 2036 is estimated at $510 million.
[GRAPHIC]
[TIFF OMITTED]
TP23MY03.009

2. Cost Per Ton for the 500 ppm Fuel Program
    The 2007 fuel program would result in large reductions of both 
SO2 and PM emissions. Roughly 98 percent of fuel sulfur is 
converted to SO2 in the engine with the remaining two 
percent being exhausted as sulfate PM. Because the majority of the 
emissions reductions associated with this program would be 
SOX, we have attributed all the control costs to 
SOX in calculating the cost per ton associated with this 
program. However, we have modeled both the SOX and PM 
reductions so that our inventory and benefits analysis fully account 
for them.
    As noted above, we have calculated both the costs and emission 
reductions of the 500 ppm fuel program as if it were to remain in place 
indefinitely. Figure V.D-1 shows the costs in each year of the program, 
the net present value of which is estimated at $510 million. We have 
estimated the 30 year net present value of the SOX emission 
reductions at 5.6 million tons.
    Table V.D-1 shows the cost per ton of emissions reduced as a result 
of the proposed 500 ppm fuel program. The cost per ton numbers include 
costs and emission reductions that would occur from both the new and 
the existing fleet (i.e., those pieces of nonroad equipment that were 
sold into the market prior to the proposed emission standards) of

[[Page 28448]]

nonroad, locomotive, and marine engines.

 Table V.D-1--500 ppm Fuel Program Aggregate Cost per Ton and Long-Term
                       Annual Cost per Ton ($2001)
------------------------------------------------------------------------
                                                 2004-2036
                                                 Discounted   Long-term
                   Pollutant                      lifetime     cost per
                                                  cost per   ton in 2036
                                                    ton
------------------------------------------------------------------------
SOX...........................................          $90          $50
------------------------------------------------------------------------

    We also considered the cost per ton of the 500 ppm fuel program 
without taking credit for the expected maintenance savings associated 
with low sulfur fuel. Without the maintenance savings, the cost per ton 
of SOX reduced would be $990 per ton for each year of the 
program. More detail on how the costs and cost per ton numbers 
associated with the 500 ppm fuel program were calculated can be found 
in the draft RIA.
3. Annual Costs for the Proposed Two-Step Fuel Program and Engine 
Program
    The costs of the total proposed engine and fuel program include 
costs associated with both steps in the fuel program--the reduction to 
500 ppm sulfur in 2007 and the reduction to 15 ppm sulfur in 2010. Also 
included are costs for the proposed 2008 engine standards for <75 
horsepower engines, the proposed 2013 standards for 25 to 75 horsepower 
engines, and costs for the proposed engine standards for £75 
horsepower engines. Included are all maintenance costs and savings 
realized by both the existing fleet (nonroad, locomotive, and marine) 
and the new fleet of engines complying with the proposed standards.
    Figure V.D-2 presents these results. All capital costs for fuel 
production and engine and equipment fixed costs have been amortized. 
The figure shows that total annual costs are estimated to be $120 
million in the first year the new engine standards apply, increasing to 
a peak of $1.7 billion in 2036 as increasing numbers of engines become 
subject to the new standards and an ever increasing amount of fuel is 
consumed. The net present value of the annualized costs over the period 
from 2007 to 2036 is $20.7 billion.
[GRAPHIC]
[TIFF OMITTED]
TP23MY03.010

4. Cost per Ton of Emissions Reduced for the Total Program
    We have calculated the cost per ton of emissions reduced associated 
with the proposed engine and fuel program. We have done this using the 
net present value of the annualized costs of the program through 2036 
and the net present value of the annual emission reductions through 
2036. We have also calculated the cost per ton of emissions in the year 
2036 using the annual costs

[[Page 28449]]

and emission reductions in that year alone. This number represents the 
long-term cost per ton of emissions reduced after all fixed costs of 
the program have been recovered by industry leaving only the variable 
costs of control. The cost per ton numbers include costs and emission 
reductions that would occur from the existing fleet (i.e., those pieces 
of nonroad equipment that were sold into the market prior to the 
proposed emission standards). These results are shown in Table V.D-2. 
We did the cost analysis using a 3% discount rate. We will also be 
conducting a similar analysis using a 7% discount rate and including 
this information in the docket.

 Table V.D-2--Total Proposed Fuel and Engine Program Aggregate Cost per
              Ton and Long-Term Annual Cost Per Ton ($2001)
------------------------------------------------------------------------
                                                 2004-2036
                                                 Discounted   Long-term
                   Pollutant                      lifetime     cost per
                                                  cost per   ton in 2036
                                                    ton
------------------------------------------------------------------------
NOX+NMHC......................................         $810         $530
PM............................................        8,700        6,900
SOX...........................................      \a\ 200         170
------------------------------------------------------------------------
Notes:
\a\ This result does not match that in Table 8.4-2 because the nonroad
  portion of the fuel is reduced to 15 ppm and does not stay at 500
  (locomotive and marine portions are kept at 500ppm). The costs to
  reduce fuel sulfur from uncontrolled to 15ppm were assigned 50/50 to
  NOX+NMHC and PM for the reduction to 15 ppm is to enable
  aftertreatment technology.

5. Comparison With Other Means of Reducing Emissions
    In comparison with other programs to control these pollutants, we 
believe that the proposed programs represent a cost effective strategy 
for generating substantial NOX+NMHC, PM, and SO2 
reductions. This can be seen by comparing the 2007 fuel program (i.e., 
a sulfur cap of 500 ppm) cost per ton and the total program cost per 
ton with a number of standards that EPA has adopted in the past. Table 
V.D-3 summarizes the cost per ton of several past EPA actions for 
NOX+NMHC. Table V.D-4 summarizes the cost per ton of several 
past EPA actions for PM.

 Table V.D-3--Cost per Ton of Previous Mobile Source Programs for NOX +
                                  NMHC
------------------------------------------------------------------------
                         Program                               $/ton
------------------------------------------------------------------------
Tier 2 Nonroad Diesel...................................       630
Tier 3 Nonroad Diesel...................................       430
Tier 2 vehicle/gasoline sulfur..........................  1,410-2,370
2007 Highway HD.........................................     2,260
2004 Highway HD.........................................   220-430
Off-highway diesel engine...............................   450-710
Tier 1 vehicle..........................................  2,160-2,930
NLEV....................................................      2030
Marine SI engines.......................................  1,230-1,940
On-board diagnostics....................................     2,430
Marine CI engines.......................................   30-190
------------------------------------------------------------------------
Note: Costs adjusted to 2001 dollars using the Producer Price Index for
  Total Manufacturing Industries.


  Table V.D-4.--Cost per Ton of Previous Mobile Source Programs for PM
------------------------------------------------------------------------
                        Program                               $/ton
------------------------------------------------------------------------
Tier 1/Tier 2 Nonroad Diesel...........................     2,410
2007 Highway HD........................................    14,280
Marine CI engines......................................  5,480-4,070
1996 urban bus.........................................  12,870-20,590
Urban bus retrofit/rebuild.............................    31,740
1994 highway HD diesel.................................  21,930-25,670
------------------------------------------------------------------------
Note: Costs adjusted to 2001 dollars using the Producer Price Index for
  Total Manufacturing Industries.

    To compare the cost per ton of SO2 emissions reduced, we 
looked at the cost per ton for the Title IV SO2 trading 
programs. This information is found in EPA report 430/R-02-004, 
``Documentation of EPA Modeling Applications (V.2.1) Using the 
Integrated Planning Model'', in Figure 9.11 on page 9-14 (
www.epa.gov/airmarkets/epa-ipm/index.html#documentation). The SO2 cost 
per ton results of the proposed program presented in Table V.D-2 
compare very favorably with the program shown in Table V.D-5.

 Table V.D-5--Cost per Ton of SO2 From EPA Base Case 2000 for the Title
                         IV SO2 Trading Programs
------------------------------------------------------------------------
                  Program                               $/ton
------------------------------------------------------------------------
Title IV SO2 Trading Programs.............  $490 in 2010 to $610 in
                                             2020.
------------------------------------------------------------------------
Note: Costs adjusted to 2001 dollars using the Producer Price Index for
  Total Manufacturing Industries.

E. Do the Benefits Outweigh the Costs of the Standards?

    Our analysis of the health and welfare benefits to be expected from 
this proposal are presented in this section. Briefly, the analysis 
projects major benefits throughout the period from initial 
implementation of the rule through 2030, the last year analyzed. As 
described below, thousands of deaths and other serious health effects 
would be prevented, yielding a net present value in 2004 of those 
benefits we could monetize of approximately $550 billion dollars. These 
benefits exceed the net present value of the social cost of the 
proposal ($17 billion) by a factor of over 30 to one.
1. What Were the Results of the Benefit-Cost Analysis?
    Table V.E-1 presents the primary estimate of reduced incidence of 
PM-related health effects for the years 2020 and 2030. In interpreting 
the results, it is important to keep in mind the limited set of effects 
we are able to monetize. Specifically, the table lists the PM-related 
benefits associated with the reduction of several health effects.\290\ 
In 2030, we estimate that there will be 9,600 fewer fatalities per year 
associated with fine PM, and the rule will result in about 5,700 fewer 
cases of chronic bronchitis, 8,300 fewer hospitalizations (for 
respiratory and cardiovascular disease combined), and result in 
significant reductions in days of restricted activity due to 
respiratory illness (with an estimated 5.7 million fewer cases). We 
also estimate substantial health improvements for children from reduced 
upper and lower respiratory illness, acute bronchitis, and asthma 
attacks.\291\
---------------------------------------------------------------------------

    \290\ Based upon recent preliminary findings by the Health 
Effects Institute, the concentration-response functions used to 
estimate reductions in hospital admissions may over or underestimate 
the true concentration-response relationship. See letter from Dan 
Greenberg, President, Health Effects Institute, May 30, 2002, 
attached to letter from Dr. Hopke, dated August 8, 2002. Docket A-
2000-01, Document IV-A-145.
    \291\ Our estimate incorporates significant reductions of 
150,000 fewer cases of lower respiratory symptoms in children ages 7 
to 14 each year, 110,000 fewer cases of upper respiratory symptoms 
(similar to cold symptoms) in asthmatic children each year, and 
14,000 fewer cases of acute bronchitis in children ages 8 to 12 each 
year. In addition, we estimate that this rule will reduce almost 
6,000 emergency room visits for asthma attacks in children each year 
from reduced exposure to particles. Additional incidents would be 
avoided from reduced ozone exposures. Asthma is the most prevalent 
chronic disease among children and currently affects over seven 
percent of children under 18 years of age.
---------------------------------------------------------------------------

    Table V.E-2 presents the total monetized benefits for the years 
2020 and 2030. This table also indicates with a ``B'' those additional 
health and environmental effects which we were unable to quantify or 
monetize. These effects are additive to estimate of total benefits, and 
EPA believes there is

[[Page 28450]]

considerable value to the public of the benefits that could not be 
monetized. A full listing of the benefit categories that could not be 
quantified or monetized in our estimate are provided in Table V.E-5.
    In summary, EPA's primary estimate of the benefits of the rule are 
approximately $81 + B billion in 2030. In 2020, total monetized 
benefits are approximately $43 + B billion. These estimates account for 
growth in real gross domestic product (GDP) per capita between the 
present and the years 2020 and 2030. As the table indicates, total 
benefits are driven primarily by the reduction in premature fatalities 
each year, which account for over 90 percent of total benefits.

   Table V.E-1.--Reductions in Incidence of PM-Related Adverse Health
   Effects Associated With the Proposed Nonroad Diesel Engine and Fuel
                                Standards
------------------------------------------------------------------------
                                          Avoided incidence \a\  (cases/
                                                       year)
                Endpoint                 -------------------------------
                                               2020            2030
------------------------------------------------------------------------
Premature mortality \b\--Base estimate:            5,200           9,600
 Long-term exposure (adults, 30 and
 over)..................................
Chronic bronchitis (adults, 26 and over)           3,600           5,700
Non-fatal myocardial infarctions                   9,200          16,000
 (adults, 18 and older).................
Hospital admissions--Respiratory                   2,400           4,500
 (adults, 20 and older) \c\.............
Hospital admissions--Cardiovascular                1,900           3,800
 (adults, 20 and older) \d\.............
Emergency Room Visits for Asthma (18 and           3,600           5,700
 younger)...............................
Acute bronchitis (children, 8-12).......           8,400          14,000
Lower respiratory symptoms (children, 7-          92,000         150,000
 14)....................................
Upper respiratory symptoms (asthmatic             77,000         110,000
 children, 9-11)........................
Work loss days (adults, 18-65)..........         650,000         960,000
Minor restricted activity days (adults,        3,900,000      5,700,000
 age 18-65).............................
------------------------------------------------------------------------
Notes:
\a\ Incidences are rounded to two significant digits.
\b\ Premature mortality associated with ozone is not separately included
  in this analysis
\c\ Respiratory hospital admissions for PM includes admissions for COPD,
  pneumonia, and asthma.
\d\ Cardiovascular hospital admissions for PM includes total
  cardiovascular and subcategories for ischemic heart disease,
  dysrhythmias, and heart failure.

     Table V.E-2.--EPA Primary Estimate of the Annual Quantified and
  Monetized Benefits Associated With Improved PM Air Quality Resulting
       From the Proposed Nonroad Diesel Engine and Fuel Standards
------------------------------------------------------------------------
                                             Monetary Benefits\a,\ \b\
                                           (millions 2000$, adjusted for
                Endpoint                          income growth)
                                         -------------------------------
                                               2020            2030
------------------------------------------------------------------------
Premature mortality \c\ Long-term                $39,000         $74,000
 exposure (adults, 30 and over).........
Chronic bronchitis (WTP valuation;                 1,600           2,600
 adults, 26 and over)...................
Non-fatal myocardial infarctions........             750           1,300
Hospital Admissions from Respiratory                  38              74
 Causes \d\.............................
Hospital Admissions from Cardiovascular               40              80
 Causes \e\.............................
Emergency Room Visits for Asthma........               1               2
Acute bronchitis (children, 8-12).......               3               5
Lower respiratory symptoms (children, 7-               2               3
 14)....................................
Upper respiratory symptoms (asthmatic                  2               3
 children, 9-11)........................
Work loss days (adults, 18-65)..........              90             130
Minor restricted activity days (adults,              210             320
 age 18-65).............................
Recreational visibility (86 Class I                1,200           1,900
 Areas).................................
                                         -----------------
    Total Monetized Benefits \f\........      43,000 + B     81,000 + B
------------------------------------------------------------------------
Notes:
\a\ Monetary benefits are rounded to two significant digits.
\b\ Monetary benefits are adjusted to account for growth in real GDP per
  capita between 1990 and the analysis year (2020 or 2030).
\c\ Valuation assumes the 5 year distributed lag structure described
  earlier. Results reflect the use of two different discount rates; a 3%
  rate which is recommended by EPA's Guidelines for Preparing Economic
  Analyses (US EPA, 2000a), and 7% which is recommended by OMB Circular
  A-94 (OMB, 1992).
\d\ Respiratory hospital admissions for PM includes admissions for COPD,
  pneumonia, and asthma.
\e\ Cardiovascular hospital admissions for PM includes total
  cardiovascular and subcategories for ischemic heart disease,
  dysrhythmias, and heart failure.
\f\ B represents the monetary value of the unmonetized health and
  welfare benefits. A detailed listing of unquantified PM, ozone, CO,
  and NMHC related health effects is provided in Table V.E-5.

    The estimated social cost (measured as changes in consumer and 
producer surplus) in 2030 to implement the final rule from Table V.F-2 
is $1.5 billion (2000$). Thus, the net benefit (social benefits minus 
social costs) of the program at full implementation is approximately 
$79 + B billion. In 2020, partial implementation of the program yields 
net benefits of $42 + B billion. Therefore, implementation of the final 
rule is expected to provide society with a net gain in social welfare 
based on economic efficiency criteria. Table V.E-3 presents a summary 
of the benefits,

[[Page 28451]]

costs, and net benefits of the proposed rule. Figure VE.1 displays the 
stream of benefits, costs, and net benefits of the Nonroad Land-based 
Diesel Vehicle Rule from 2007 to 2030. In addition, Table V-E.4 
presents the net present value of the stream of benefits, costs, and 
net benefits associated with the rule for this 23 year period (using a 
three percent discount rate). The total net present value in 2004 of 
the stream of net benefits (benefits minus costs) is $530 billion.

    Table V.E-3.--Summary of Benefits, Costs, and Net Benefits of the Proposed Nonroad Diesel Engine and Fuel
                                                    Standards
----------------------------------------------------------------------------------------------------------------
                                            2020 \a\  (billions of 2000          2030 \a\  (billions of 2000
                                                      dollars)                             dollars)
----------------------------------------------------------------------------------------------------------------
Social Costs \b\......................  $1.4...............................  $1.5.
Social Benefits \b,\ \c,\ \d\:
    CO, VOC, Air Toxic-related          Not monetized......................  Not monetized.
     benefits.
    Ozone-related benefits............  Not monetized......................  Not monetized.
    PM-related Welfare benefits.......  $1.2...............................  $1.9.
    PM-related Health benefits........  $42+ B.............................  $79 + B.
    Net Benefits (Benefits-Costs) \c\.  $42 + B............................  $79 + B.
----------------------------------------------------------------------------------------------------------------
Notes:
\a\ All costs and benefits are rounded to two significant digits.
\b\ Note that costs are the total costs of reducing all pollutants, including CO, VOCs and air toxics, as well
  as NOX and PM. Benefits in this table are associated only with PM, NOX and SO3 reductions.
\c\ Not all possible benefits or disbenefits are quantified and monetized in this analysis. Potential benefit
  categories that have not been quantified and monetized are listed in Table V.E-5. B is the sum of all
  unquantified benefits and disbenefits.

[[Page 28452]]
[GRAPHIC]
[TIFF OMITTED]
TP23MY03.011


   Table V.E-4.--Net Present Value in 2004 of the Stream of Benefits,
 Costs, and Net Benefits for the Proposed Nonroad Diesel Engine and Fuel
                                Standards
                           [Billions of 2000$]
------------------------------------------------------------------------

------------------------------------------------------------------------
Social Costs............................................             $17
Social Benefits.........................................             550
Net Benefits............................................         \a\ 530
------------------------------------------------------------------------
Notes:
\a\ Numbers do not add due to rounding.

2. What Was Our Overall Approach to the Benefit-Cost Analysis?
    The basic question we sought to answer in the benefit-cost analysis 
was, ``What are the net yearly economic benefits to society of the 
reduction in mobile source emissions likely to be achieved by this 
proposed rulemaking?'' In designing an analysis to address this 
question, we selected two future years for analysis (2020 and 2030) 
that are representative of the stream of benefits and costs at partial 
and full-implementation of the program.
    To quantify benefits, we evaluated PM-related health effects 
(including directly emitted PM, SO3, and NOX 
contributions to fine particulate matter). Our approach requires the 
estimation of changes in air quality expected from the rule and then 
estimating the resulting impact on health. In order to characterize the 
benefits of today's action, given the constraints on time and resources 
available for the analysis, we adopted a benefits transfer technique 
that relies on air quality and benefits modeling for a preliminary 
control option for nonroad diesel engines and fuels. Results from the 
modeled preliminary control option in 2020 and 2030 are then scaled and 
transferred to the emission reductions expected from the proposed rule. 
We also transferred modeled results by using scaling factors associated 
with time to examine the stream of benefits in years other than 2020 
and 2030.
    More specifically, our health benefits assessment is conducted in 
two phases. Due to the time requirements for running the sophisticated 
emissions and air quality models needed to obtain estimates of the 
benefits expected to result from implementation of the rule, it is 
often necessary to select an example set of emission reductions to use 
for the purposes of emissions and air quality modeling. In phase one, 
we evaluate the PM and ozone related health effects associated with a 
modeled preliminary control option that was a close approximation of 
the proposed standards in the years 2020 and 2030. Using information 
from the modeled preliminary control option on the changes in ambient 
concentrations of PM and ozone, we then conduct a

[[Page 28453]]

health assessment to estimate the number of reduced incidences of 
illnesses, hospitalizations, and premature fatalities associated with 
this scenario and estimate the total economic value of these health 
benefits. The standards we are proposing in this rulemaking, however, 
are slightly different in the amount of emission reductions expected to 
be achieved in 2020 and 2030 relative to the modeled scenario. Thus, in 
phase two of the analysis we apportion the results of the phase one 
analysis to the underlying NOX, SO3, and PM 
emission reductions and scale the apportioned benefits to reflect 
differences in emissions reductions between the modeled preliminary 
control option and the proposed standards. The sum of the scaled 
benefits for the PM, SO3, and NOX emission 
reductions provide us with the total benefits of the rule.
    The benefit estimates derived from the modeled preliminary control 
option in phase one of our analysis uses an analytical structure and 
sequence similar to that used in the benefits analyses for the Heavy 
Duty Engine/Diesel Fuel final rule and in the ``section 812 studies'' 
to estimate the total benefits and costs of the full Clean Air 
Act.\292\ We used many of the same models and assumptions used in the 
Heavy Duty Engine/Diesel Fuel analysis as well as other Regulatory 
Impact Analyses (RIAs) prepared by the Office of Air and Radiation. By 
adopting the major design elements, models, and assumptions developed 
for the section 812 studies and other RIAs, we have largely relied on 
methods which have already received extensive review by the independent 
Science Advisory Board (SAB), by the public, and by other federal 
agencies. In addition, we will be working through the next section 812 
study process to enhance our methods.\293\ Interested parties will 
therefore be able to obtain further information from the section 812 
study on the kinds of methods we are likely to use for estimating 
benefits and costs in the final nonroad diesel rule.
---------------------------------------------------------------------------

    \292\ The section 812 studies include: (1) US EPA, Report to 
Congress: The Benefits and Costs of the Clean Air Act, 1970 to 1990, 
October 1997 (also known as the ``Section 812 Retrospective 
Report''); and (2) the first in the ongoing series of prospective 
studies estimating the total costs and benefits of the Clean Air Act 
(see EPA report number: EPA-410-R-99-001, November 1999). See Docket 
A-99-06, Document II-A-21.
    \293\ We anticipate a public SAB meeting June 11-13, 2003, in 
Washington, DC, regarding components of our analytical blueprint. 
Interested parties may want to consult the Web page: 
http://www.epa.gov/science1.
---------------------------------------------------------------------------

    The benefits transfer method used in phase two of the analysis is 
similar to that used to estimate benefits in the recent analysis of the 
Nonroad Large Spark-Ignition Engines and Recreational Engines standards 
(67 FR 68241, November 8, 2002). A similar method has also been used in 
recent benefits analyses for the proposed Industrial Boilers and 
Process Heaters NESHAP and the Reciprocating Internal Combustion 
Engines NESHAP.
    On September 26, 2002, the National Academy of Sciences (NAS) 
released a report on its review of the Agency's methodology for 
analyzing the health benefits of measures taken to reduce air 
pollution. The report focused on EPA's approach for estimating the 
health benefits of regulations designed to reduce concentrations of 
airborne particulate matter (PM).
    In its report, the NAS said that EPA has generally used a 
reasonable framework for analyzing the health benefits of PM-control 
measures. It recommended, however, that the Agency take a number of 
steps to improve its benefits analysis. In particular, the NAS stated 
that the Agency should:
    ? Include benefits estimates for a range of regulatory 
options;
    ? Estimate benefits for intervals, such as every five years, 
rather than a single year;
    ? Clearly state the projected baseline statistics used in 
estimating health benefits, including those for air emissions, air 
quality, and health outcomes;
    ? Examine whether implementation of proposed regulations 
might cause unintended impacts on human health or the environment;
    ? When appropriate, use data from non-U.S. studies to broaden 
age ranges to which current estimates apply and to include more types 
of relevant health outcomes;
    ? Begin to move the assessment of uncertainties from its 
ancillary analyses into its Base analyses by conducting probabilistic, 
multiple-source uncertainty analyses. This assessment should be based 
on available data and expert judgment.
    Although the NAS made a number of recommendations for improvement 
in EPA's approach, it found that the studies selected by EPA for use in 
its benefits analysis were generally reasonable choices. In particular, 
the NAS agreed with EPA's decision to use cohort studies to derive 
benefits estimates. It also concluded that the Agency's selection of 
the American Cancer Society (ACS) study for the evaluation of PM-
related premature mortality was reasonable, although it noted the 
publication of new cohort studies that should be evaluated by the 
Agency.
    EPA has addressed many of the NAS comments in our analysis of the 
proposed rule. We provide benefits estimates for each year over the 
rule implementation period for a wide range of regulatory alternatives, 
in addition to our proposed emission control program. We use the 
estimated time path of benefits and costs to calculate the net present 
value of benefits of the rule. In the RIA, we provide baseline 
statistics for air emissions, air quality, population, and health 
outcomes. We have examined how our benefits estimates might be impacted 
by expanding the age ranges to which epidemiological studies are 
applied, and we have added several new health endpoints, including non-
fatal heart attacks, which are supported by both U.S. studies and 
studies conducted in Europe. We have also improved the documentation of 
our methods and provided additional details about model assumptions.
    Several of the NAS recommendations addressed the issue of 
uncertainty and how the Agency can better analyze and communicate the 
uncertainties associated with its benefits assessments. In particular, 
the Committee expressed concern about the Agency's reliance on a single 
value from its analysis and suggested that EPA develop a probabilistic 
approach for analyzing the health benefits of proposed regulatory 
actions. The Agency agrees with this suggestion and is working to 
develop such an approach for use in future rulemakings. EPA plans to 
hold a meeting of its Science Advisory Board (SAB) in early Summer 2003 
to review its plans for addressing uncertainty in its analyses. Our 
likely approach will incorporate short-term elements intended to 
provide interim methods in time for the final Nonroad rule to address 
uncertainty in important analytical parameters such as the 
concentration-response relationship for PM-related premature mortality. 
Our approach will also include longer-term elements intended to provide 
scientifically sound, peer-reviewed characterizations of the 
uncertainty surrounding a broader set of analytical parameters and 
assumptions, including but not limited to emissions and air quality 
modeling, demographic projections, population health status, 
concentration-response functions, and valuation estimates.
3. What Are the Significant Limitations of the Benefit-Cost Analysis?
    Every benefit-cost analysis examining the potential effects of a 
change in

[[Page 28454]]

environmental protection requirements is limited to some extent by data 
gaps, limitations in model capabilities (such as geographic coverage), 
and uncertainties in the underlying scientific and economic studies 
used to configure the benefit and cost models. Deficiencies in the 
scientific literature often result in the inability to estimate 
quantitative changes in health and environmental effects, such as 
potential increases in premature mortality associated with increased 
exposure to carbon monoxide. Deficiencies in the economics literature 
often result in the inability to assign economic values even to those 
health and environmental outcomes which can be quantified. While these 
general uncertainties in the underlying scientific and economics 
literatures, which can cause the valuations to be higher or lower, are 
discussed in detail in the Regulatory Support Document and its 
supporting documents and references, the key uncertainties which have a 
bearing on the results of the benefit-cost analysis of this final rule 
include the following:
    ? The exclusion of potentially significant benefit categories 
(such as health and ecological benefits of reduction in CO, VOCs, air 
toxics, and ozone);
    ? Errors in measurement and projection for variables such as 
population growth;
    ? Uncertainties in the estimation of future year emissions 
inventories and air quality;
    ? Uncertainties associated with the scaling of the results of 
the modeled benefits analysis to the proposed standards, especially 
regarding the assumption of similarity in geographic distribution 
between emissions and human populations and years of analysis;
    ? Variability in the estimated relationships of health and 
welfare effects to changes in pollutant concentrations;
    ? Uncertainties in exposure estimation;
    ? Uncertainties associated with the effect of potential 
future actions to limit emissions.
    Despite these uncertainties, we believe the benefit-cost analysis 
provides a reasonable indication of the expected economic benefits of 
the proposed rulemaking in future years under a set of assumptions.
    One significant limitation to the benefit transfer method applied 
in this analysis is the inability to scale ozone-related benefits. 
Because ozone is a homogeneous gaseous pollutant, it is not possible to 
apportion ozone benefits to the precursor emissions of NOX 
and VOC. Coupled with the potential for NOX reductions to 
either increase or decrease ambient ozone levels, this prevents us from 
scaling the benefits associated with a particular combination of VOC 
and NOX emissions reductions to another. Because of our 
inability to scale ozone benefits, we do not include ozone benefits as 
part of the monetized benefits of the proposed standards. For the most 
part, ozone benefits contribute substantially less to the monetized 
benefits than do benefits from PM, thus their omission will not 
materially affect the conclusions of the benefits analysis. Although we 
expect economic benefits to exist, we were unable to quantify or to 
value specific changes in ozone, CO or air toxics because we did not 
perform additional air quality modeling.
    There are also a number of health and environmental effects which 
we were unable to quantify or monetize. A full appreciation of the 
overall economic consequences of the proposed rule requires 
consideration of all benefits and costs expected to result from the new 
standards, not just those benefits and costs which could be expressed 
here in dollar terms. A complete listing of the benefit categories that 
could not be quantified or monetized in our estimate are provided in 
Table V.E-5. These effects are denoted by ``B'' in Table V.E-3 above, 
and are additive to the estimates of benefits.

Table V.E-5.--Additional, Non-monetized Benefits of the Proposed Nonroad
                    Diesel Engine and Fuel Standards
------------------------------------------------------------------------
          Pollutant                       Unquantified effects
------------------------------------------------------------------------
Ozone Health.................  Premature mortality.\a\
                               Increased airway responsiveness to
                                stimuli.
                               Inflammation in the lung.
                               Chronic respiratory damage.
                               Premature aging of the lungs.
                               Acute inflammation and respiratory cell
                                damage.
                               Increased susceptibility to respiratory
                                infection.
                               Non-asthma respiratory emergency room
                                visits.
                               Increased school absence rates.
Ozone Welfare................  Decreased yields for commercial forests
                                (for example, Western US).
                               Decreased yields for fruits and
                                vegetables.
                               Decreased yields for non-commercial
                                crops.
                               Damage to urban ornamental plants.
                               Impacts on recreational demand from
                                damaged forest aesthetics.
                               Damage to ecosystem functions.
PM Health....................  Infant mortality.
                               Low birth weight.
                               Changes in pulmonary function.
                               Chronic respiratory diseases other than
                                chronic bronchitis.
                               Morphological changes.
                               Altered host defense mechanisms.
                               Cancer.
                               Non-asthma respiratory emergency room
                                visits.
PM Welfare...................  Visibility in many Class I areas.
                               Residential and recreational visibility
                                in non-Class I areas.
                               Soiling and materials damage.
                               Damage to ecosystem functions.

[[Page 28455]]

Nitrogen and Sulfate           Impacts of acidic sulfate and nitrate
 Deposition Welfare.            deposition on commercial forests.
                               Impacts of acidic deposition to
                                commercial freshwater fishing.
                               Impacts of acidic deposition to
                                recreation in terrestrial ecosystems.
                               Reduced existence values for currently
                                healthy ecosystems.
                               Impacts of nitrogen deposition on
                                commercial fishing, agriculture, and
                                forests.
                               Impacts of nitrogen deposition on
                                recreation in estuarine ecosystems.
                               Damage to ecosystem functions.
CO Health....................  Premature mortality.\a\
                               Behavioral effects.
HC Health \b\................  Cancer (benzene, 1,3-butadiene,
                                formaldehyde, acetaldehyde).
HC Welfare...................  Direct toxic effects to animals.
                               Bioaccumulation in the food chain.
                               Damage to ecosystem function.
                               Odor.
------------------------------------------------------------------------
Notes:
\a\ Premature mortality associated with ozone and carbon monoxide is not
  separately included in this analysis. In this analysis, we assume that
  the ACS/Krewski, et al. C-R function for premature mortality captures
  both PM mortality benefits and any mortality benefits associated with
  other air pollutants. A copy of Krewski, et al., can be found in
  Docket A-99-06, Document No. IV-G-75.
\b\ Many of the key hydrocarbons related to this rule are also hazardous
  air pollutants listed in the Clean Air Act.

F. Economic Impact Analysis

    An Economic Impact Analysis (EIA) was prepared to estimate the 
economic impacts of this proposal on producers and consumers of nonroad 
engines and equipment and related industries. The Nonroad Diesel 
Economic Impact Model (NDEIM), developed for this analysis, was used to 
estimate market-level changes in price and outputs for affected engine, 
equipment, fuel, and application markets as well as the social costs 
and their distribution across economic sectors affected by the program. 
This section presents the results of the economic impact analysis. A 
detailed description of the NDEIM, the model inputs, and several 
sensitivity analyses can be found in chapter 10 of the Draft Regulatory 
Impact Analysis prepared for this proposal.
1. What Is an Economic Impact Analysis?
    Regulatory agencies conduct economic impact analyses of potential 
regulatory actions to inform decision makers about the effects of a 
proposed regulation on society's current and future well-being. In 
addition to informing decision makers within the Agency, economic 
impact analyses are conducted to meet the statutory and administrative 
requirements imposed by Congress and the Executive office. The Clean 
Air Act requires an economic impact analysis under section 317, while 
Executive Order 12866--Regulatory Planning and Review requires 
Executive Branch agencies to perform benefit-costs analyses of all 
rules it deems to be ``significant'' (typically over $100 million 
annual social costs) and submit these analyses to the Office of 
Management and Budget (OMB) for review. This economic impact analysis 
estimates the potential market impacts of the proposed rule's 
compliance costs and provides the associated social costs and their 
distribution across stakeholders for comparison with social benefits 
(as presented in Section V.E).
2. What Is EPA's Economic Analysis Approach for This Proposal?
    The underlying objective of an EIA is to evaluate the effect of a 
proposed regulation on the welfare of affected stakeholders and society 
in general. Using information on the expected compliance costs of the 
proposed program as presented in the preceding discussion, this EIA 
explores how the companies that produce nonroad diesel engines, 
equipment, or fuel may change their production behavior in response to 
the costs of complying with the standards. It also explores how the 
consumers who use the affected products may change their purchasing 
decisions. For example, the construction industry may reduce purchases 
if the prices of nonroad diesel equipment increase, thereby reducing 
the volume of equipment sold (or market demand) for such equipment. 
Alternatively, the construction industry may pass along these 
additional costs to the consumers of their final goods and services by 
increasing prices, which would mitigate the potential impacts on the 
purchases of nonroad diesel equipment.
    The conceptual approach of the NDEIM is to link significantly 
affected markets to mimic how compliance costs will potentially ripple 
through the economy. The compliance costs will be directly borne by 
engine manufacturers, equipment manufacturers, and petroleum 
refineries. Depending on market characteristics, some or all of these 
compliance costs will be passed on through the supply chain in the form 
of higher prices extending to producers and consumers in the 
application markets (i.e., construction, agriculture, and 
manufacturing). The NDEIM explicitly models these linkages and 
estimates behavioral responses that lead to new equilibrium prices and 
output for all related markets and the resulting distribution of costs 
across stakeholders.
    The NDEIM uses a multi-market partial equilibrium approach to track 
changes in price and quantity for 60 integrated product markets, as 
follows:
    ? 7 diesel engine markets (less than 25 hp, 26 to 50 hp, 51 
to 75 hp, 76 to 100 hp, 101 to 175 hp, 176 to 600 hp, and greater than 
600 hp; the EIA includes more horsepower categories than the standards, 
allowing more efficient use of the engine compliance cost estimates 
developed for this proposal).
    ? 42 diesel equipment markets (7 horsepower categories within 
7 application categories: agricultural, construction, general 
industrial, pumps and compressors, generator and welder sets, 
refrigeration and air conditioning, and lawn and garden; there are 7 
horsepower/application categories that did not have sales in 2000 and 
are not included in the model, so the total number of diesel equipment 
markets is 42 rather than 49).
    ? 3 application markets (agricultural, construction, and 
manufacturing).
    ? 8 nonroad diesel fuel markets (2 sulfur content levels of 
15 ppm and 500 ppm for each of 4 PADDs; PADDs 1 and

[[Page 28456]]

3 are combined for the purpose of this analysis). It should be noted 
that PADD 5 includes Alaska and Hawaii. Because those two states are 
geographically separate from the rest of PADD 5, we seek comment on 
whether they should be considered as separate fuel markets.
    The NDEIM uses an intermediate run time frame and assumes perfect 
competition in the market sectors. It is a computer model comprised of 
a series of spreadsheet modules that define the baseline 
characteristics of the supply and demand for the relevant markets and 
the relationships between them. A detailed description of the model 
methodology, inputs, and parameters is provided in chapter 10 of the 
draft RIA prepared for this proposal. The model methodology is firmly 
rooted in applied microeconomic theory and was developed following the 
OAQPS Economic Analysis Resource Document.\294\ Based on the specified 
market linkages, the model is shocked by applying the engineering 
compliance cost estimates to the appropriate market suppliers and then 
numerically solved using an iterative auctioneer approach by ``calling 
out'' new prices until a new equilibrium is reached in all markets 
simultaneously.
---------------------------------------------------------------------------

    \294\ U.S. Environmental Protection Agency, Office of Air 
Quality Planning and Standards, Innovative Strategies and Economics 
Group, OAQPS Economic Analysis Resource Document, April 1999. A copy 
of this document can be found in Docket A-2001-28, Document No. II-
A-14.
---------------------------------------------------------------------------

    The actual economic impacts of the proposed rule will be determined 
by the ways in which producers and consumers of the engines, equipment, 
and fuels affected by the proposal change their behavior in response to 
the costs incurred in complying with the standards. In the NDEIM, these 
behaviors are modeled by the demand and supply elasticities. The supply 
elasticities for the engine and equipment markets and the demand 
elasticities for the application markets were estimated using 
econometric methods. The procedures and results are reported in 
Appendix 10.1 of the draft RIA. Literature-based estimates were used 
for the supply elasticities in the application and fuel markets.
    There are two ways to handle the demand elasticities for the 
engine, equipment, and fuel markets. In the approach used in NDEIM, 
these demand elasticities are internally derived based on the specified 
market linkages, i.e., the demand for engines, equipment, and fuel are 
modeled as directly related to the supply and demand of goods and 
services supplied by the final application markets. In other words, the 
supply of those goods and services determines the demand for equipment 
and fuel, and the supply of equipment determines the demand for 
engines. Using this approach, the NDEIM predicts that engine and 
equipment production will decrease by only a small amount: 0.013% and 
0.014% respectively (see Table V.F-1). Also, please see draft RIA 
Appendices 10A and 10B for more detailed estimates on the price 
increase estimates. Because the application markets are modeled with 
inelastic or unit elastic demand and supply elasticities (quantity 
supplied/demanded is expected to be fairly insensitive to price changes 
or they will vary directly with price changes), the model predicts that 
engine and equipment manufacturers will pass along virtually all of 
their costs to end users.
    An alternative approach could be used in which the demand 
elasticities for the equipment, engine, and fuel markets are not 
derived as part of the model. They could be estimated separately or a 
sensitivity analysis could be conducted that assumes more elastic 
values than those generated by the NDEIM. We are continuing to 
investigate this matter and will be placing additional information 
about elasticities in the docket during the comment period for this 
rule. We request comment on that information as well as on the 
methodology and other aspects of this EIA.
    The estimated engine and equipment market impacts are based solely 
on the expected increase in variable costs associated with the proposed 
standards. Fixed costs associated with the engine emission standards 
are not included in the market analysis reported in Table IV-F-1. This 
is because in an analysis of competitive markets the industry supply 
curve is based on its marginal cost curve, and fixed costs are not 
reflected in changes in the marginal cost curve. In addition, fixed 
costs are primarily R&D costs associated with design and engineering 
changes, and firms in the affected industries currently allocate funds 
for these costs. Therefore, fixed costs are not likely to affect the 
prices of engines or equipment. This assumption is described in greater 
detail in section 10.2 of the draft RIA. R&D costs are a long-run 
concern and decisions to invest or not invest in R&D are made in the 
long run. If funds have to be diverted from some other activity into 
R&D needed to meet the environmental regulations, then these costs 
represent a component of the social costs of the rule. Therefore, fixed 
costs are included in the welfare impact estimates reported in Table 
V.F-2 as additional costs on producers. We also performed a sensitivity 
analysis, included in chapter 10 of the draft RIA for this proposal, 
that includes fixed costs as part of the model. This results in a 
transfer of welfare losses from engine and equipment markets to the 
application markets, but does not change the overall welfare losses 
associated with the proposal.
    Economic theory indicates that, in the long run, prices are 
expected to reflect the average total costs of the marginal producer in 
a market and not just variable costs. This suggests that it may be 
necessary to treat fixed costs differently for a long-run analysis. We 
will continue to investigate this effect and intend to place additional 
information in the docket during the comment period for this rule. We 
request comment on that information as well as on how fixed costs and 
R&D expenditures are handled in the NDEIM.
    In addition to the variable and fixed costs described above, there 
are three additional costs components that are included in the total 
social cost estimates of the proposed regulation but that are not 
explicitly included in the NDEIM. These are operating savings (costs), 
fuel marker costs, and spillover from 15 ppm fuel to higher sulfur 
fuel. We request comment on how best to incorporate each of these costs 
in the analysis.
    Operating savings (costs) refers to changes in operating costs that 
are expected to be realized by users of both existing and new nonroad 
diesel equipment as a result of the reduced sulfur content of nonroad 
diesel fuel. These include operating savings (cost reductions) due to 
fewer oil changes, which accrue to nonroad engines, and marine and 
locomotive engines, that are already in use as well as new nonroad 
engines that will comply with the proposed standards (see section 
V.B.). These savings (costs) also include any extra operating costs 
associated with the new PM emission control technology which may accrue 
to new engines that use this new technology. These savings (costs) are 
not included directly in the model because some of the savings accrue 
to existing engines and because these savings (costs) are not expected 
to affect consumer decisions with respect to new engines. Instead, they 
are added into the estimated welfare impacts as additional costs to the 
application markets, since it is the users of these engines that will 
see these savings (costs). Nevertheless, a sensitivity analysis was 
also performed in which these savings (costs) are included as inputs to 
the NDEIM, where they are modeled as benefits accruing to the 
application producers. The results of

[[Page 28457]]

this analysis are presented in Chapter 10 of the draft RIA.
    Fuel marker costs refers to costs associated with marking high 
sulfur diesel fuel in the locomotive, marine, and heating oil markets 
between 2007 and 2014. Marker costs are not included in the market 
analysis because locomotive, marine, and heating oil markets are not 
explicitly modeled in the NDEIM. Similar to the operating savings 
(costs), marker costs are added into the estimated welfare impacts 
separately.
    The costs of fuel that spills over from the 15 ppm market to higher 
grade sulfur fuel are also not included in the NDEIM but, instead, are 
added into the estimated welfare impacts separately. As described in 
section IV above, refiners are expected to produce more 15 ppm fuel 
than is required for the nonroad diesel fuel market. This excess 15 ppm 
fuel will be sold into markets that allow fuel with a higher sulfur 
level (e.g., locomotive, marine diesel, or home heating fuel). Because 
this spillover fuel will meet the 15 ppm limit, it is necessary to 
count the costs of sulfur reduction processes against those fuels.
    Consistent with the engine and equipment cost discussion in section 
V.C. of this preamble, the EIA does not include any cost savings 
associated with the proposed equipment transition flexibility program 
or the proposed nonroad engine ABT program. As a result, the results of 
this EIA can be viewed as somewhat conservative, in this respect.
3. What Are the Results of this Analysis?
    The economic analysis consists of two parts: a market analysis and 
welfare analysis. The market analysis looks at expected changes in 
prices and quantities for directly and indirectly affected market 
commodities. The welfare analysis looks at economic impacts in terms of 
annual and present value changes in social costs. For this proposed 
rule, the social costs are computed as the sum of market surplus offset 
by operating cost savings. Market surplus is equal to the aggregate 
change in consumer and producer surplus based on the estimated market 
impacts associated with the proposed rule. Operating cost savings are 
associated with the decreased sulfur content of diesel fuel. These 
include maintenance savings (cost reductions) and changes in fuel 
efficiency. Increased maintenance costs may also be incurred for some 
technologies. Operating costs are not included in the market analysis 
but are instead listed as a separate category in the social cost 
results tables.
    Economic impact results for 2013, 2020, and 2030 are presented in 
this section. The first of these years, 2013, corresponds to the first 
year in which the standards affect all engines, equipment, and fuels. 
It should be noted that, as illustrated in Table V.D-2, above, 
aggregate program costs peak in 2014; increases in costs after that 
year are due to increases in the population of engines over time. The 
other years, 2020 and 2030, correspond to years analyzed in our 
benefits analysis. Detailed results for all years are included in 
Appendix 10.E. for this chapter.
a. Expected Market Impacts
    The market impacts of this rule suggest that the overall economic 
impact of the proposed emission control program on society is expected 
to be small, on average. According to this analysis, the average prices 
of goods and services produced using equipment and fuel affected by the 
proposal are expected to increase by about 0.02 percent. The estimated 
price increases and quantity reductions for engines and equipment vary 
depending on compliance costs. In general, we would expect for price 
increases to be higher (lower) as a result of a high (low) relative 
level of compliance costs to market price. We would also expect the 
change in price to be highest when compliance costs are highest.
    The estimated market impacts for 2013, 2020, and 2030 are presented 
in Table V.F-1. The market-level impacts presented in this table 
represent production-weighted averages of the individual market-level 
impact estimates generated by the model: the average expected price 
increase and quantity decrease across all of the units in each of the 
engine, equipment, fuel, and final application markets. For example, 
the model includes seven individual engine markets that reflect the 
different horsepower size categories. The 23 percent price change for 
engines shown in Table V.F-1 for 2013 is an average price change across 
all engine markets weighted by the number of production units. 
Similarly, equipment impacts presented in Table V.F-1 are weighted 
averages of 42 equipment-application markets, such as small (< 25hp) 
agricultural equipment and large (£600hp) industrial 
equipment. It should be noted that price increases and quantity 
decreases for specific types of engines, equipment, application 
sectors, or diesel fuel markets are likely to be different. But the 
data in this table provide a broad overview of the expected market 
impacts that is useful when considering the impacts of the proposal on 
the economy as a whole. The individual market-level impacts are 
presented in Chapter 10 of the draft RIA for this proposal.
    Engine Market Results: Most of the variable costs associated with 
the proposed rule are passed along in the form of higher prices. The 
average price increase in 2013 for engines is estimated to be about 23 
percent. This percentage is expected to decrease to about 19.5 percent 
for 2020 and later. This expected price increase varies by engine size 
because compliance costs are a larger share of total production costs 
for smaller engines. In 2013, the year of greatest compliance costs 
overall, the largest expected percent price increase is for engines 
between 25 and 50 hp: 34 percent or $852; the average price for an 
engine in this category is about $2,500. However, this price increase 
is expected to drop to 26 percent, or about $647, for 2016 and later. 
The smallest expected percent price increase in 2013 is for engines in 
the greater than 600 hp category. These engines are expected to see 
price increases of about 3 percent increase in 2013, increasing to 
about 5.6 percent in 2014 and beyond. The expected price increase for 
these engines is about $4,211 in 2013, increasing to about $6,950 in 
2014 and later, for engines that cost on average about $125,000.
    The market impact model predicts that even with these increases in 
engine prices, total demand is not expected to change very much. The 
expected average change in quantity is only about 69 engines per year 
in 2013, out of total sales of more than 500,000 engines. The estimated 
change in market quantity is small because as compliance costs are 
passed along the supply chain they become a smaller share of total 
production costs. In other words, firms that use these engines and 
equipment will continue to purchase them even at the higher cost 
because the increase in costs will not have a large impact on their 
total production costs. Diesel equipment is only one factor of 
production for their output of construction, agricultural, or 
manufactured goods. The average decrease in the quantity of all engines 
produced as a result of the regulation is estimated to be about 0.013 
percent. This decrease ranges from 0.010 percent for engines less than 
25 hp to 0.016 percent for engines 175 to 600 hp.
    Equipment Market Results: Estimated price changes for the equipment 
markets reflect both the direct costs of the proposed standards on 
equipment production and the indirect cost through increased engine 
prices. In 2013, the average price increase for nonroad diesel 
equipment is estimated

[[Page 28458]]

to be about 5.2 percent. This percentage is expected to decrease to 
about 4.5 percent for 2020 and beyond. The range of estimated price 
increases across equipment types parallels the share of engine costs 
relative to total equipment price, so the estimated percentage price 
increase among equipment types also varies. The market price in 2013 
for agricultural equipment between 175 and 600 hp is estimated to 
increase about 1.4 percent, or $1,835 for equipment with an average 
cost of $130,000. This compares with an estimated engine price increase 
of about $1,754 for engines of that size. The largest expected price 
increase in 2013 for equipment is $4,335, or 4.9 percent, for pumps and 
compressors over 600 hp. This compares with an estimated engine price 
increase of about $4,211 for engines of that size. The smallest 
expected price increase in 2013 for equipment is $125, or 3.6 percent, 
for construction equipment less than 25 hp. This compares with an 
estimated engine price increase of about $124 for engines of that size. 
The price changes for the equipment are less than that for engines 
because the engine is only one input in the production of equipment.
    The output reduction for nonroad diesel equipment is estimated to 
be very small and to average about 0.014 percent for all years. This 
decrease ranges from 0.005 percent for general manufacturing equipment 
to 0.019 percent for construction equipment. The largest expected 
decrease in quantity in 2013 is 13 units of construction equipment per 
year for construction equipment between 100 and 175 hp, out of about 
62,800 units. The smallest expected decrease in quantity in 2013 is 
less than one unit per year in all hp categories of pumps and 
compressors.

                                 Table V.F-1.--Summary of Market Impacts ($2001)
----------------------------------------------------------------------------------------------------------------
                                    Engineering           Change in price               Change in quantity
                                       cost      ---------------------------------------------------------------
             Market              ----------------    Absolute
                                     Per unit       ($million)        Percent        Absolute         Percent
----------------------------------------------------------------------------------------------------------------
                                                      2013
----------------------------------------------------------------------------------------------------------------
Engines.........................          $1,087            $840            22.9           -69 a          -0.013
Equipment.......................           1,021           1,017             5.2            -118          -0.014
Application Markets b...........  ..............  ..............            0.02  ..............          -0.010
No. 2 Distillate Nonroad........           0.039           0.038             4.1         -1.38 c          -0.013
---------------------------------
                                                      2020
----------------------------------------------------------------------------------------------------------------
Engines.........................          $1,028            $779            19.5           -79 a          -0.013
Equipment.......................           1,018           1,013             4.4            -135          -0.014
Application Markets b...........  ..............  ..............            0.02  ..............          -0.010
No. 2 Distillate Nonroad........           0.039           0.039             4.1         -1.58 c          -0.014
---------------------------------
                                                      2030
----------------------------------------------------------------------------------------------------------------
Engines.........................          $1,027            $768            19.4           -92 a          -0.013
Equipment.......................           1,004             999             4.5            -156          -0.014
Application Markets b...........  ..............  ..............            0.02  ..............          -0.010
No. 2 Distillate Nonroad........           0.039           0.039             4.1         -1.84 c         -0.014
----------------------------------------------------------------------------------------------------------------
Notes:
a The absolute change in the quantity of engines represents only engines sold on the market. Reductions in
  engines consumed internally by integrated engine/equipment manufacturers are not reflected in this number but
  are captured in the cost analysis. For this reason, the absolute change in the number of engines and equipment
  does not match.
b The model uses normalized commodities in the application markets because of the great heterogeneity of
  products. Thus, only percentage changes are presented.
c Units are in million of gallons.

    Application Market Results: The estimated price increase associated 
with the proposed standards in all three of the application markets is 
very small and averages about 0.02 percent for all years. In other 
words, on average, the prices of goods and services produced using the 
engines, equipment, and fuel affected by this proposal are expected to 
increase only negligibly. This is because in all of the application 
markets the compliance costs passed on through price increases 
represent a very small share of total production costs. For example, 
the construction industry realizes an increase in production costs of 
approximately $468 million in 2013 because of the price increases for 
diesel equipment and fuel. However, this represents only 0.03 percent 
of the $1,392 billion value of shipments in the construction industry 
in 2001. The estimated average commodity price increase in 2013 ranges 
from 0.06 percent in the agricultural application market to about 0.01 
percent in the manufacturing application market. The percentage change 
in output is also estimated to be very small and averages about 0.01 
percent. This reduction ranges from less than a 0.01 percent decrease 
in manufacturing to about a 0.02 percent decrease in construction. Note 
that these estimated price increases and quantity decreases are average 
for these sectors and may vary for specific subsectors. Also, note that 
absolute changes in price and quantity are not provided for the 
application markets in Table V.F-1 because normalized commodity values 
are used in the market model. Because of the great heterogeneity of 
manufactured or agriculture products, a normalized commodity ($1 unit) 
is used in the application markets. This has no impact on the estimated 
percentage change impacts but makes interpretation of the absolute 
changes less informative.
    Fuel Markets Results: The estimated average price increase across 
all nonroad diesel fuel is about 4 percent for all years. For 15 ppm 
fuel, the estimated price increase for 2013 ranges from 3.2 percent in 
the East Coast region (PADD 1&3) to 9.3 percent in the mountain region 
(PADD 4). The average

[[Page 28459]]

national output decrease for all fuel is estimated to be about 0.01 
percent for all years, and is relatively constant across all four 
regional fuel markets.
b. Expected Welfare Impacts
    Social cost impact estimates are presented in Table V.F-2. A time 
series of social costs from 2007 through 2030 is presented in Table 
IV.F-3. As described above, the total social cost of the regulation is 
the sum of the changes in producer and consumer surplus estimated by 
the model plus engine maintenance savings (negative costs) resulting 
from using fuel with a lower sulfur content. Total social costs in 2013 
are projected to be 1,202.4 million ($2001). About 82 percent of the 
total social costs is expected to be borne by producers and consumers 
in the application markets, indicating that the majority of the costs 
are expected to be passed on in the form of higher prices. When these 
estimated impacts are broken down, 58 percent are expected to be borne 
by consumers in the application markets and 42 percent are expected to 
be borne by producers in the application markets. Equipment 
manufacturers are expected to bear about 10 percent of the total social 
costs. Engine manufacturers and diesel fuel refineries are expected to 
bear 2.5 percent and 0.5 percent, respectively. The remaining 5.0 
percent is accounted for by fuel marker costs and the additional costs 
of 15 ppm fuel being sold in to markets such as marine diesel, 
locomotive, and home heating fuel that do not require it.
    In 2030, the total social costs are projected to be about $1,509.6 
million ($2001). The increase is due to the projected annual growth in 
the engine and equipment populations. As in earlier years, producers 
and consumers in the application markets are expected to bear the large 
majority of the costs, approximately 94 percent. This is consistent 
with economic theory, which states that, in the long run, all costs are 
passed on to the consumers of goods and services.
    The present value of total social costs through 2030 is estimated 
to be $16.5 billion ($2001). This present value is calculated using a 
social discount rate of 3 percent from 2004 through 2030. We also 
performed an analysis using an alternative 7 percent social discount 
rate. Using that discount rate, the present value of the social costs 
through 2030 is estimated to be $9.9 billion ($2001).

                          Table V.F-2.--Summary of Social Costs Estimates Associated With Primary Program: 2013, 2020, and 2030
                                                                      [$million]a,b
--------------------------------------------------------------------------------------------------------------------------------------------------------
                                                   Maximum cost year (2013)                    Year 2020                       Final year (2030)
                                             -----------------------------------------------------------------------------------------------------------
                                                Market     Operating                Market     Operating                Market     Operating
                                               surplus      savings      Total     surplus      savings      Total     surplus      savings      Total
                                               ($10\6\)    ($10\6\)                ($10\6\)    ($10\6\)                ($10\6\)    ($10\6\)
--------------------------------------------------------------------------------------------------------------------------------------------------------
Engine Producers Total......................       30.2  ............       30.2        0.1  ............        0.1        0.1  ............        0.1
Equipment Producers Total...................      116.1  ............      116.1      102.6  ............      102.6        5.3  ............        5.3
    Agricultural Equipment..................       39.9  ............       39.9       33.2  ............       33.2        1.3  ............        1.3
    Construction Equipment..................       53.0  ............       53.0       48.2  ............       48.2        3.8  ............        3.8
    Industrial Equipment....................       23.2  ............       23.2       21.2  ............       21.2        0.2  ............        0.2
Application Producers and Consumers Total...    1,231.8       (241.9)      989.8    1,386.5       (190.1)    1,196.3    1,598.9       (174.5)    1,424.5
    Total Producer..........................      515.7  ............  .........      583.4  ............  .........      672.9  ............  .........
    Total Consumer..........................      716.1  ............  .........      803.1  ............  .........      926.0  ............  .........
    Agriculture.............................      348.7        (44.7)      304.0      339.2        (35.2)      364.0      416.5        (32.3)      429.2
    Construction............................      468.3        (77.9)      390.4      550.4        (61.2)      489.3      635.7        (56.1)      579.5
    Manufacturing...........................      414.8       (119.3)      295.5      436.8        (93.8)      343.0      501.8        (86.0)      415.7
Fuel Producers Total........................        7.8  ............        7.8        9.0  ............        9.0       10.5  ............       10.5
    PADD I&III..............................        3.6  ............        3.6        4.1  ............        4.1        4.8  ............        4.8
    PADD II.................................        2.9  ............        2.9        3.3  ............        3.3        3.9  ............        3.9
    PADD IV.................................        0.8  ............        0.8        0.9  ............        0.9        1.0  ............        1.0
    PADD V..................................        0.5  ............        0.5        0.6  ............        0.6        0.8  ............        0.8
Nonroad Spillover...........................  .........         51.2   .........  .........         58.6   .........  .........         69.2
Marker Costs................................  .........          7.3   .........  .........  ............  .........  .........  ............  .........
                                             ------------
--------------------------------------------------------------------------------------------------------------------------------------------------------
Notes:
\a\ Figures are in 2001 dollars.
\b\ Operating savings are shown as negative costs.

[[Page 28460]]

  Table IV.F-3--National Engineering Compliance Costs and Social Costs
               Estimates for the Proposed Rule: 2004-2030
                               [$10 \6\]
a
------------------------------------------------------------------------
                                               Engineering      Total
                    Year                       compliance       social
                                                  costs         costsb
------------------------------------------------------------------------
2004.......................................            0.00         0.00
2005.......................................            0.00         0.00
2006.......................................            0.00         0.00
2007.......................................           39.61        39.61
2008.......................................          130.41       130.40
2009.......................................          132.25       132.25
2010.......................................          262.02       262.01
2011.......................................          641.12       641.07
2012.......................................        1,010.37     1,010.27
2013.......................................        1,202.52     1,202.40
2014.......................................        1,329.14     1,329.01
2015.......................................        1,260.74     1,260.62
2016.......................................        1,298.40     1,298.27
2017.......................................        1,318.75     1,318.62
2018.......................................        1,325.02     1,324.89
2019.......................................        1,339.30     1,339.16
2020.......................................        1,366.79     1,366.66
2021.......................................        1,351.08     1,350.94
2022.......................................        1,349.58     1,349.44
2023.......................................        1,365.53     1,365.38
2024.......................................        1,371.60     1,371.45
2025.......................................        1,395.98     1,395.83
2026.......................................        1,419.79     1,419.64
2027.......................................        1,442.91     1,442.76
2028.......................................        1,465.41     1,465.26
2029.......................................        1,487.68     1,487.53
2030.......................................        1,509.77     1,509.61
--------------------------------------------
NPV at 3%..................................       16,524.29    16,522.66
NPV at 7%..................................        9,894.02    9,893.06
------------------------------------------------------------------------
Notes:
a Figures are in 2001 dollars.
b Figures in this column do not include the human health and
  environmental benefits of the proposal.

VI. Alternative Program Options

    Our proposed emission control program consists of a two-step 
program to reduce the sulfur content of nonroad diesel fuel in 
conjunction with the proposed Tier 4 engine standards. As we developed 
this proposal, we evaluated a number of alternative options with regard 
to the scope, level, and timing of the standards. This section presents 
a summary of our analysis of several alternative control scenarios. A 
complete discussion of all the alternatives, their feasibility, and 
their inventory, benefits, and cost impacts can be found in Chapter 12 
of the draft Regulatory Impact Analysis for this proposal.
    While we are interested in comments on all of the alternatives 
presented, we are especially interested in comments on two alternative 
scenarios which EPA believes merit further consideration in developing 
the final rule: a program in which sulfur levels are required to be 
reduced to 15 ppm in essentially a single step, and a variation on the 
proposed two-step fuel control program, in which the second step of 
sulfur control to 15 ppm in 2010 would apply to locomotive and marine 
diesel fuel in addition to nonroad diesel fuel. This section describes 
these two options in greater detail; additional information can be 
found in Chapter 12 of the draft Regulatory Impact Analysis for this 
proposal.

A. Summary of Alternatives

    We developed emissions, benefits, and cost analyses for a number of 
alternatives. The alternatives we considered can be categorized 
according to the structure of their fuel requirements: whether the 15 
ppm fuel sulfur limit is reached in two-steps, like the proposed 
program, or one-step.
    One-step alternatives are those in which the fuel sulfur standard 
is applied in a single step: there are no fuel-based phase-ins. We 
evaluated three one-step alternatives. Option 1 is described in detail 
in Section VI.B, below. We considered two other one-step alternatives 
which differ from Option 1 in the timing of the fuel option (2006 or 
2008) and the engines standards (level of the standards and when they 
are introduced). As described in Table IV-1, Option 1b differs from 
Option 1 regarding the timing of the fuel standards, while Option 1a 
differs from Option 1 in terms of the engine standards. Both Option 1a 
and 1b would also extend the 15 ppm fuel sulfur limit to locomotive and 
marine diesel fuel as well.
    Two-step alternatives are those in which the fuel sulfur standard 
is set first at 500 ppm and then is reduced to 15 ppm. The two-step 
alternatives vary from the proposal in terms of both the timing and 
levels of the engine standards and the timing of the fuel standards. 
Option 2a is the same as the proposed program except the 500 ppm fuel 
standard is introduced a year earlier, in 2006. Option 2b is the same 
as the proposed program except the 15 ppm fuel standard is introduced a 
year earlier in 2009 and the trap-based PM standards begin earlier for 
all engines. Option 2c is the same as the proposed program except the 
15 ppm fuel standard is introduced a year earlier in 2009 and the trap-
based PM standards begin earlier for engines 175-750 hp. Option 2d is 
the same as the proposed program except the NOX standard is 
reduced to 0.30 g/bhp-hr for engines 25-75 hp, and this standard is 
phased in. Finally, Option 2e is the same as the proposed program 
except there are no new Tier 4 NOX limits.
    Options 3 and 4 are identical to the proposed program, except 
Option 3 would exempt mining equipment over 750 hp from the Tier 4 
standards, and Option 4 would include applying the 15 ppm sulfur limit 
to both locomotive and marine diesel fuel. Option 4 is discussed in 
detail in Section IV.C, below.
    Option 5a and 5b are identical to the proposal except for the 
treatment of engines less than 75 hp. Option 5a is identical to the 
proposal except that no new program requirements would be set in Tier 4 
for engines under 75 hp. Instead Tier 2 standards and testing 
requirements for engines under 50 hp, and Tier 3 standards and testing 
requirements for 50-75 hp engines, would continue indefinitely. The 
Option 5b program is identical to the proposal except that for engines 
under 75 hp only the 2008 engine standards would be set. There would be 
no additional PM filter-based standard in 2013 for 25-75 hp engines, 
and no additional NOX+NMHC standard in 2013 for 25-50 hp 
engines.
    Table VI-1 contains a summary of a number of these alternatives and 
the expected emission reductions, costs, and monetized benefits 
associated with them in comparison to the proposal. These alternatives 
cover a broad range of possible approaches and serve to provide insight 
into the many other program design alternatives not expressly evaluated 
further. The analysis was done using a 3% discount rate. If we were to 
use another rate, the values would change but not to such a degree as 
to change our conclusions regarding the various options. A complete 
discussion of all the alternatives, their feasibility, and their 
inventory, benefits, and cost impacts can be found in Chapter 12 of the 
draft Regulatory Impact Analysis for this proposal.

[[Page 28461]]
[GRAPHIC]
[TIFF OMITTED]
TP23MY03.012
[[Page 28462]]
[GRAPHIC]
[TIFF OMITTED]
TP23MY03.013
[[Page 28463]]

B. Introduction of 15 ppm Nonroad Diesel Sulfur Fuel in One Step

    EPA carefully evaluated and is seeking comment on alternative 
regulatory approaches. Instead of the proposed two-step reduction in 
nonroad diesel sulfur, one alternative would require that the nonroad 
diesel sulfur level be reduced to 15ppm beginning June 1, 2008. This 
alternative would have the advantage of enabling use of high efficiency 
exhaust emission control technology for nonroad engines as early as the 
2009 model year. It also would have several disadvantages which have 
prompted us not to propose it. The disadvantages in comparison to the 
proposal include inadequate lead-time for engine and equipment 
manufacturers and refiners, leading to increased costs and potential 
market disruptions. In this section, we describe this alternative in 
greater detail and discuss potential engine and fuel impacts. We also 
present our estimated emission and benefit impacts. Two other one-step 
fuel options which are variations of the alternative discussed in this 
section, Options 1a and 1b in Table VI-1, are presented in Chapter 12 
of the draft RIA for this proposal.
1. Description of the One-Step Alternative
    While numerous engine standards and phase-in schedules are 
possible, we considered the standards shown in Tables VI-2 and VI-3 as 
being the most stringent one-step program that could be considered 
potentially feasible considering cost, lead-time, and other factors. 
These standards are similar to those in our proposed option, the 
primary difference being the generally earlier phase-in dates for the 
PM standards.

                               Table VI-2.--PM Standards for 1-Step Fuel Scenario
                                                   [g/bhp-hr]
----------------------------------------------------------------------------------------------------------------
                                                                           Model year
                 Engine power                  -----------------------------------------------------------------
                                                   2009       2010       2011       2012       2013       2014
----------------------------------------------------------------------------------------------------------------
hp < 25.......................................       0.30  .........  .........  .........  .........  .........
25 <= hp <50..................................      10.22  .........  .........  .........       0.02  .........
50 <= hp <75..................................  .........  .........  .........  .........       0.02  .........
75 <= hp <175.................................  .........  .........       0.01  .........  .........  .........
                                                .........    \a\ 50%    \a\ 50%   \a\ 100%  .........  .........
175 <= hp <750................................  .........       0.01  .........  .........  .........  .........
                                                  \a\ 50%    \a\ 50%   \a\ 100%  .........  .........  .........
hp £= 750...........................  .........  .........  .........  .........       0.01  .........
                                                .........  .........    \a\ 50%    \a\ 50%    \a\ 50%  \a\ 100%
----------------------------------------------------------------------------------------------------------------
Notes:
\a\ Percentages are the model year sales required to comply with the indicated standard.

                          Table VI-3.--NOX and NMHC Standards for 1-Step Fuel Scenario
                                                   [g/bhp-hr]
----------------------------------------------------------------------------------------------------------------
                                                                                     Model year
                           Engine power                            ---------------------------------------------
                                                                       2011       2012       2013        2014
----------------------------------------------------------------------------------------------------------------
25 <= hp < 75.....................................................  .........  .........      a 3.5  ...........
                                                                              ------------
                                                                                            0.30 NOX
75 <= hp <175.....................................................                         0.14 NMHC
                                                                              ------------
                                                                                   b 50%      b 50%       b 100%
                                                                   ------------
                                                                                      0.30 NOX
175 <= hp <750....................................................                    0.14 NMHC
                                                                   ------------
                                                                        b 50%      b 50%      b 50%       b 100%
                                                                   ------------
                                                                                       0.30 NOX
hp £=750................................................                    0.14 NMHC
                                                                   ------------
                                                                        b 50%      b 50%      b 50%      b 100%
----------------------------------------------------------------------------------------------------------------
Notes:
a A 3.5 NMHC + NOX standard would apply to the 25-50 hp engines. Engines greater than 50hp are already subject
  to this standard in 2008 under the existing Tier 3 program.
b Percentages are the model year sales required to comply with the indicated standards.

2. Engine Emission Impacts
    The main advantage associated with this one-step approach is 
pulling ahead the long-term PM engine standards. By making 15 ppm 
sulfur fuel widely available by late 2008, we could accelerate the 
long-term PM engine standards, leading to the introduction of precious 
metal catalyzed PM traps as early as 2009, two years earlier than 
possible under the two-step sulfur reduction approach. Some 
stakeholders have expressed the concern that a two-step approach leads 
to later than desired introduction of high-efficiency exhaust emissions 
controls on nonroad diesels because this cannot happen until the 15 ppm 
fuel standard goes into effect. As shown in Table VI-1, there would be 
additional public health benefits associated with this one-step 
approach. However, in comparison to the proposal, the additional 
benefits are

[[Page 28464]]

relatively small, less than one percent or about $3 billion more than 
the proposed program.\295\
---------------------------------------------------------------------------

    \295\ A variation on this one-step approach would be to also 
require the sulfur content of locomotive and marine fuel to meet the 
15 ppm standard in 2008. The decision of whether or not to require 
the sulfur content of locomotive and marine fuel to also be reduced 
to 15 ppm, however, is not unique to the one step approach, and, as 
discussed below is an alternative also being evaluated under our 
proposed 2-step program. Were we to require locomotive and marine 
diesel fuel to also meet the 15 ppm standard in 2008 under a one-
step approach, there would be additional inventory reductions of 
about 10,000 tons of PM and 128,000 tons of SO3 (NPV 3% 
through 2030).
---------------------------------------------------------------------------

    Even though 15 ppm fuel would be available beginning June 1, 2008 
under this one-step approach, we do not believe it would be feasible to 
propose an aggressive turnover of new engines to trap-equipped versions 
in 2009. Nor would it be possible to introduce NOX controls 
any earlier than we are already proposing, model year 2011. The 
proposed standards need to be coordinated with Tier 3 standards, and 
with the heavy duty highway diesel standards. The coordination of Tier 
4 standards with Tier 3 standards and with the development of emissions 
control technology for highway diesel engines is of critical importance 
to successful implementation of the Tier 4 standards. Even those 
manufacturers who do not make highway engines are expected to gain 
substantially from the highway PM and NOX control 
development work, provided they can plan for standards set at a similar 
level of stringency and timed in a way to allow for the orderly 
migration of highway engine technology to nonroad applications.
    Thus, although the application of high-efficiency exhaust PM 
emission controls to nonroad diesels would be enabled with the 
introduction of 15 ppm sulfur nonroad fuel in 2008 under a one-step 
program, we believe that to require the application of PM controls 
across the wide spectrum of nonroad engines shortly thereafter would 
raise serious feasibility concerns that could only be resolved, if at 
all, through a very large additional R&D effort undertaken roughly in 
parallel with the similarly large highway R&D effort, a duplication of 
effort we wish to avoid for reasons discussed in Section III. Nonroad 
engine designers would need to accomplish much of this development well 
before the diesel experience begins to accumulate in earnest in 2007, 
in order to be ready for a 2009 first introduction date. Waiting until 
2007 before initiating 2009 model year design work would risk the 
possibility of product failures, limited product availability and major 
market disruptions. At the same time, for those engine manufacturers 
who participate in both the highway and nonroad diesel engine markets, 
attempting to have concurrent engine product developments for highway 
and nonroad, could result in the possibility of product failures, 
limited product availability and major disruptions for the highway 
market as well. Thus, in balancing their costs and burden, many 
manufacturers may be forced to choose which products would be available 
for 2009 and which products would be delayed for release. Manufacturers 
would also incur large additional costs to redesign hundreds of engine 
models and thousand of machine types to meet Tier 4 standards only one 
to three years after Tier 3 standards take effect in 2006-2008. These 
cost impacts are reflected in Table VI-1 and their derivation is 
explained in chapter 12 of the draft RIA. This extra expenditure could 
only be modestly mitigated by phasing in the standards, since a crash 
R&D effort with limited benefit from highway experience would still be 
necessary.
    Moreover, with respect to NOX, it would be impractical 
or simply infeasible to pull the standards ahead on the same schedule. 
This is because EPA's highway diesel program allows manufacturers to 
phase in NOX technology over 2007-2010. As a result, we do 
not expect that the high-efficiency NOX control technology 
could reasonably be applied to nonroad engines any earlier under a one-
step program than under a two-step program (i.e., beginning in 2011).
    In summary, this option would lead us to apply PM and 
NOX standards in two different model years, or else forgo 
any opportunity to apply PM traps in 2009. Redesigning engines and 
emission controls for early PM control and then again a couple of years 
later for NOX control, on top of shortened Tier 3 stability 
periods, would likely add substantial costs to the program. As 
manufacturers attempt to avoid these costs and optimize their 
development they may simply have to restrict product offerings for some 
period, leading to price spikes and shortages due to lack of product 
availability. Having the NOX and PM standards phase in 
simultaneously under our proposed approach avoids cost and design 
stability issues for both engine and equipment manufacturers. In 
addition, the longer leadtime for the engine standards under our 
proposed program will allow greater economic efficiencies for engine 
manufacturers as they transfer highway emission reduction technology to 
nonroad engines.
3. Fuel Impacts
    In addition to the challenges associated with pulling ahead the PM 
standards described above, there are also some concerns regarding the 
practicality of an early 15 ppm nonroad diesel sulfur standard. A one-
step approach may result in several economic inefficiencies that would 
increase the cost of the program. For example, refiners will have 
little opportunity to take advantage of the newer desulfurization 
technologies currently being developed. As described in sections IV and 
V, refiners will only begin to be able to take advantage of these new 
technologies in 2008. By 2010, the ability to incorporate them into 
their refinery modifications is expected to double. If refiners have to 
take steps to reduce the sulfur content of nonroad diesel fuel earlier, 
they will likely have to use more expensive current technology. The 
cost impacts of this decision will persist, since the choice of 
technology is a long term decision. If a refiner is forced by the 
effective date of the standards to employ a more expensive technology, 
that choice will affect that refiner's output indefinitely, since the 
cost of upgrading to the new technologies will be prohibitive. As 
presented in section 5.2 of the Draft RIA, we estimate that the costs 
of achieving a 15 ppm standard in 2008 is approximately 0.4 c/gal 
greater than for the proposal. While difficult to quantify there are 
also considerable advantages to allowing refiners some operating time 
in producing 15 ppm diesel fuel for the highway program prior to 
requiring them to solidify their designs for producing nonroad diesel 
fuel to 15 ppm. The primary advantage is that the design of 
desulfurization equipment used to produce 15 ppm nonroad diesel fuel 
can reflect the operating experience of the equipment used to produce 
15 ppm highway diesel fuel starting in 2006. This extra time would also 
provide current refiners of high sulfur diesel fuel with highly 
confident estimates of the cost of producing 15 ppm diesel fuel, 
reducing uncertainty and increasing their likelihood of investing to 
produce this fuel. With a start date of June 1, 2008 refiners would 
have to solidify their designs and start construction prior to getting 
any data on the performance of their highway technology. This would 
increase the cost of producing 15 ppm nonroad diesel fuel for the life 
of the new desulfurization equipment, as well as potentially delaying 
some refiners' decision to invest in new desulfurization equipment due 
to uncertainties in cost, performance, etc.

[[Page 28465]]

4. Emission and Benefit Impacts
    We used the nonroad model to estimate the emission inventory 
impacts associated with this one-step option, as well as the other 
options listed in Table VI-1. As for all the alternatives, we then used 
the benefits transfer method to estimate the monetized benefits of the 
alternative.\296\ The results are shown in Table VI-1. As is evidenced 
by the values in Table VI-1, the one-step alternative would achieve 
slightly greater PM and NOX emission reductions through 2030 
than the proposed 2-step program, with 6,000 and 11,000 additional tons 
reduced, respectively (or less than 0.5 percent). Unlike the proposed 
2-step program, however, there would be no SO2 emission 
reductions in 2007 due to the delay in fuel sulfur control, although 
2009 and later emission are slightly greater due primarily to the 
earlier introduction of engines using PM filters. Nevertheless, the 
SO2 benefits of the one-step program are slightly less than 
the proposed 2-step program in the long run, by about 191,000 tons 
(about 4 percent) through 2030.
---------------------------------------------------------------------------

    \296\ The results that were obtained for Option 1a were 
extrapolated based on the emission inventory changes to the proposed 
program and were obtained for the other alternatives by assuming the 
air quality changes between the alternative and the actual case run 
were small enough to allow for such extrapolation. An explanation of 
the benefits transfer method is contained in Chapter 9 of the draft 
RIA.
---------------------------------------------------------------------------

    After careful consideration of these matters, we have decided to 
propose the two-step approach in today's notice. The two-step program 
avoids adverse risks to the smooth implementation of the entire Tier 4 
nonroad program that could be caused by the significantly shortened 
lead-time and stability of the one-step program. There are also 
concerns about the potential negative impacts the one-step option may 
have on the 2007 highway program, including the implications of the 
overlap of implementation schedules (see above and Chapter 12 of the 
draft RIA). Nevertheless, we believe that the one-step approach is a 
regulatory alternative worth considering. In addition to seeking 
comment on our proposed program, we also seek comment on the relative 
merits and shortcomings of a one-step approach to regulating nonroad 
diesel fuel and the associated schedule for implementing the engine 
standards.

C. Applying 15 ppm Requirement to Locomotive and Marine Diesel Fuel

    To enable the high efficiency exhaust emission control technology 
to begin to be applied to nonroad diesel engines beginning with the 
2011 model year, we are proposing that all nonroad diesel fuel produced 
or imported after June 1, 2010 would have to meet a 15 ppm sulfur cap. 
Although locomotive and marine diesel engines are similar in size to 
some of the diesel engines covered in this proposal, there are many 
differences that have caused us to treat them separately in past EPA 
programs.\297\ These include differences in duty cycles and exhaust 
system design configurations, size, and rebuild and maintenance 
practices. Because of these differences, we are not proposing new 
engine standards today for these engine categories. Since we are not 
proposing more stringent emission standards, we are also not proposing 
that the second step of sulfur control to 15 ppm in 2010 be applied to 
locomotive and marine diesel fuel. Instead, we are proposing to set a 
sulfur fuel content standard of 500 ppm for diesel fuel used in 
locomotive and marine applications. This fuel standard is expected to 
provide considerable sulfate PM and SO2 benefits even 
without establishing more stringent emission standards for these 
engines. We estimate that, cumulatively through 2030, reducing the 
sulfur content of locomotive and marine diesel fuel would eliminate 
about 102,000 tons of sulfate PM (net present value, based on a 3 
percent discount rate).
---------------------------------------------------------------------------

    \297\ Locomotives, in fact, are treated separately from other 
nonroad engines and vehicles in the Clean Air Act, which contains 
provisions regarding them in section 213(a)(5). Less than 50 hp 
marine engines were included in the 1998 final rule for nonroad 
diesel engines, albeit with some special provisions to deal with 
marine-specific engine characteristics and operating cycles.
---------------------------------------------------------------------------

    As discussed in section IV, we are seriously considering the option 
of extending the 15 ppm sulfur standard to locomotive and marine fuel 
as early as June 1, 2010, including them in the second step of the 
proposed two-step program. There are several advantages associated with 
this alternative. First, as reflected in Table VI-1, it would provide 
important additional sulfate PM and SO2 emission reductions 
and the estimated benefits from these reductions would outweigh the 
costs by a considerable margin. Second, in some ways it would simplify 
the fuel distribution system and the design of the fuel program 
proposed today since a marker would not be required for locomotive and 
marine diesel fuel. Furthermore, the prices for locomotive and marine 
diesel fuel may be virtually unaffected. Under the proposal, we expect 
that a certain amount of marine fuel will be 15 ppm sulfur fuel 
regardless of the standard due to limitations in the production and 
distribution of unique fuel grades. Where 500 ppm fuel is available, 
the possible suppliers of fuel will likely be more constrained, 
limiting competition and allowing prices to approach that of 15 ppm 
fuel. If we were to bring locomotive and marine fuel to 15 ppm, the 
pool of possible suppliers could expand beyond those today, since 
highway diesel fuel will also be at the same standard. Third, it would 
help reduce the potential opportunity for misfueling of 2007 and later 
model year highway vehicles and 2011 and later model year nonroad 
equipment with higher sulfur fuel. Finally, it would allow refiners to 
coordinate plans to reduce the sulfur content of all of their nonroad, 
locomotive, and marine diesel fuel at one time. While in many cases 
this may not be a significant advantage, it may be a more important 
consideration here since it is probably not a question of whether 
locomotive and marine fuel must meet a 15 ppm cap, but merely when. As 
discussed in section IV, it is the Agency's intention to propose action 
in the near future to set new emission standards for locomotive and 
marine engines that could require the use of high efficiency exhaust 
emission control technology, and thus, also require the use of 15 ppm 
sulfur diesel fuel.\298\ We anticipate that such engine standards would 
likely take effect in the 2011-13 timeframe, requiring 15 ppm 
locomotive and marine diesel fuel in the 2010-12 timeframe. We intend 
to publish an advance notice of proposed rulemaking for such standards 
by the Spring of 2004 and finalize those standards by 2007.
---------------------------------------------------------------------------

    \298\ EPA established the most recent new standards for 
locomotives and marine diesel engines (including those under 50 hp) 
in separate actions (63 FR 18977, April 16, 1998, and 67 FR 68241, 
November 8, 2002).
---------------------------------------------------------------------------

    However, discussions with refiners have suggested there are 
significant advantages to leaving locomotive and marine diesel fuel at 
500 ppm, at least in the near-term and until we set more stringent 
standards for those engines. The locomotive and marine diesel fuel 
markets could provide an important market for off-specification 
product, particularly during the transition to 15 ppm for highway and 
nonroad diesel fuel in 2010. Waiting just a year or two beyond 2010 
would address the critical near-term needs during the transition. In 
addition, waiting just another year or two beyond 2010 is also 
projected to allow virtually all refiners to take advantage of the new 
lower cost technology.
    After careful consideration of these matters, we have decided not 
to propose

[[Page 28466]]

to apply the second step of sulfur control of 15 ppm to locomotive and 
marine diesel fuel at this time. Nevertheless, for the reasons 
described above, we are carefully weighing whether it would be 
appropriate to do so. Therefore, we seek comment on this alternative 
and the various advantages, disadvantages, and implications of it.

D. Other Alternatives

    We have also analyzed a number of other alternatives, as summarized 
in Table VI-1. Some of these focus on control options more stringent 
than our proposal while others reflect modified engine requirements 
that result in less stringent control. EPA has evaluated these options 
in terms of the feasibility, emissions reductions, costs, and other 
relevant factors. EPA believes the proposed approach is the proper one 
with respect to these factors, and believes the options discussed above 
while having possible merit in some areas, raise what we believe are 
different and significant concerns with respect to these factors 
compared to the proposed approach. Hence we did not include these 
options. These concerns are discussed in more detail in Chapter 12. 
These concerns are discussed in more detail in Chapter 12 of the draft 
RIA. Hence, we did not include these options as part of our proposal 
for nonroad fuel and engine controls. We are interested in comment on 
these alternatives, especially information regarding their feasibility, 
costs, and other relevant concerns.

VII. Requirements for Engine and Equipment Manufacturers

    This section describes the regulatory changes proposed for the 
engine and equipment compliance program. First, the proposed 
regulations for Tier 4 engines have been written in plain language. 
They are structured to contain the provisions that are specific to 
nonroad CI engines in a new proposed part 1039, and to apply the 
general provisions of existing parts 1065 and 1068. The proposed plain 
language regulations, however, are not intended to significantly change 
the compliance program, except as specifically noted in today's notice 
(and we are not soliciting comment on any part of the rule that remains 
unchanged substantively). As proposed, these plain language regulations 
would only apply for Tier 4 engines. The changes from the existing 
nonroad program are described below along with other notable aspects of 
the compliance program.

A. Averaging, Banking, and Trading

1. Are We Proposing To Keep the ABT Program for Nonroad Diesel Engines?
    EPA has included averaging, banking, and trading (ABT) programs in 
most mobile source emission control programs adopted in recent years. 
Our existing regulations for nonroad diesel engines include an ABT 
program (Sec.  89.201 through Sec.  89.212). We are proposing to retain 
the basic structure of the existing nonroad diesel ABT program with 
today's notice, though we are proposing a number of changes to 
accommodate implementation of the proposed emission standards. Behind 
these changes is the recognition that the proposed standards represent 
a major technological challenge to the industry. The proposed ABT 
program is intended to enhance the ability of engine manufacturers to 
meet the stringent standards proposed today. The proposed program is 
also structured to limit production of very high-emitting engines and 
to avoid unnecessary delay of the transition to the new exhaust 
emission control technology.
    We view the proposed ABT program as an important element in setting 
emission standards that are appropriate under CAA section 213 with 
regard to technological feasibility, lead time, and cost. The ABT 
program helps to ensure that the stringent standards we are proposing 
are appropriate under section 213(a) given the wide breadth and variety 
of engines covered by the standards. For example, if there are engine 
families that will be particularly costly or have a particularly hard 
time coming into compliance with the standard, this flexibility allows 
the manufacturer to adjust the compliance schedule accordingly, without 
special delays or exceptions having to be written into the rule. 
Emission-credit programs also create an incentive (for example, to 
generate credits in early years to create compliance flexibility for 
later engines) for the early introduction of new technology, which 
allows certain engine families to act as trailblazers for new 
technology. This can help provide valuable information to manufacturers 
on the technology before they apply the technology throughout their 
product line. This early introduction of clean technology improves the 
feasibility of achieving the standards and can provide valuable 
information for use in other regulatory programs that may benefit from 
similar technologies. Early introduction of such engines also secures 
earlier emission benefits.
    In an effort to make information on the ABT program more available 
to the public, we intend to issue periodic reports summarizing use of 
the proposed ABT program by engine manufacturers. The information 
contained in the periodic reports would be based on the information 
submitted to us by engine manufacturers, and summarized in a way that 
protects the confidentiality of individual engine manufacturers. We 
believe this information will also be helpful to engine manufacturers 
by giving them a better indication of the availability of credits. 
Again, our periodic reports would not contain any confidential 
information submitted by individual engine manufacturers, such as sales 
figures. Also, the information would be presented in a format that 
would not allow such confidential information to be determined from the 
reports.
2. What Are the Provisions of the Proposed ABT Program?
    The following section describes the changes proposed to the 
existing ABT program. In addition to those areas specifically 
highlighted, we are soliciting comments on all aspects of the proposed 
ABT changes, including comments on the need for and benefit of these 
changes to manufacturers in meeting the proposed emission standards.
    The ABT program has three main components. Averaging means the 
exchange of emission credits between engine families within a given 
engine manufacturer's product line. (Engine manufacturers divide their 
product line into ``engine families'' that are comprised of engines 
expected to have similar emission characteristics throughout their 
useful life.) Averaging allows a manufacturer to certify one or more 
engine families at levels above the applicable emission standard, but 
below a set upper limit. However, the increased emissions must be 
offset by one or more engine families within that manufacturer's 
product line that are certified below the same emission standard, such 
that the average emissions from all the manufacturer's engine families, 
weighted by engine power, regulatory useful life, and production 
volume, are at or below the level of the emission standard. (The 
inclusion of engine power, useful life, and production volume in the 
averaging calculations is designed to reflect differences in the in-use 
emissions from the engines.) Averaging results are calculated for each 
specific model year. The mechanism by which this is accomplished is 
certification of the engine family to a ``family emission limit'' (FEL) 
set by the manufacturer, which may be above or below the standard. An 
FEL that is established

[[Page 28467]]

above the standard may not exceed an upper limit specified in the ABT 
regulations. Once an engine family is certified to an FEL, that FEL 
becomes the enforceable emissions limit for all the engines in that 
family for purposes of compliance testing. Averaging is allowed only 
between engine families in the same averaging set, as defined in the 
regulations.
    Banking means the retention of emission credits by the engine 
manufacturer for use in future model year averaging or trading. Trading 
means the exchange of emission credits between nonroad diesel engine 
manufacturers which can then be used for averaging purposes, banked for 
future use, or traded to another engine manufacturer.
    The existing ABT program for nonroad diesel engines covers 
NMHC+NOX emissions as well as PM emissions. With today's 
notice we are proposing to make the ABT program available for the 
proposed NOX standards and proposed PM standards. (For 
engines less than 75 horsepower where we are proposing combined 
NMHC+NOX standards, the ABT program would continue to be 
available for the proposed NMHC+NOX standards as well as the 
proposed PM standards.) ABT would not be available for the proposed 
NMHC standards for engines above 75 horsepower or for the proposed CO 
standards for any engines.
    As noted earlier, the existing ABT program for nonroad diesel 
engines includes FEL caps--limits on how high the emissions from 
credit-using engine families can be. No engine family may be certified 
above these FEL caps. These limits provide the manufacturers compliance 
flexibility while protecting against the introduction of unnecessarily 
high-emitting engines. When we propose new standards, we typically 
propose new FEL caps for the new standards. In the past, we have 
generally set the FEL caps at the emission levels allowed by the 
previous standard, unless there was some specific reason to do 
otherwise. We are proposing to do otherwise here because the proposed 
standard levels in today's notice are so much lower than the current 
standards levels, especially the Tier 4 standards for engines above 75 
horsepower. The transfer to new technology is feasible and appropriate. 
Thus, to ensure that the ABT provisions are not used to continue 
producing old-technology high-emitting engines under the new program, 
the proposed FEL caps would not, in general, be set at the previous 
standards. An exception is for the proposed NMHC+NOX 
standard for engines between 25 and 50 horsepower effective in model 
year 2013, where we are proposing to use the previously applicable 
NMHC+NOX standard for the FEL cap since the gap between the 
previous and proposed standards is approximately 40 percent (rather 
than 90 percent for engines above 75 horsepower).
    For engines above 75 horsepower certified during the phase-in 
period, there would be two separate sets of engines with different FEL 
caps. For engines certified to the existing (Tier 3) 
NMHC+NOX standards during the phase-in, the FEL cap would 
necessarily continue to be the existing FEL caps as adopted in the 
October 1998 rule. For engines certified to the proposed Tier 4 
NOX standard during the phase-in, the FEL cap would be 3.3 
g/bhp-hr for engines between 75 and 100 horsepower, 2.8 g/bhp-hr for 
engines between 100 and 750 horsepower, and 4.6 g/bhp-hr for engines 
above 750 horsepower. These proposed NOX FEL caps represent 
an estimate of the NOX emission level that is expected under 
the combined NMHC+NOX standards that apply with the existing 
previous tier standards. Beginning in model year 2014 when the proposed 
Tier 4 NOX standard for engines above 75 horsepower take 
full effect, we are proposing a NOX FEL cap of 0.60 g/bhp-hr 
for engines above 75 horsepower. (As described below, we are proposing 
to allow a small number of engines greater than 75 horsepower to have 
NOX FELs above the 0.60 g/bhp-hr cap beginning in model year 
2014.) Given the fact that the proposed Tier 4 NOX standard 
is approximately a 90 percent reduction from the existing standards for 
engines above 75 horsepower, we do not believe the previous standard 
would be appropriate as the FEL cap for all engines once the Tier 4 
standards are fully phased-in. We believe that the proposed 
NOX FEL caps will ensure that manufacturers adopt 
NOX aftertreatment technology across all of their engine 
designs (with the exception of a limited number) but will also allow 
for some meaningful use of averaging during the phase-in period. When 
compared to the proposed 0.30 g/bhp-hr NOX standard, the 
proposed NOX FEL cap of 0.60 g/bhp-hr (effective when the 
Tier 4 standards are fully phased-in) is consistent with FEL caps set 
in previous rulemakings.
    For the transitional PM standards being proposed for engines 
between 25 and 75 horsepower effective in model year 2008 and for the 
Tier 4 PM standards for engines below 25 horsepower, we are proposing 
the previously applicable Tier 2 PM standards (which do vary within the 
25 to 75 horsepower category) for the FEL caps since the gap between 
the previous and proposed standards is approximately 50 percent (rather 
than in excess of 90 percent for engines above 75 horsepower). For the 
proposed Tier 4 PM standard effective in model year 2013 for engines 
between 25 and 75 horsepower, we are proposing a PM FEL cap of 0.04 g/
bhp-hr, and for the proposed Tier 4 PM standard effective in model 
years 2011 and 2012 for engines between 75 and 750 horsepower, we are 
proposing a PM FEL cap of 0.03 g/bhp-hr. (As described below, we are 
proposing to allow a small number of Tier 4 engines greater than 25 
horsepower to have PM FELs above these caps.) Given the fact that the 
proposed Tier 4 PM standards for engines above 25 horsepower are less 
than 10 percent of the previous standards, we do not believe the 
previous standards would be appropriate as FEL caps once the Tier 4 
standards take effect. We believe that the proposed PM FEL caps will 
ensure that manufacturers adopt PM aftertreatment technology across all 
of their engine designs (except for a limited number of engines), yet 
will still provide substantial flexibility in meeting the standards.
    For the proposed Tier 4 PM standards for engines above 750 
horsepower there is a phase-in period during model years 2011 through 
2013. During the phase-in period, there would be two separate sets of 
engines with different FEL caps. For engines certified to the existing 
Tier 2 PM standard, the FEL cap would continue to be the existing PM 
FEL cap adopted in the October 1998 rule. For engines certified to the 
proposed Tier 4 PM standard during the phase-in, the FEL cap would be 
0.15 g/bhp-hr (the PM standard for the previous tier). Beginning in 
model year 2014, when the proposed Tier 4 PM standard for engines above 
750 horsepower takes full effect, consistent with the proposed caps for 
lower horsepower categories, we are proposing a PM FEL cap of 0.03 g/
bhp-hr. (As described below, we are proposing to allow a small number 
of engines greater than 750 horsepower to have PM FELs above the 0.03 
g/bhp-hr cap beginning in model year 2014.) We believe that the 
proposed PM FEL caps for engines above 750 horsepower will ensure that 
manufacturers adopt PM aftertreatment technology across all of their 
engine designs once the standard is fully phased-in (with the exception 
of a limited number) while allowing for some meaningful use of 
averaging during the phase-in period.
    Table VII.A-1 contains the proposed FEL caps and the effective 
model year

[[Page 28468]]

for the FEL caps (along with the associated standards proposed for Tier 
4). We request comment on the need for and the levels of these proposed 
FEL caps. It should be noted that for Tier 4, where we are proposing a 
new transient test, as well as retaining the current steady-state test, 
the FEL established by the engine manufacturer would be used as the 
enforceable limit for the purpose of compliance testing under both test 
cycles. In addition, under the NTE requirements, the FEL times the 
appropriate multiplier would be used as the enforceable limit for the 
purpose of such compliance testing.

                                 Table VII.A-1.--Proposed FEL Caps for the Proposed Tier 4 Standards in the ABT Program
                                                                       [g/bhp-hr]
--------------------------------------------------------------------------------------------------------------------------------------------------------
                                                                              NOX
            Power category                    Effective model year          standard                 NOX FEL cap                PM standard   PM FEL cap
--------------------------------------------------------------------------------------------------------------------------------------------------------
hp < 25 (kW < 19).....................  2008+...........................        (\a\)  (\a\)..................................     \b\ 0.30         0.60
25 <= hp < 50 (19 <= kW < 37).........  2008-2012.......................        (\a\)  (\a\)..................................         0.22         0.45
25 <= hp < 50 (19 <= kW < 37).........  2013+\d\........................      \e\ 3.5  5.6 \e\................................         0.02     \f\ 0.04
50 <= hp < 75 (37 <= kW < 56).........  2008-2012.......................        (\a\)  (\a\)..................................         0.22         0.30
50 <= hp < 75 (37 <= kW < 56).........  2013+...........................        (\a\)  (\a\)..................................         0.02     \f\ 0.04
75 <= hp <175 (56 <= kW <130).........  2012-2013 \g\...................         0.30  3.3 for hp < 100 2.8 for hp = 100.
75 <= hp <175 (56 <= kW <130).........  2014+...........................         0.30  0.60 \f\...............................         0.01     \f\ 0.03
175 <= hp <=750 (130 <= kW <=560).....  2011-2013.......................         0.30  2.8....................................         0.01     \f\ 0.03
175 <= hp <=750 (130 <= kW <=560).....  2014+...........................         0.30  0.60 \f\...............................         0.01     \f\ 0.03
hp £750 (kW £560).  2011-2013.......................         0.30  4.6....................................         0.01         0.15
hp £750 (kW £560).  2014+...........................         0.30  0.60 \f\...............................         0.01     \f\ 0.03
--------------------------------------------------------------------------------------------------------------------------------------------------------
Notes:
\a\ The existing NMHC+NOX standard and FEL cap apply (see CFR Title 40, section 89.112).
\b\ A PM standard of 0.45 g/bhp-hr would apply to air-cooled, hand-startable, direct injection engines under 11 horsepower, effective in 2010.
\c\ The proposed FEL caps do not apply if the manufacturer elects to comply with the optional standards. The existing FEL caps continue to apply.
\d\ FEL caps apply in model year 2012 if the manufacturer elects to comply with the optional standards.
\e\ These are a combined NMHC+NOX standard and FEL cap.
\f\ As described in this section, a small number of engines are allowed to exceed these FEL caps.
\g\ This period would extend through the first nine months of 2014 under the alternative, reduced phase-in requirement (see Section III.B.1. for a
  description of the proposed alternative).

    As noted above, we are proposing to allow a limited number of 
engines to have a higher FEL than the caps noted in Table VII.A-1 in 
certain instances. Under this proposal, the allowance to certify up to 
these higher FEL caps would apply to Tier 4 engines at or above 25 
horsepower. The provisions are intended to provide some limited 
flexibility for engine manufacturers as they transition to the 
stringent standards while ensuring that the vast majority of engines 
are converted to the advanced low-emission technologies expected under 
the Tier 4 program. This additional lead time appears appropriate, 
given the potential that a limited set of nonroad engines may face 
especially challenging difficulties in complying, and considering 
further that the same amount of overall emission reductions would be 
achieved through the need for credit-generating nonroad engines.
    Beginning the first year Tier 4 standards apply in each power 
category above 25 horsepower, an engine manufacturer would be allowed 
to certify up to ten percent of its engines in each power category with 
PM FELs above the caps shown in Table VII.A-1. The PM FEL cap for such 
engines would instead be the applicable previous tier PM standard. The 
ten percent allowance would be available for the first four years the 
Tier 4 standards apply. For the power categories in which we are 
proposing a phase-in requirement for the Tier 4 NOX 
standards, the allowance to use a higher FEL cap would apply only to PM 
during the phase-in years. Once the phase-in period is complete, the 
allowance would apply to NOX as well. (For engines above 750 
horsepower, where we are proposing a phase-in for both NOX 
and PM, the allowance to use a higher FEL cap would not take effect 
until model year 2014 when the phase-in was complete.)
    After the fourth year the Tier 4 standards apply, the allowance to 
certify engines using the higher FEL caps would still be available but 
for no more than five percent of a manufacturer's engines in each power 
category. (For the power category between 25 and 75 horsepower, this 
allowance would apply beginning with the 2013 model year and would 
apply to PM. The allowance to use the higher FEL caps is not necessary 
for the 2008 proposed standards or the 2013 proposed 
NMHC+NOX standards because the FEL caps for those standards 
are set at the previously applicable tier standards.)
    Table VII.A-2 presents the model years, percent of engines, and 
higher FEL caps that would apply under this allowance. Because the 
engines certified with the higher FEL caps are certified to the Tier 4 
standards (albeit through the use of credits), they would be considered 
Tier 4 engines and all other requirements for Tier 4 engines would also 
apply, including the Tier 4 NMHC standard. We invite comment on whether 
additional provisions may be necessary for the limited number of 
engines certified to the higher FELs, including whether an averaging 
program for NMHC would be needed.

[[Page 28469]]

                                 Table VII.A-2.--Allowance for Limited Use of an FEL Cap Higher Than the Tier 4 FEL Caps
--------------------------------------------------------------------------------------------------------------------------------------------------------
                                                                             Engines
                                                                           allowed to
            Power category                        Model years              have higher    NOX FEL cap (g/bhp-hr)           PM FEL cap (g/bhp-hr)
                                                                              FELs
--------------------------------------------------------------------------------------------------------------------------------------------------------
25 <= hp <75 (19 <= kW < 56).........  2013-2016.......................              10  Not applicable.........  0.22.
                                       2017+...........................               5                           ......................................
--------------------------------------
75 <= hp <175 (56 <= kW <130)........  2012-2013a......................              10  Not applicable.........  0.30 for hp <100.
                                      ---------------------------------------------------------------------------
                                       2014-2015.......................              10  3.3 for hp <100........  0.22 for hp £=100.
                                      --------------------------------------------------
                                       2016+...........................               5  2.8 for hp =100.
--------------------------------------
175 <= hp <=750 (130 <= kW <= 560)...  2011-2013.......................              10  Not applicable.........  0.15.
                                      --------------------------------------------------
                                       2015+...........................               5                           ......................................
--------------------------------------
hp £750 (kW £      2014-2017.......................              10  4.6....................  0.15.
 560).
                                      --------------------------------------------------
                                       2018+...........................               5                           ......................................
--------------------------------------------------------------------------------------------------------------------------------------------------------
a This period would extend through the first nine months of 2014 under the alternative, reduced phase-in requirement (see Section III.B.1. for a
  description of the proposed alternative).

    We request comment on the proposed provisions to allow higher FELs 
on a limited number of Tier 4 engines, including whether the proposed 
allowance limits of 10 percent and 5 percent have been set at the right 
levels and whether the allowance to use a higher FEL cap is appropriate 
for the Tier 4 program. We also request comment on allowing 
manufacturers to use the allowances in a slightly different manner over 
the first four years. Instead of allowing manufacturers to certify up 
to ten percent for each of the first four years, manufacturers could 
certify up to 40 percent of one year's production but spread it out 
over four years in an unequal manner (e.g., 15 percent in the first and 
second years, and 5 percent in the third and fourth years). Last of 
all, we request comment on whether the allowance should be available 
for NOX during the years we a proposing a phase-in for the 
Tier 4 NOX standards. As proposed, we would not cover 
NOX during the phase-in years because manufacturers already 
can certify up to 50 percent of their engines to the Tier 3 
NMHC+NOX standards.
    Under the proposed Tier 4 program, for engines above 75 horsepower 
there will be two different groups of engines during the phase-in 
period. In one group, engines would certify to the applicable Tier 3 
NMHC+NOX standard (or Tier 2 standard for engines above 750 
horsepower), and would be subject to the ABT restrictions and 
allowances previously established for those tiers. In the other group, 
engines would certify to the 0.30 g/bhp-hr NOX standard, and 
would be subject to the restrictions and allowances in this proposed 
program. While engines in each group are certified to different 
standards, we are proposing to allow manufacturers to transfer credits 
across these two groups of engines with the following adjustment. As 
proposed, manufacturers could use credits generated during the phase-
out of engines subject to the Tier 3 NMHC+NOX standard (or 
Tier 2 NMHC+NOX standard for engines above 750 horsepower) 
to average with engines subject to the 0.30 g/bhp-hr NOX 
standard, but these credits will be subject to a 20 percent discount. 
In other words, each gram of NMHC+NOX credits from the 
phase-out engines would be worth 0.8 grams of NOX credits in 
the new ABT program. The ability to average credits between the two 
groups of engines will give manufacturers a greater opportunity to gain 
experience with the low-NOX technologies before they are 
required to meet the final Tier 4 standards across their full 
production. (The 20 percent discount would also apply to 
NMHC+NOX credits generated on less than 75 horsepower 
engines and used for averaging purposes with the NOX 
standards for engines greater than 75 horsepower.)
    We are proposing the 20 percent discount for two main reasons. 
First, the discounting addresses the fact that NMHC reductions can 
provide substantial NMHC+NOX credits, which are then treated 
as though they were NOX credits. For example, a 2010 model 
year engine (between 175 and 750 horsepower) emitting at 2.7 g/bhp-hr 
NOX and 0.3 g/bhp-hr NMHC meets the 3.0 g/bhp-hr 
NMHC+NOX standard in that year, but gains no credits. In 
2011, that engine, equipped with a PM trap to meet the new PM standard, 
will have very low NMHC emissions because of the trap, an emission 
reduction already accounted for in our assessment of the air quality 
benefit of this program. As a result, without substantially redesigning 
the engine to reduce NOX or NMHC, the manufacturer could 
garner a windfall of nearly 0.3 g/bhp-hr of NMHC+NOX credit 
for each of these engines produced. (Engines designed at lower 
NOX levels than this in 2010 can gain even more credits.) 
Allowing these NMHC-derived credits to be used undiscounted to offset 
NOX emissions on the phase-in engines in 2011 (for which 
each 0.1 g/bhp-hr of margin can make a huge difference in facilitating 
the design of engines to meet the 0.30 g/bhp-hr NOX 
standard) would be inappropriate. Second, the discounting would work 
toward providing a net environmental benefit from the ABT program, such 
that the more that manufacturers use banked and averaged credits, the 
greater the potential emission reductions overall.
    Some foreign engine manufacturers have commented that it is 
difficult for them to accurately predict the number of engines that 
eventually end up in the U.S., especially when they sell to a number of 
different equipment manufacturers who may import equipment. This would 
make it difficult for the engine manufacturer to ensure they are 
complying with the proposed NOX phase-in requirements for 
engines above 75 horsepower and the proposed PM phase-in requirements 
for engines above 750 horsepower. Therefore, we are proposing to allow 
engine

[[Page 28470]]

manufacturers to demonstrate compliance with the NOX phase 
in requirements for engines above 75 horsepower and the PM phase in 
requirements for engines above 750 horsepower by certifying ``split'' 
engine families (i.e., an engine family that is split into two equal-
sized subfamilies, one that generates a number of credits and one that 
uses an equal number of credits). In order to facilitate compliance 
with the proposed standards, we are proposing that this option be 
available to all engine manufacturers (i.e., both foreign and domestic 
manufacturers). Manufacturers would be allowed to certify split engine 
families with FELs no higher than the levels specified in Table VII.A-
3. The maximum NOX FEL values specified in Table VII.A-3 
were set at the level which would result in NOX ABT credits 
from engines above the Tier 4 standards offsetting ABT credits from 
engines below the previously applicable NMHC+NOX standards, 
including the 20 percent discount for using NMHC+NOX credits 
on Tier 4 engines. The maximum PM FEL value for engines above 750 
horsepower was set at the level halfway between the Tier 2 and proposed 
Tier 4 PM standard for engines above 750 horsepower. Manufacturers 
certifying split engine families would exclude those engines from end 
of the year ABT calculations (and therefore would not need to determine 
actual U.S. sales of such engine families for ABT credit calculation 
purposes). Manufacturers certifying split engine families would also 
exclude those engines from the calculations demonstrating compliance 
with the phase-in percentage requirements as well.

 Table VII.A-3.--Maximum FEL for Engine Families Certified as ``Split''
                             Engine Families
------------------------------------------------------------------------
                                                                Maximum
           Power category                   Pollutant          FEL,  g/
                                                                bhp-hr
------------------------------------------------------------------------
75 <= hp £175 (56 <= kW    NOX....................    \a\ 1.7
 <130).
175 <= hp <=750 (130 <= kW <560)...  NOX....................        1.5
hp £750 (kW X....................        2.3
 eq>560).
hp £750 (kW 560).
------------------------------------------------------------------------
Notes:
\a\ A limit of 2.5 g/bhp-hr would apply under the alternative, reduced
  phase-in requirement (see Section III.B.1. for a description of the
  proposed alternative).

    We are proposing one additional restriction on the use of credits 
under the ABT program. For the proposed Tier 4 standards we are 
proposing that manufacturers may only use credits generated from other 
Tier 4 engines or from engines certified to the previous tier of 
standards (i.e., Tier 2 for engines below 50 horsepower, Tier 3 for 
engines between 50 and 750 horsepower, and Tier 2 engines above 750 
horsepower). (As discussed in more detail below, we are proposing 
slightly different restrictions on the use of previous tier credits for 
engines between 75 and 175 horsepower.) We currently have a similar 
provision that prohibits the use of Tier 1 credits to demonstrate Tier 
3 compliance, and given the levels of the final Tier 4 standards being 
proposed today, we believe it is appropriate to apply a similar 
restriction. Otherwise, we would be concerned about the possibility 
that credits from engines certified to relatively high standards could 
be used to significantly delay the implementation of the final Tier 4 
program and its benefits.
    For reasons explained in Section III.B.1.b. of today's notice, we 
are proposing unique phase-in requirements for engines between 75 and 
175 horsepower in order to ensure appropriate lead time for these 
engines. Because of these unique phase-in provisions for engines 
between 75 and 175 horsepower, we are proposing slightly different 
provisions regarding the use of previous-tier credits. Under this 
proposal, manufacturers that choose to demonstrate compliance with the 
proposed phase-in requirements (i.e., 50 percent in 2012 and 2013 and 
100 percent in 2014) would be allowed to use Tier 2 NMHC+NOX 
credits generated by engines above 50 horsepower (along with any other 
allowable credits) to demonstrate compliance with the Tier 4 standards 
for engines between 75 and 175 horsepower during model years 2012, 2013 
and 2014 only. These Tier 2 credits would be subject to the power 
rating conversion already established in our ABT program, and to the 
20% credit adjustment we are proposing for use of NMHC+NOX 
credits as NOX credits. Manufacturers that choose to 
demonstrate compliance with the optional reduced phase-in requirement 
for engines between 75 and 175 horsepower, would not be allowed to use 
Tier 2 credits generated by engines above 50 horsepower to demonstrate 
compliance with the Tier 4 standards. (Use of credits other than banked 
Tier 2 credits from engines above 50 horsepower would still be allowed, 
in accordance with other ABT program provisions.) In addition, 
manufacturers choosing the reduced phase-in option would not be allowed 
to generate NOX credits from engines in this power category 
in 2012, 2013, and the first 9 months of 2014, except for use in 
averaging within this power category (i.e., no banking or trading, or 
averaging with engines in other power categories would be permitted). 
This restriction would apply throughout this period even if the reduced 
phase-in option is exercised during only a portion of this period. We 
believe that this restriction is important to avoid potential abuse of 
the added flexibility allowance, considering that larger engine 
categories will be required to demonstrate substantially greater 
compliance levels with the 0.30 g/bhp-hr NOX standard 
several years earlier than engines built under this option.
    Under this proposal, we are not proposing any averaging set 
restrictions for Tier 4 engines. An averaging set is a group of 
engines, defined by EPA in the regulations, within which manufacturers 
may use credits under the ABT program. In the current nonroad diesel 
ABT program, there are averaging set restrictions. The current 
averaging sets consist of engines less than 25 horsepower and engines 
greater than or equal to 25 horsepower. The restriction was adopted 
because of concerns over the ability of manufacturers to generate 
significant credits from the existing engines and use the credits to 
delay compliance with the newly adopted standards. (See 63 FR 56977.) 
We believe the proposed Tier 4 standards are sufficiently protective to 
limit the ability of manufacturers to generate significant credits from 
their current engines. In addition, we believe the proposed FEL caps 
provide sufficient assurance that low-emissions technologies will be 
introduced in a timely manner. Therefore, under this proposal, 
averaging would be allowed between all engine power categories without 
restriction effective with the Tier 4 standards. The averaging set 
restriction placed on credits generated from Tier 2 and Tier 3 engines 
would continue to apply if they are used to demonstrate compliance for 
Tier 4 engines.
    As described in section III.B.1.d.i. of today's notice, we are also 
proposing a separate PM standard for air-cooled, hand-startable, direct 
injection engines under 11 horsepower. In order to avoid potential 
abuse of this standard, engines certified under this proposed 
requirement would not be allowed to generate credits as part of the ABT 
program. Credit use by these engines

[[Page 28471]]

would be allowed. The restriction should be no burden to manufacturers, 
as it would apply only to those air-cooled, hand-startable, direct 
injection engines under 11 horsepower that are certified under the 
special standard, and the production of credit-generating engines would 
be contrary to the standard's purpose.
    The current ABT program contains a restriction on trading credits 
generated from indirect injection engines greater than 25 horsepower. 
The restriction was originally adopted because of concerns over the 
ability of manufacturers to generate significant credits from existing 
technology engines. (See 63 FR 56977.) Under this proposal, we are not 
proposing the restriction which prohibits manufacturers from trading 
credits generated on Tier 4 indirect fuel injection engines greater 
than 25 horsepower. Based on the certification levels of indirect 
injection engines, we do not believe there is the potential for 
manufacturers to generate significant credits from their currently 
certified engines against the proposed Tier 4 standards. Therefore, we 
are not proposing to restrict the trading of credits generated on Tier 
4 indirect injection engines to other manufacturers. The restriction 
placed on the trading of credits generated from Tier 2 and Tier 3 
indirect injection engines would continue to apply in the Tier 4 
timeframe.
    We are not proposing to apply a specific discount to Tier 3 PM 
credits used to demonstrate compliance with the Tier 4 standards. PM 
credits generated under the Tier 3 standards are based on testing 
performed over a steady-state test cycle. Under the proposed Tier 4 
standards, the test cycle is being supplemented with a transient test 
(see Section III.C above and VII.F below). Because in-use PM emissions 
from Tier 3 engines will vary depending on the type of application in 
which the engine is used (some having higher in-use PM emissions, some 
having lower in-use PM emissions), the relative ``value'' of the Tier 3 
PM credits in the Tier 4 timeframe will differ. Instead of requiring 
manufacturers to gather information to estimate the level of in-use PM 
emissions compared to the PM level of the steady-state test, we believe 
allowing manufacturers to bring Tier 3 PM credits directly into the 
Tier 4 time frame without any adjustment is appropriate because it 
discounts their value for use in the Tier 4 timeframe (since the 
initial baseline being reduced is probably higher than measured in the 
Tier 2 test procedure).
3. Should We Expand the Nonroad ABT Program To Include Credits From 
Retrofit of Nonroad Engines?
    We are considering expanding the scope of the standards by setting 
voluntary new engine standards applicable to the retrofit of nonroad 
diesel engines, and allowing these nonroad diesel engines to generate 
PM and NOX credits available for use by other nonroad diesel 
engines. This program could achieve greater emission reductions of 
these pollutants than could otherwise be achieved, in a cost-effective 
manner. Specifically, we would allow existing in-use nonroad diesel 
engines that are retrofitted to achieve more stringent levels of 
emissions than are otherwise required to generate credits available for 
use in the ABT program by new nonroad engines. Credit-generating 
engines electing to participate in the program would be considered new 
nonroad diesel engines, subject to the normal compliance mechanisms 
applicable to other new nonroad diesel engines. These new nonroad 
engines could generate credits that could be used in the ABT program 
for other new nonroad diesel engines. Any such program would also have 
to ensure that credits are surplus, verifiable, quantifiable, and 
enforceable. We request comment on whether such a program would be 
feasible and appropriate for the Tier 4 nonroad standards, and on how 
such a program might be structured.
    We are considering an approach for credit generation based on the 
use of advanced exhaust emission control technology/engine system 
combinations that would provide significant emissions reductions. To 
accomplish this, simple changes that are easy to circumvent 
accidentally or to defeat intentionally would not be eligible to 
generate credits, and essentially, only changes involving introduction 
of post combustion emissions control technology would be eligible. 
Thus, we would structure the program such that engine recalibration as 
the sole mechanism to reduce emissions would not be eligible for 
retrofit credits. Also, as noted, for purposes of a nonroad retrofit 
ABT program, in order to generate credits, the manufacturer of the 
nonroad retrofit engine system choosing to participate in the program 
would accept that the retrofit engine would be considered a new nonroad 
engine, subject to enforceable standards and normal certification and 
compliance requirements. We have outlined in a memorandum to the docket 
our ideas for meeting these objectives, including possible ways to 
structure the program.\299\ This memorandum describes potential 
procedures for credit generation, credit use, and a number of 
compliance, implementation, and enforcement measures.
---------------------------------------------------------------------------

    \299\ Memorandum to the Docket, Chris Lieske and Joseph 
McDonald, EPA, Additional Information on Nonroad Retrofit Engine ABT 
Credit Concepts, Docket A-2001-28.
---------------------------------------------------------------------------

    We recognize that expanding the ABT program in this way would 
introduce new issues and complexities to the nonroad Tier 4 program, 
and that there are several ways to structure the program. We are 
seeking comment on whether such an expansion of the ABT program is 
feasible and appropriate, as well as on the details of how a program 
could be structured. We have considered and described a possible 
framework for nonroad retrofit credits in an effort to help commenters 
provide input. The level of detail provided below and in the memorandum 
to the docket does not indicate that we have made any decisions on 
whether nonroad retrofit credits are appropriate for the ABT program or 
about how the program should function. We invite comment not only on 
the provisions described below and in the memorandum to the docket, but 
also on alternative approaches that commenters believe would lead to a 
better overall program.
    We are also seeking comment on the timing of a retrofit credits 
approach. We believe that if such a program were adopted, credit 
generation could start in 2004 at the earliest, and request comment on 
ending the program in the 2015 time frame. We view this as primarily a 
transitional program which could be most useful in the early years of 
the nonroad program. Ending the program in 2015 may also ease concerns 
about long-term impact of such a program on the environment.
    We encourage commenters to carefully address all aspects of a 
nonroad retrofit credits program including its usefulness, feasibility, 
compliance and enforcement measures, environmental benefits, and 
potential cost savings. We specifically request comment on the 
potential for such a program to provide additional emissions reductions 
than would otherwise be obtained and request comment on the potential 
impacts such provisions would have on emissions reductions associated 
with the proposed nonroad standards. We are also interested in comments 
on practical issues and details regarding how the program would operate 
and be enforced.
    a. What would be the environmental impact of allowing ABT nonroad 
retrofit credits?

[[Page 28472]]

    We would structure any nonroad credit ABT program in a way that 
provides greater overall emissions reductions over the life of the 
group of nonroad engines involved than would otherwise be achieved. 
These additional overall reductions would be achieved by applying a 
discount of 20 percent to ABT retrofit credits that are used to meet 
nonroad standards. The result of applying a discount would be that each 
ABT retrofit credit generated would translate to less than one nonroad 
engine credit available for consumption in the nonroad program. For 
example, a discount of 20 percent would reduce the consumable credits 
by 20 percent. The discount would provide greater overall net emissions 
reductions from the use of an ABT retrofit program, and the amount of 
this environmental benefit would increase with increased use of the 
program. Also, applying a discount would be consistent with past Agency 
actions (see additional discussion in the memorandum to the docket 
noted above).
    A discount would be an essential element of the nonroad retrofit 
credit provisions, since one of our objectives if we promulgated such 
an expanded ABT program would be to create greater net emission 
reductions. The absence of a discount would result in no net 
environmental impact, as the generation of credits would lead to 
emissions reductions which would be offset by the increase in emissions 
when the credits were used. A discount would also serve to mitigate the 
potential for net environmental detriments due to uncertainties in 
credit calculation and use.
    We request comment on whether a discount of 20 percent would be 
appropriate given the expectation that the discount will generate cost-
effective emissions reductions that would otherwise not occur, as well 
as the more prevalent uncertainties associated with trading credits 
between nonroad retrofits and new nonroad engines.
    b. How would EPA ensure compliance with retrofit emissions 
standards?
    If this program were adopted, we would expect to require the 
retrofit manufacturer to specify all emissions related maintenance and 
to list the type of fuel used to certify its retrofit-engine system and 
whether a particular fuel sulfur level is necessary to meet the 
standard and to maintain emissions compliance of the retrofit-engine 
system in-use. If such a fuel is necessary to maintain emissions 
compliance in-use, EPA would also consider the fuel to be ``critical 
emission related scheduled maintenance'' under a retrofit engine 
program. As a result of such classification, the manufacturer would be 
required to demonstrate that proper fueling will be performed in-use. 
Such a demonstration would include a showing that the required fuel is 
available to, and would be used by, the ultimate consumer or fleet 
operator receiving the retrofitted engines. Such retrofitted engines 
would also have to be labeled appropriately to reflect the new engine 
family and may also require labeling for the type of fuel to be used. 
In general, we would require the manufacturer to submit a plan for 
implementing all relevant aspects of the retrofit to ensure proper 
installation and emissions compliance throughout the useful life 
period. A full discussion of compliance issues and possible compliance 
provisions, such as recall, in-use testing, useful life, and warranty 
is provided in the memorandum to the docket, noted above. We request 
comment on these approaches for ensuring in-use compliance with 
possible nonroad retrofit emissions standards and requirements.
    c. What is the legal authority for a nonroad ABT retrofit program?
    Allowing use by new nonroad engines of credits generated by 
retrofit of in-use nonroad engines is justified legally as an aspect of 
EPA's standard setting authority. As we envision a program, a retrofit 
nonroad engine would be considered to be a new nonroad engine when the 
manufacturer opts into a voluntary retrofit program (if established). 
Upon such opt-in, this new engine would be subject to enforceable 
standards under CAA section 213, somewhat similar to opting into the 
voluntary Blue Sky series standards (see Section VII.E.2). Thus, the 
generation of credits by nonroad retrofits and their use by new engines 
subject to Tier 4 would be similar to conventional ABT. Put another 
way, the generation of credits by retrofitting in-use non-road engines 
and their subsequent use by new nonroad engines subject to the Tier 4 
standards is an averaging program involving emission credits generated 
by one type of new nonroad engine and used by other new nonroad 
engines, similar to conventional ABT programs. With a nonroad retrofit 
credit program, and the emissions reductions associated with it, the 
overall emission reductions from Tier 4 nonroad engines and nonroad 
retrofit engines, taken together, would be the greatest achievable 
considering cost, noise, safety and energy factors, and would also be 
appropriate after considering those same factors. See also NRDC v. 
Thomas, 805 F.2d 410, 425 (D.C. Cir. 1986) (averaging provisions upheld 
against challenge that they are inconsistent with NCP provisions), and 
Husqvarna AB v. EPA, 254 F.3d 195, 202 (D.C. Cir 2001) (averaging, 
banking, and trading provisions cited as an element supporting EPA's 
selection of lead time under section 213(b)). At the same time, we also 
note that the proposed standards are the greatest achievable (taking 
all statutory factors into account) and appropriate independent of the 
nonroad retrofit program, as explained elsewhere in this preamble.\300\
---------------------------------------------------------------------------

    \300\ There is one minor exception to this analysis. Retrofits 
involving use of new nonroad engines as replacement engines in older 
nonroad equipment would be justified primarily as an aspect of EPA's 
lead time authority under section 213(d). This is because credits 
would not be generated from an engine certifying to a more stringent 
standard, so that the credit is effectively generated by equipment 
rather than by an engine, i.e. generated by something other than a 
new non-road engine.
---------------------------------------------------------------------------

B. Transition Provisions for Equipment Manufacturers

1. Why Are We Proposing Transition Provisions for Equipment 
Manufacturers?
    As EPA developed the 1998 Tier 2/3 standards for nonroad diesel 
engines, we determined that provisions were needed to avoid unnecessary 
hardship for equipment manufacturers. The specific concern is the 
amount of work required and the resulting time needed for equipment 
manufacturers to incorporate all of the necessary equipment redesigns 
into their applications in order to accommodate engines that have been 
redesigned to meet the new emission standards. We therefore adopted a 
set of provisions for equipment manufacturers to provide them with 
reasonable leadtime for the transition process to the newly adopted 
standards. The program consisted of four major elements: (1) A percent-
of-production allowance, (2) a small-volume allowance, (3) availability 
of hardship relief, and (4) continuance of the allowance to use up 
existing inventories of engines. See 63 at FR 56977-56978 (Oct. 23, 
1998).
    Given the level of the proposed Tier 4 standards, we believe that 
there will be engine design changes comparable in magnitude to those 
involved during the transition to Tier 2/3. We thus believe that at 
least some equipment manufacturers will face comparable challenges 
during the transition to the Tier 4 standards. This is confirmed by 
comments to EPA by a number of the equipment Small Entity 
Representatives during the SBREFA process, which indicated that the 
Tier 2/3 transition provisions were proving beneficial in providing 
adequate leadtime and urging

[[Page 28473]]

EPA to adopt comparable provisions in a Tier 4 rule. See Report of the 
Small Business Advocacy Review Panel, section 8.4.1 (Dec. 23, 2002). 
Therefore, with a few exceptions described in more detail below, we are 
proposing to adopt transition provisions for Tier 4 in this notice that 
are similar to those adopted with the previous Tier 2/3 rulemaking. The 
following section describes the proposed transition provisions 
available to equipment manufacturers. (Section VII.C. of today's notice 
describes all of the proposed provisions that would be available 
specifically for small businesses.)
    Our experience to date with the transition provisions for the Tier 
2/3 standards above 50 horsepower is limited. In the one power category 
where manufacturers have been required to submit information on the 
number of engines using the allowances (engines between 300 and 600 
horsepower), approximately 20 percent of the engines in the category 
are relying on the allowances in the first year that the Tier 2 
standards apply. (For the power categories below 50 horsepower, 
manufacturers are reporting that there are very few engines using 
allowances. However, given the level of the Tier 1 standards, we would 
not expect there to have been much need for equipment redesign to 
handle Tier 1 engines.) While this information is useful, we do not 
believe there is enough information available to determine if the level 
of the existing allowances should be revised for the Tier 4 proposal. 
For this reason, we are primarily relying on the provisions of the Tier 
2/3 equipment manufacturer transition provisions for the Tier 4 
proposal. However, as described in more detail below, we are proposing 
to add notification, reporting, and labeling requirements to the Tier 4 
proposal, which are not required in the existing transition provisions 
for equipment manufacturers. We believe these additional proposed 
provisions are necessary for EPA to gain a better understanding of the 
extent to which these provisions will be used and to ensure compliance 
with the Tier 4 transition provisions. We are also proposing new 
provisions dealing specifically with foreign equipment manufacturers 
and the special concerns raised by the use of the transition provisions 
for equipment imported into the U.S.
    As under the existing provisions, equipment manufacturers would not 
be obligated to use any of these provisions, but all equipment 
manufacturers would be eligible to do so. Also, as under the existing 
program, we are proposing that all entities under the control of a 
common entity, and that meet the definition in the regulations of a 
nonroad vehicle or nonroad equipment manufacturer contained in the 
regulations, would have to be considered together for the purposes of 
applying exemption allowances. This would not only provide certain 
benefits for the purpose of pooling exemptions, but would also preclude 
the abuse of the small-volume allowances that would exist if companies 
could treat each operating unit as a separate equipment manufacturer.
2. What Transition Provisions Are We Proposing for Equipment 
Manufacturers?
    a. Percent-of-Production Allowance
    Under the proposed percent-of-production allowance, each equipment 
manufacturer may install engines not certified to the proposed Tier 4 
emission standards in a limited percentage of machines produced for the 
U.S. market. Equipment manufacturers would need to provide written 
assurance to the engine manufacturer that such engines are being 
procured for the purpose of the transition provisions for equipment 
manufacturers. These engines would instead have to be certified to the 
standards that would apply in the absence of the Tier 4 standards 
(i.e., Tier 2 for engines below 50 horsepower, Tier 3 for engines 
between 50 and 750 horsepower,\301\ and Tier 2 for engines above 750 
horsepower). This percentage would apply separately to each of the 
proposed Tier 4 power categories (engines below 25 horsepower, engines 
between 25 and 75 horsepower, engines between 75 and 175 horsepower, 
engines between 175 and 750 horsepower, and engines above 750 
horsepower) and is expressed as a cumulative percentage of 80 percent 
over the seven years beginning when the Tier 4 standards first apply in 
a category. No exemptions would be allowed after the seventh year. For 
example, an equipment manufacturer could install engines certified to 
the Tier 3 standards in 40 percent of its entire 2011 production of 
nonroad equipment that use engines rated between 175 and 750 
horsepower, 30 percent of its entire 2012 production in this horsepower 
category, and 10 percent of its entire 2013 production in this 
horsepower category. (During the transitional period for the Tier 4 
standards, the fifty percent of engines that would be allowed to 
certify to the previous tier NOX standard but meet the Tier 
4 PM standard would be considered as Tier 4-compliant engines for the 
purpose of the equipment manufacturer transition provisions.) If the 
same manufacturer were to produce equipment using engines rated above 
750 horsepower, a separate cumulative percentage allowance of 80 
percent would apply to these machines during the seven years beginning 
in 2011. This proposed percent-of-production allowance is almost 
identical to the percent-of-production allowance adopted in the October 
1998 final rule, the difference being, as explained earlier, that we 
are proposing to have fewer power categories associated with the 
proposed Tier 4 standards.
---------------------------------------------------------------------------

    \301\ Under this proposal, for engines between 50 and 75 
horsepower, the NMHC+NOX standard that would apply in 
Tier 4 is the same as the existing Tier 3 NMHC+NOX 
standard.
---------------------------------------------------------------------------

    The proposed 80 percent exemption allowance, were it to be used to 
its maximum extent by all equipment manufacturers, would bring about 
the introduction of cleaner engines several months later than would 
have occurred if the new standards were to be implemented on their 
effective dates. However, the equipment manufacturer flexibility 
program has been integrated with the standard-setting process from the 
initial development of this proposal, and as such we believe it is a 
key factor in assuring that there is sufficient lead time to initiate 
the Tier 4 standards according to the proposed schedule.\302\
---------------------------------------------------------------------------

    \302\ For emissions modeling purposes, we have assumed that 
manufacturers take full advantage of the existing allowances under 
the transition program for equipment manufacturers in establishing 
the emissions baseline. This assumption is based on information 
provided to us by engine manufacturers for model year 2001, which 
shows that approximately 20 percent of the engines in the 300-600 
horsepower category are relying on the allowances in the first year 
that the Tier 2 standards apply. In modeling the Tier 4 program, 
because the program will not take effect for many years and it is 
not possible to accurately forecast use of the proposed transition 
program for equipment manufacturers and to assess costs in a 
conservative manner, we have assumed that all engines will meet the 
Tier 4 standards in the timeframe proposed. As discussed in section 
V.C., this is consistent with our cost analysis, which assumes no 
use of the proposed transition program for equipment manufacturers.
---------------------------------------------------------------------------

    Machines that use engines built before the effective date of the 
proposed Tier 4 standards would not be included in an equipment 
manufacturer's percent of production calculations under this allowance. 
Machines that use engines certified to the previous tier of standards 
under our Small Business provisions (as described in Section VII.C. of 
this proposal) would not be included in an equipment manufacturer's 
percent of production calculations under this allowance. All engines 
certified to the Tier 4 standards, including those engines that produce 
emissions at higher levels than the

[[Page 28474]]

standards, but for which an engine manufacturer uses ABT credits to 
demonstrate compliance, would count as Tier 4 complying engines and 
would not be included in an equipment manufacturer's percent of 
production calculations. As noted earlier, engines that meet the 
proposed Tier 4 PM standards but are allowed to meet the Tier 3 
NMHC+NOX standards during the phase-in period would also 
count as Tier 4 complying engines and would not be included in an 
equipment manufacturer's percent of production calculations. And, as 
also noted earlier, all engines used under the percent-of-production 
allowance would have to certify to the standards that would be in 
effect in the absence of the Tier 4 standards (i.e., the Tier 3 
standards for engines between 50 and 750 horsepower and the Tier 2 
standards for engines below 50 horsepower and above 750 horsepower).
    The choice of a cumulative percent allowance of 80 percent is based 
on our best estimate of the degree of reasonable leadtime needed by 
equipment manufacturers. We believe the 80 percent allowance responds 
to the need for flexibility identified by equipment manufacturers, 
while ensuring a significant level of emission reductions in the early 
years of the proposed program.
    We are also proposing to allow manufacturers to start using a 
limited number of the new Tier 4 flexibilities once the seven-year 
period for the existing Tier 2/Tier 3 program expires (and so continue 
producing engines meeting Tier 1 or Tier 2 standards). In this way, a 
manufacturer could potentially continue exempting the most difficult 
applications once the seven-year period of the current Tier 2/3 
flexibility provisions is finished. (Under the existing transition 
program for equipment manufacturers, any unused allowances expire after 
the seven year period. We are not reopening this provision with this 
proposal.) However, opting to start using Tier 4 allowances once the 
seven-year period from the current Tier 2/Tier 3 program expires would 
reduce the available percent of production exemptions available from 
the Tier 4 standards. We are proposing that equipment manufacturers may 
use up to a total of 10 percent of their Tier 4 allowances prior to the 
effective date of the proposed Tier 4 standards. (The early use of Tier 
4 allowances would be allowed in each Tier 4 power category.) This 
percentage of equipment utilizing the early Tier 4 allowances would be 
subtracted from the proposed Tier 4 allowance of 80 percent for the 
appropriate power category, resulting in fewer allowances once the Tier 
4 standards take effect. For example, if an equipment manufacturer used 
the maximum amount of early Tier 4 allowances of 10 percent, then the 
manufacturer would have a cumulative total of 70 percent remaining when 
the Tier 4 standards take effect (i.e., 80 percent production allowance 
minus 10 percent). We are also requesting comment on requiring 
equipment manufacturers to take a two-for-one loss of Tier 4 allowances 
for each allowance used prior to the Tier 4 effective date. This would 
reduce the number of overall engines that could be exempted under the 
Tier 4 allowance program and result in greater environmental benefits 
than would be realized if manufacturers used all of the Tier 4 
allowances in the Tier 4 timeframe.
    We view this proposed provision on early use of Tier 4 allowances 
as providing reasonable leadtime for introducing Tier 4 engines, since 
it should result in earlier introduction of Tier 4-compliant engines 
(assuming that the 80% allowance would otherwise be utilized) with 
resulting net environmental benefit (notwithstanding longer utilization 
of earlier Tier engines, due to the stringency of the Tier 4 standards) 
and should do so at net reduction in cost by providing cost savings for 
the engines that have used the Tier 4 allowances early. As discussed 
above, once the Tier 4 implementation model year begins, engines which 
use the transition provision allowances must be certified to the 
standards that would apply in the absence of the Tier 4 standards.
    b. Small-Volume Allowance
    The percent-of-production approach described above may provide 
little benefit to businesses focused on a small number of equipment 
models. Therefore we are proposing to allow any equipment manufacturer 
to exceed the percent-of-production allowances described above during 
the same seven year period, provided the manufacturer limits the number 
of exempted engines to 700 total over the seven years, and to 200 in 
any one year. As noted earlier, equipment manufacturers would need to 
provided written assurance to the engine manufacturer when it purchases 
engines under the transition provisions for equipment manufacturers. 
The limit of 700 exempted engines would apply separately to each of the 
proposed Tier 4 power categories (engines below 25 horsepower, engine 
between 25 and 75 horsepower, engines between 75 and 175 horsepower, 
engines between 175 and 750 horsepower, and engines above 750 
horsepower). In addition, manufacturers making use of this provision 
must limit exempted engines to a single engine family in each Tier 4 
power category.
    As with the proposed percent-of-production allowance, machines that 
use engines built before the effective date of the proposed Tier 4 
standards would not be included in an equipment manufacturer's count of 
engines under the small-volume allowance. Similarly, machines that use 
engines certified to the previous tier of standards under our Small 
Business provisions (as described in Section VII.C. of this proposal) 
would not be included in an equipment manufacturer's count of engines 
under the small-volume allowance. All engines certified to the Tier 4 
standards, including those that produce emissions at higher levels than 
the standards but for which an engine manufacturer uses ABT credits to 
demonstrate compliance, would be considered as Tier 4 complying engines 
and would not be included in an equipment manufacturer's count of 
engines under the small-volume allowance. Engines that meet the 
proposed Tier 4 PM standards but are allowed to meet the Tier 3 
NMHC+NOX standards during the phase-in period would also be 
considered as Tier 4 complying engines and would not be included in an 
equipment manufacturer's count of engines under the small-volume 
allowance. All engines used under the small-volume allowance would have 
to certify to the standards that would be in effect in the absence of 
the Tier 4 standards (i.e., the Tier 3 standards for engines between 50 
and 750 horsepower and the Tier 2 standards for engines below 50 
horsepower and above 750 horsepower).
    In discussions regarding the current small-volume allowance, some 
manufacturers expressed the desire to be able to exempt engines from 
more than one engine family, but still fall under the number of 
exempted engine limit. (Under the current rules, although equipment 
manufacturers are allowed to exempt up to 700 units over seven years, 
they must all use the same engine family. In many cases, a 
manufacturer's largest sales volume model does not even sell 700 units 
over seven years. As a result, the maximum number of units a 
manufacturer can exempt under the small-volume allowance is less than 
the 700 unit limit.) We are concerned, however, that allowing 
manufacturers to exempt engines in more than one family, but retaining 
the current 700-unit allowance, could lead to significantly higher 
numbers of engines being exempted from the Tier 4 program.

[[Page 28475]]

    Using data of equipment sales by equipment manufacturers that 
qualify as small businesses under Small Business Administration (SBA) 
guidelines, we have analyzed the effects of a small-volume allowance 
program that would set an exempted engine allowance lower than 700 
units over seven years but allow manufacturers to exempt engines from 
more than one engine family. Based on sales information for small 
businesses, we believe we could revise the small-volume allowance 
program to include lower caps and allow manufacturers to exempt more 
than one engine family while still keeping the total number of engines 
eligible for the allowance at roughly the same overall level as the 
700-unit program described above.\303\ Such a program would in general 
provide sufficient leadtime for equipment manufacturers, allowing them 
to temporarily exempt greater numbers of equipment models from the 
proposed Tier 4 standards, but, as noted above, keeping the total 
number of engines eligible for the allowance at roughly the same 
overall level as the existing program would allow (and so not allow 
more leadtime than necessary). Based on our analysis, the small-volume 
allowance program could be revised to allow equipment manufacturers to 
exempt 525 machines over seven years (with a maximum of 150 in any 
given year) for each of the three power categories below 175 
horsepower, and 350 machines over seven years (with a maximum of 100 in 
any given year) for the two power categories above 175 horsepower. 
Concurrent with the revised caps, manufacturers would be allowed to 
exempt engines from more than one engine family under the small-volume 
allowance program. Table VII.B-1 compares the proposed small-volume 
allowance program to the variation described in this paragraph.
---------------------------------------------------------------------------

    \303\ ``Analysis of Small Volume Equipment Manufacturer 
Flexibilities,'' EPA memo from Phil Carlson to Docket A-2001-28.

                            Table VII.B-1.--Small-Volume Allowance Program Comparison
----------------------------------------------------------------------------------------------------------------
                                                                            Maximum
                                                                            exempted
                                           Engines exempted over 7 years    engines      Single engine family
                                                                             in one          restriction?
                                                                              year
----------------------------------------------------------------------------------------------------------------
Proposed program........................  --700 for each power category..        200  --Yes
Variation under consideration...........  --525 for power categories <           100  --No
                                           175 hp.
                                          --350 for power categories  175 hp.
----------------------------------------------------------------------------------------------------------------

    We request comment on adopting a small-volume allowance program 
with the lower caps noted above that allows manufacturers to exempt 
more than one engine family in each power category. We specifically 
request comment on allowing equipment manufacturers to choose between 
the two small-volume allowance programs described above. Alternatively, 
we request comment on whether we should replace the current program 
(which allows 700 units over seven years with a one engine family 
restriction) with this revised small-volume allowance program (which 
would allow fewer units over seven years but without the single engine 
family restriction). Our analysis of small businesses noted above did 
show that there were a very limited number of companies that could 
potentially get fewer total allowances under a revised program with the 
lower caps compared to the existing program (i.e., a company that sells 
an equipment model that utilizes one engine family whose sales over a 
seven year period are above the revised limits noted above but less 
than 700). Allowing an equipment manufacturer to choose between the two 
programs would help to ensure that manufacturers are able to retain the 
current level of flexibility they have under the current program.
    Because we are proposing fewer power categories for the Tier 4 
standards, the proposed equipment flexibility program is designed to 
reflect those changes. Therefore, under the proposed small-volume 
allowance, the specified unit allowances will apply separately to each 
of the five power categories being proposed for the Tier 4 standards.
    As noted earlier, we are also proposing to allow manufacturers to 
start using a limited number of the new Tier 4 flexibilities once the 
seven-year period for the existing Tier 2/Tier 3 program expires (and 
so continue producing engines meeting Tier 1 or Tier 2 standards). 
Under the proposed small-volume allowance, any engines used by the 
manufacturer prior to Tier 4 would be subtracted from the proposed 700 
unit allowance (for the appropriate Tier 4 power category), resulting 
in fewer allowances once the Tier 4 standards take effect. As with the 
proposed percent-of-production allowance, we are proposing to limit the 
number of Tier 4 small-volume allowances that can be used prior to the 
effective dates of the Tier 4 standards to a total of 100 units in each 
of the Tier 4 power categories. We are taking comment on requiring 
equipment manufacturers to take a two-for-one loss of Tier 4 small-
volume allowances for each allowance used prior to the Tier 4 effective 
date. As explained above, we view this proposal as providing reasonable 
leadtime for introduction of Tier 4 engines by providing the 
possibility of earlier introduction of such engines with a net cost 
savings.
    c. Hardship Relief Provision
    We are proposing to extend the availability of the ``hardship 
relief provision'' with the Tier 4 transition provisions for equipment 
manufacturers. Under the proposal, an equipment manufacturer that does 
not make its own engines could obtain limited additional relief by 
providing evidence that, despite its best efforts, it cannot meet the 
implementation dates, even with the proposed equipment flexibility 
program provisions outlined above. Such a situation might occur if an 
engine supplier without a major business interest in the equipment 
manufacturer were to change or drop an engine model very late in the 
implementation process. As with other equipment manufacturer transition 
provisions, the equipment Small Entity Representatives indicated that 
the availability this allowance was useful to them in the transition to 
the Tier 2/3 standards, and they urged that it be continued in any Tier 
4 rule. Report of the Small Business Advocacy Panel, section 8.4.1.

[[Page 28476]]

    Applications for hardship relief would have to be made in writing, 
and would need to be submitted before the earliest date of 
noncompliance. The application would also have to include evidence that 
failure to comply was not the fault of the equipment manufacturer (such 
as a supply contract broken by the engine supplier), and would need to 
include evidence that serious economic hardship to the company would 
result if relief is not granted. We would work with the applicant to 
ensure that all other remedies available under the flexibility 
provisions were exhausted before granting additional relief, if 
appropriate, and would limit the period of relief to no more than one 
year. Applications for hardship relief generally will only be accepted 
during the first year after the effective date of an applicable new 
emission standard.
    The Agency expects this provision would be rarely used. This 
expectation has been supported by our initial experience with the Tier 
2 standards in which only one equipment manufacturer has applied under 
the hardship relief provisions. Requests for hardship relief would be 
evaluated by EPA on a case-by-case basis, and may require, as a 
condition of granting the applications, that the equipment manufacturer 
agree (in writing) to some appropriate measure to recover the lost 
environmental benefit.
    d. Existing Inventory Allowance
    The current program for nonroad diesel engines includes a provision 
for equipment manufacturers to continue to use engines built prior to 
the effective date of new standards, until the older engine inventories 
are depleted. It also prohibits stockpiling of previous tier engines. 
We are proposing to extend these provisions as manufacturers transition 
to the standards contained in this proposal. We are also proposing to 
extend the existing provision that provides an exception to the 
applicable compliance regulations for the sale of replacement engines. 
In proposing to extend this provision, we are requiring that engines 
built to replace certified engines be identical in all material 
respects to an engine of a previously certified configuration that is 
of the same or later model year as the engine being replaced. The term 
``identical in all material respects'' would allow for minor 
differences that would not reasonably be expected to affect emissions.
3. What Are the Recordkeeping, Notification, Reporting, and Labeling 
Requirements Associated With the Equipment Manufacturer Transition 
Provisions?
    a. Recordkeeping Requirements for Engine and Equipment 
Manufacturers
    We are proposing to extend the recordkeeping requirements from the 
current equipment manufacturer transition program. Under the proposed 
requirements, engine manufacturers would be allowed to continue to 
build and sell previous tier engines needed to meet the market demand 
created by the equipment manufacturer flexibility program, provided 
they receive written assurance from the engine purchasers that such 
engines are being procured for this purpose. We are proposing that 
engine manufacturers would be required to keep copies of the written 
assurance from the engine purchasers for at least five full years after 
the final year in which allowances are available for each power 
category.
    Equipment manufacturers choosing to take advantage of the proposed 
Tier 4 allowances would be required to: (1) Keep records of the 
production of all pieces of equipment excepted under the allowance 
provisions for at least five full years after the final year in which 
allowances are available for each power category; (2) include in such 
records the serial and model numbers and dates of production of 
equipment and installed engines, and the rated power of each engine, 
(3) calculate annually the number and percentage of equipment made 
under these transition provisions to verify compliance that the 
allowances have not been exceeded in each power category; and (4) make 
these records available to EPA upon request.
b. Notification Requirements for Equipment Manufacturers
    We are also proposing some new notification requirements for 
equipment manufacturers with the Tier 4 program. Under this proposal, 
equipment manufacturers wishing to participate in the Tier 4 transition 
provisions would be required to notify EPA prior to their use of the 
Tier 4 transition provisions. Equipment manufacturers would be required 
to submit their notification before the first calendar year in which 
they intend to use the transition provisions. We believe that prior 
notification will not be a significant burden to the equipment 
manufacturer, but will greatly enhance our ability to ensure 
compliance. Indeed, EPA believes that in order for an equipment 
manufacturer to properly use either of the allowances provided, it 
would already have the information required in the notification. Thus 
we are not requiring additional planning or information gathering 
beyond that which the equipment manufacturer must already be doing in 
order to ensure its compliance with the regulations. Under the proposed 
notification requirements, each equipment manufacturer would be 
required to notify EPA in writing and provide the following 
information:
    (1) The nonroad equipment manufacturer's name, address, and contact 
person's name, phone number;
    (2) the allowance program that the nonroad equipment manufacturer 
intends to use by power category;
    (3) the calendar years in which the nonroad equipment manufacturer 
intends to use the exception;
    (4) an estimation of the number of engines to be exempted under the 
transition provisions by power category;
    (5) the name and address of the engine manufacturer from whom the 
equipment manufacturer intends to obtain exempted engines; and
    (6) identification of the equipment manufacturer's prior use of 
Tier 2/3 transition provisions.
    EPA is requesting comment on whether the notification provisions 
should also apply to the current Tier 2/Tier 3 transition program, and 
if so, how these provisions should be phased in for equipment 
manufacturers using the current Tier 2/Tier 3 transition provisions. 
EPA believes such a notification provision could be implemented as soon 
as 2005 and requests comments on the appropriate start date should we 
adopt such a notification provision for equipment manufacturers for the 
Tier 2/Tier 3 transition program.
c. Reporting Requirements for Engine and Equipment Manufacturers
    As with the current program, engine manufacturers who participate 
in the proposed Tier 4 program would be required to annually submit 
information on the number of such engines produced and to whom the 
engines are provided, in order to help us monitor compliance with the 
program and prevent abuse of the program.
    We are proposing new reporting requirement for equipment 
manufacturers participating in the Tier 4 equipment manufacturer 
transition provisions. Under this proposal, equipment manufacturers 
participating in the program would be required to submit an annual 
written report to EPA that calculates its annual number of exempted 
engines under the transition provisions by power category in the

[[Continued on page 28477]] 

 
 


Local Navigation


Jump to main content.