You are on page 1of 152

Accepted Manuscript

High Temperature Materials for Heavy Duty Diesel Engines: Historical and
Future Trends

Dean Pierce, Allen Haynes, Jeff Hughes, Ron Graves, Phil Maziasz,
Govindarajan Muralidharan, Amit Shyam, Ben Wang, Roger England, Claus
Daniel

PII: S0079-6425(18)30104-X
DOI: https://doi.org/10.1016/j.pmatsci.2018.10.004
Reference: JPMS 538

To appear in: Progress in Materials Science

Received Date: 27 January 2018


Revised Date: 22 October 2018
Accepted Date: 23 October 2018

Please cite this article as: Pierce, D., Haynes, A., Hughes, J., Graves, R., Maziasz, P., Muralidharan, G., Shyam, A.,
Wang, B., England, R., Daniel, C., High Temperature Materials for Heavy Duty Diesel Engines: Historical and
Future Trends, Progress in Materials Science (2018), doi: https://doi.org/10.1016/j.pmatsci.2018.10.004

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers
we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting proof before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.
This manuscript has been authored by UT-Battelle, LLC under Contract No. DE-AC05-00OR22725

with the U.S. Department of Energy. The United States Government retains and the publisher, by accepting

the article for publication, acknowledges that the United States Government retains a non-exclusive, paid-up,

irrevocable, world-wide license to publish or reproduce the published form of this manuscript, or allow others

to do so, for United States Government purposes. The Department of Energy will provide public access to

these results of federally sponsored research in accordance with the DOE Public Access Plan

(http://energy.gov/downloads/doe-public-access-plan).

1
High Temperature Materials for Heavy Duty Diesel Engines: Historical and Future Trends

Dean Piercea, Allen Haynesa, Jeff Hughesb, Ron Gravesa, Phil Maziasza, Govindarajan Muralidharana, Amit

Shyama, Ben Wangb, Roger Englandc, Claus Daniela

a
Oak Ridge National Laboratory, 1 Bethel Valley Rd, Oak Ridge, TN, USA
b
Cummins, Inc., 500 Jackson St, Columbus, IN, USA
c
Valvoline, Inc., 100 Valvoline Way, Lexington, KY, USA

ABSTRACT

Heavy duty (HD) vehicles are projected to be the largest fuel-use subsector in transportation, with

current demand for diesel fuel projected to grow 30% by 2040. Historically, a primary strategy for increasing

diesel engine efficiency has been to increase peak cylinder pressure (PCP). However, increasing PCP imparts

greater mechanical and thermal loads on engine components and materials. In recent years, the material

property limits for many components have been reached and further increases in PCP above ~20 MPa have

been difficult, while still maintaining the necessary affordability and longevity of on-road HD diesel engines.

This paper reviews the historical evolution and major metallurgical advancements of high temperature

materials in HD on road diesel engines (10-15 L displacement) up to the current state of the art, focusing on

materials in the engine block, cylinder heads, pistons, valves, and exhaust components. These components

cover a wide range of material classes, including cast iron, ferritic steel, austenitic steel, titanium alloys,

nickel based super-alloys, and high temperature coatings. The microstructural degradation and failure

mechanisms of the materials associated with the complex mechanical and thermal loading during service are

discussed and key areas for future materials research are suggested that overcome technical barriers.

Key Words

diesel engines; cast iron; martensitic steel; austenitic steel; microalloyed steel; nickel based superalloys

2
ABSTRACT ................................................................................................................................... 2

1 INTRODUCTION ................................................................................................................... 6

2 DIESEL ENGINE BACKGROUND AND TECHNICAL OVERIVEW ................................... 8

2.1 Diesel Engines and Freight Transport .............................................................................. 8

2.2 Diesel Engine Technical Background ............................................................................ 10

2.3 Diesel Emissions Control and Alternative Fuels for Heavy Duty Engines ....................... 19

2.4 Summary of Heavy Duty Diesel Engine Materials and Limitations................................. 21

3 CAST IRON ENGINE BLOCKS AND CYLINDER HEADS ................................................ 23

3.1 Cylinder Heads – Design and Manufacturing ................................................................. 23

3.2 Engine Block – Design and Manufacturing .................................................................... 25

3.3 Cast Iron Evolution ....................................................................................................... 27

3.4 Life Limiting Mechanisms at Elevated Temperatures During Operation ......................... 38

3.5 Pathways for Future Materials Development .................................................................. 41

4 TURBOCHARGER AND EXHAUST MANIFOLD HOUSINGS .......................................... 43

4.1 Design and Manufacturing............................................................................................. 43

4.2 Current Exhaust Manifold and Turbocharger Housing Materials .................................... 45

4.3 Candidate Materials for Future Exhaust Manifold and Turbocharger Housings ............... 49

4.3.1 Thermal Properties .................................................................................................... 50

4.3.2 Static Strength........................................................................................................... 51

4.3.3 Fatigue Strength ........................................................................................................ 53

4.3.4 Creep Strength .......................................................................................................... 57

3
4.3.5 Oxidation Resistance ................................................................................................. 59

4.4 Life Limiting Mechanisms at Elevated Temperatures During Operation ......................... 65

4.5 Pathways for Future Materials Development .................................................................. 67

5 PISTONS .............................................................................................................................. 70

5.1 Evolution of Materials and Manufacturing ..................................................................... 70

5.2 Life Limiting Mechanisms at Elevated Temperatures During Operation ......................... 76

5.3 Pathways for Future Materials Development .................................................................. 79

6 TURBINE WHEEL ............................................................................................................... 82

6.1 Design and Manufacturing............................................................................................. 82

6.2 Current Turbine Wheel Material .................................................................................... 82

6.3 Candidate Materials for Future Turbine Wheels ............................................................. 86

6.4 Life Limiting Mechanisms at Elevated Temperatures During Operation ......................... 90

6.4.1 Inconel 713C ............................................................................................................ 90

6.4.2 Titanium Aluminide .................................................................................................. 92

6.5 Pathways for Future Materials Development .................................................................. 94

7 INTAKE AND EXHAUST VALVES .................................................................................... 95

7.1 Valve Design and Manufacturing................................................................................... 95

7.2 Exhaust Valve Materials Evolution ................................................................................ 97

7.2.1 Precipitation Strengthened Austenitic Steels .............................................................. 99

7.2.2 Nickel Based Superalloys ........................................................................................ 102

7.2.3 Intermediate Nickel Content Alloys ......................................................................... 104

7.3 Intake Valve Materials Evolution ................................................................................. 106

4
7.4 Life Limiting Mechanisms at Elevated Temperatures During Operation ....................... 107

7.4.1 Valve Seat Wear ..................................................................................................... 108

7.4.2 High Cycle Fatigue ................................................................................................. 113

7.4.3 Oxidation, Sulfidation Corrosion, and Guttering ...................................................... 114

7.4.4 Long Term Material Aging...................................................................................... 118

7.4.5 Failure Analysis Case Studies .................................................................................. 121

7.5 Pathways for Future Materials Development ................................................................ 123

8 SUMMARY AND CONCLUSIONS ................................................................................... 126

9 ACKNOWLEDGEMENTS ................................................................................................. 126

10 DISCLAIMER ................................................................................................................ 126

11 REFERENCES................................................................................................................ 127

5
1 INTRODUCTION

Heavy duty diesel (HDD) freight trucks (class 7-8) consumed ~4.6 quadrillion Btus of energy in the

US in the year 2014, or approximately 18% of all U.S transportation energy use [1]. The average fuel

economy of class 7-8 trucks was ~5.8 miles per gallon in 2014 and the amount of annual miles driven by

freight trucks in the US is projected to increase from 276 billion in 2016 to 332 billion in 2030, a ~20%

increase [2]. Therefore, improving freight truck efficiency could have enormous economic and environmental

benefits, including significantly reduced cost for the transport of goods, combined with the reduction and

avoidance of substantial greenhouse gas (GHG) emissions and pollutants such as NOx and particulate matter.

Enhancing diesel engine efficiency is one pathway to increase the fuel economy of HDD freight trucks.

Increasing peak cylinder pressure (PCP) has historically enabled increased diesel engine efficiency but also

increases temperatures and loads [3]. Consequently, in recent years, the materials limits within some

components have been reached and further increases in PCP above ~20 MPa and temperature have been

difficult, while still maintaining the necessary affordability and longevity of on-road heavy duty diesel

engines (HDDE). Longevity goals for HDDE typically approach two million miles.

The intent of this review paper is to describe the evolution of high temperature materials in HDDE,

the current state of the art, technical barriers of materials to achieving higher PCPs, and pathways for future

higher temperature diesel engine materials development. The review will aim to show how particular

materials innovations historically enabled higher PCPs, improved engine efficiencies, and reduced emissions.

This review will focus on materials used in select components directly exposed to the combustion gas stream

and/or high temperatures (e.g., cylinder heads, engine blocks, pistons, cylinder liners, valves, exhaust

manifolds, and turbochargers) in HD on-road diesel engines with displacement in the 10-15L range. Particular

attention will be paid to select cast irons, ferritic steels, precipitation strengthened austenitic steels, titanium

aluminides, and nickel based superalloys which are used in the aforementioned components. High

temperature diesel engine materials are exposed to highly complex and dynamic mechanical and thermal

loading cycles as well as oxidative and corrosive atmospheres. As such, the links between important material

degradation/failure mechanisms and microstructural /metallurgical parameters will be elucidated. This

approach will provide the fundamental understanding of the microstructural properties which are preventing

6
or limiting the use of current materials in future applications with higher peak cylinder pressures, higher

operating temperatures, and more corrosive/oxidizing environments. Subsequently, promising research areas

and strategies for materials development that could enable higher PCPs, operating temperatures, and resultant

efficiencies will be discussed.

Section 2 of this manuscript will provide background and technical overview of diesel engines. This

section will provide general energy consumption data for diesel engines used for freight transport as well as

the high level thermodynamic principles which are guiding increases in peak cylinder pressures and

temperatures in order to realize higher efficiencies. Sub-section 2.4 gives a brief overview of the current

materials used in components exposed to high temperatures, the limits of these materials in terms of PCP and

temperature, along with the major life limiting mechanisms for selected components and materials. After

section 2, the remaining sections will be devoted to reviewing materials used in individual components

exposed to high temperatures and/or exhaust gases. Each “component section” will provide a brief overview

of manufacturing processes used to fabricate the component, a detailed subsection covering the evolution and

current state of the art for materials used in that particular component, the material degradation and failure

mechanisms observed in that component during operation, and suggested areas for future materials research to

enable operation of that component at higher peak cylinder pressures and temperatures. Section 3 will treat

the engine block and cylinder head and be primarily devoted to gray cast iron and compacted graphite iron

(CGI). Section 4 will detail the materials used in exhaust manifolds and turbocharger housings, with a

primary focus on ductile cast irons and candidate materials to replace ductile cast iron in future engines.

Section 5 will be devoted to materials in pistons, which have historically been aluminium alloys, microalloyed

steels, and martensitic steels. Section 6 will detail the current and candidate materials for turbine wheels.

Current and historical materials for intake and exhaust valves, which have generally been manufactured from

martensitic steels (intake valves) and austenitic steels and nickel based super-alloys (exhaust valves), will be

discussed in Section 7. A brief summary and concluding remarks are contained in section 8.

7
2 DIESEL ENGINE BACKGROUND AND TECHNICAL OVERIVEW

2.1 DIESEL ENGINES AND FREIGHT TRANSPORT

Invented in the mid-1890s by Rudolph Diesel, the internal combustion engine that bears his name is

now the mainstay of freight transport and construction equipment worldwide. It is the most efficient

combustion engine on the planet for propelling vehicles and is the choice for applications where minimizing

fuel consumption is imperative. About 70% of all freight-moving trucks in the US are diesel powered, and

nearly all of the Class 7-8 large tractor trailer rigs use diesel engines [4]. Diesel’s successful engine

configuration of 1897 had an efficiency of 26%, and though modest-sounding we note that competing steam

engines of the time were 10% efficient [5]. The first diesel car in the US was a Packard in 1929 powered by a

Cummins engine from the company started by Clessie Cummins in 1919, which remains a leading

manufacturer of diesel engines today. In Europe, the Benz Company was producing diesel powered cars at

scale in 1936, many used in taxis because of their high fuel economy. Today’s production diesel engines for

large sea-going ships (oil tankers etc.) have a record-setting 51% efficiency [6], and recent short-term

demonstrations of experimental HDDE for on-road applications have reached 51% [7,8].

Even these impressive benchmarks are not at the theoretical efficiency limits for internal combustion

engines, and numerous factors mandate continuing improvement:

 Medium and HD vehicles1 account for over 22% of the US transportation energy consumption [1],

and are the fastest growing contributor to GHG emissions in the transportation sector [9].

Worldwide demand for diesel fuel is following this trend. By 2040, HD vehicles are expected to be

the largest fuel-use subsector in transportation, with truck and marine demand for diesel fuel growing

by about 30% from 2000-2040, as shown in Figure 1 [10]. Diesel fuel accounts for 90% of fuel

consumed by medium and heavy-duty highway vehicles [1]. In the US, even with an expected 15%

improvement in truck fuel economy, total energy use for freight is expected to rise 30% by 2040 [2].

1
Medium and heavy duty vehicles here indicate all vehicles greater than 8500 pounds gross vehicle
weight rating (GVWR). For regulatory purpose, they are all called heavy-duty vehicles and range from large
pickups and vans to tractor-trailer rigs weighing up to 80,000 pounds when fully loaded.

8
 HDDEs are likely to experience increased competition from alternative powertrains in the coming

decades for powering class 7-8 HD trucks for long-distance transport. However, significant cost,

technical, and infrastructure obstacles must be overcome before widescale implementation of electric

powertrains can occur which utilize hydrogen or batteries for energy storage can occur. For instance,

class 8 vehicle electrification is currently hindered by the relatively high cost and low specific energy

of current Li-ion battery technology (relative to the specific energy of diesel fuel), which could

necessitate significant compromises in vehicle range and/or payload capacity assuming current

driving patterns [11].

 Cost-effective freight movement is vital to national economies, with fuel cost and driver

compensation leading contributors to costs.

 The first-ever regulations for heavy vehicle fuel economy in the US were enacted in 2011, and

similar laws have been placed in other countries, invoking the legal requirement for higher

efficiency. A second more stringent phase of fuel consumption regulations was published in 2016, to

take effect in 2018 [12].

 Today’s engines do not achieve their peak efficiency over their full range of operating conditions, so

there is considerable opportunity for improvement.

 Diesel engines for heavy trucks are expected to have useful lives of over 1,000,000 miles.

9
Figure 1 – Projected world energy consumption in million oil-equivalent barrels per day (MBDOE) between 2000 and
2040 for specific transportation sectors. The largest world growth by volume is for HD vehicles. These data are
reproduced from [10].

2.2 DIESEL ENGINE TECHNICAL BACKGROUND

Advancements in diesel engine efficiency and reduced emissions have been enabled in part by

development of new materials, and the critical role of advanced materials is expected to continue. Most of the

technical paths to higher engine efficiencies require a higher combustion pressure (to achieve greater torque

per engine size) which in general, leads to higher component temperatures. Higher combustion pressures and

temperatures impose greater mechanical and/or thermal loads on engine components (e.g., engine blocks,

cylinder heads, pistons, valves, and exhaust components), which may exceed the limits of the materials in

these components. The characteristics and design parameters that govern the efficiency of diesel engines are

summarized here.

Diesel engines, in the family of reciprocating piston/crankshaft architectures, are distinguished by

their combustion process which relies on compressing the air in the combustion chamber to a high

temperature sufficient to ignite the diesel fuel when sprayed at high pressure into the combustion chamber.

Hence diesel engines operate via compression ignition (CI) in contrast to gasoline engines that use a spark to

10
ignite a pre-mixed volume of fuel and air (spark ignition, SI). In CI engines, combustion continues as the fuel

and air mix and the spray process is completed, which is termed diffusion burning. Diesel fuel has specific

reactive properties that support CI. With the CI engine process, the overall combustion always occurs fuel-

lean (meaning with excess air, or air-fuel ratio greater than stoichiometric), which favorably impacts engine

efficiency by virtue of the effect on the specific heat ratio (k) of the combustion products, but creates a strong

oxidizing environment for components exposed to these gases (e.g., pistons, heads, valves, exhaust

components, etc.). In larger CI engines for trucks, the fuel is typically injected via a single central injector that

has five or more holes which distribute the fuel radially. The combustion occurs around these multiple spray

plumes, creating temporal and spatial variations in chemistry and temperature that can create similar

variations in component temperatures.

A pressure-volume graph representation of a diesel engine operation is shown in Figure 2, noting the

key events in the cycle [13]. This diagram, based on a real diesel engine, shows that modern diesels operate

somewhat in-between an ideal Otto cycle (combustion and heat addition at constant volume), and the

theoretical Diesel cycle (constant pressure combustion/heat addition). For a given compression ratio, it can be

shown that constant volume heat addition (Otto cycle) yields higher efficiency.

Figure 2 - Pressure - volume process in a four-stroke turbocharged diesel engine noting key events [13]. Reprinted with
permission Copyright © 2017 SAE International. Further distribution of this material is not permitted without prior
permission from SAE.

11
Analysis of idealized engine cycles via First Law of Thermodynamics shows that the two most

impactful parameters governing Otto cycle efficiency are the compression ratio (r) (higher is better) and the

ratio of specific heats of the combustion products (k) (higher is better), as shown in Figure 3 and Equation (1)

[14]. Lean combustion in diesels gives a higher k. Diesels actually require high compression ratio for

compression ignition, and with diffusion dominated combustion their compression ratios are not limited by

combustion knock as is the case for SI engines. CI engines use the fuel injection quantity to control load and

do not throttle air flow as in SI engines, avoiding so-called pumping losses characteristic of SI engines. In

real engines, friction and heat transfer losses, plus structural limitations on peak cylinder pressure, set a

maximum practical limit on compression ratio. Figure 4 illustrates the relationship between compression ratio

and PCP. Figure 5 shows that the efficiency gains from increasing compression ratio by a few points [8],

which is a substantial mechanical adjustment, are typically single digit %, yet competitiveness in the market

places high importance on even this degree of improvement.

(1)

Figure 3 - Compression ratio (r) (ratio = maximum cylinder volume/minimum cylinder volume for piston stroke) vs
efficiency ( th, Otto). k is the ratio of specific heats of the working fluid (combustion products), cp/cv. Lean combustion in
diesels gives a higher k.

12
Figure 4 – Peak cylinder pressure vs. compression ratio for different RPM and amount of exhaust gas recirculation (EGR).
Data replotted from [3].

Figure 5 – Closed cycle efficiency gain as a function of peak cylinder pressure. Source: Cummins, Inc. and [8].

We can illustrate the relationship between engine compression ratio, ideal constant volume

combustion, and peak pressure/temperature by examining the thermodynamic path in an engine cycle. Since

the key compression and expansion processes in a piston engine are near isentropic, where pvk=constant,

logarithmic plots are instructive. Representative pressure-volume and pressure-temperature diagrams for

diesel engines are shown in Figure 6 and Figure 7 [13], respectively. In these thermodynamic processes, there

are well-defined relationships between the temperature and pressure of the air and combustion products in the

engine processes.

13
Figure 6 - Log P vs Log V plot of turbocharged diesel engine showing key events [13]. Since this is a turbocharged
engine, the intake pressure between events H and A is well above atmospheric pressure. The peak cylinder pressure
would be higher for higher compression ratio and higher manifold pressure, and also if the heat addition (combustion) all
occurred at the minimum volume (piston at top dead center). These measures would yield higher efficiency, yet create
higher peak cylinder pressures and temperatures. Reprinted with permission Copyright © 2017 SAE International. Further
distribution of this material is not permitted without prior permission from SAE.

Figure 7 - Log P vs Log T chart for the same turbocharged diesel engine as in Figure 6 [13]. It is apparent from both
Figures that if all of the heat release from the fuel occurred at the ideal constant volume, the peak temperature and
resulting peak pressure would be considerably higher. Reprinted with permission Copyright © 2017 SAE International.
Further distribution of this material is not permitted without prior permission from SAE.

14
Known and documented (e.g.,[15,16]) technical paths to higher engine efficiency suggest the need

for continued development of materials for major engine components. Key limitations are the maximum

cylinder pressure and the thermal loading, which are related as shown in Figure 6 and Figure 7 [13]. Figure 8

shows historic and projected trends in the efficiency of diesel engines for HD vehicles. Two of the

Department of Energy (DOE) funded SuperTruck teams were noted as incorporating higher PCP as an engine

efficiency strategy [7].

Figure 8- Historical (solid blue line [16]) and future projected (dotted red line [17]) of brake thermal efficiency for heavy-
duty on highway engines. Decreases in some years are due to new emissions regulations implemented in those years.
Green data points represent demonstrated and targeted brake thermal efficiency for SuperTruck I and II research
programs, respectively [18].

As already noted, higher compression ratio is a fundamental thermodynamic route to higher

efficiency and will result in higher peak cylinder pressure. Higher compression ratio increases the starting

pressure before combustion, creating higher peak pressure after combustion for a given fuel-air combustion

heat input. Calculations [13] showed that increasing compression ratio from 8 to 16 would essentially double

the PCP with accompanying modest increases in peak temperature, comparing at the same fuel-air ratio.

Caterpillar Inc. reported experimental data that showed a 1-3% reduction in fuel consumption by increasing

compression ratio from 18 to 22 [19]. However, higher fuel injection pressure was required and the increased

PCP was noted to challenge the engine durability.

15
For a given compression ratio, engines that approximate the Otto cycle will achieve maximum

efficiency when combustion and heat release of the fuel occurs at constant volume, when the piston is at the

top of its stroke and stationary. In practice, the combustion event occurs over 30-50 degrees of crankshaft

rotation (at least) and thus real peak cylinder pressures are reduced from theoretical. Ideally, it is optimal from

an efficiency standpoint to approach constant volume combustion and its high peak cylinder pressure. This is

optimized with fuel injection timing (or spark timing in gasoline engines), and is a trade-off with noise,

emissions, and durability. Advancements in diesel fuel injection over the decades have been numerous and

critical, mostly as seen in the increase in fuel injection pressure, now approximately 250 MPa. The rates of

injection and number of possible injection events per cycle (nine) have been increased and play a crucial role

in emissions control as well as achieving rapid combustion. Fully electronically controlled fuel injection for

truck diesel engines came in the mid-1980s with the Detroit Diesel Series 60 engine (whereas electronic fuel

injection for passenger cars came about 20 years earlier).

Down-sizing and down-speeding are measures to reduce parasitic losses such as friction and heat

transfer in proportion to engine output. Enabling these measures require increasing the PCP, power density,

and/or or torque density of the engine, also described by the brake mean effective pressure (BMEP) parameter

(engine torque normalized per engine displacement) in order to maintain equivalent performance of larger

engines or engines operating at higher RPM. Technical aspects and enabling technologies for down-sizing and

down-speeding are discussed in detail by Stanton [16]. For a given engine configuration (displacement,

compression ratio, bore/stroke), BMEP is increased by adding boosting (turbocharging) and and/or operating

at higher fuel-air ratios. Researchers at the Technical University of Braunschweig [20] modelled the impact

of taking BMEP to extreme levels (up to 8 MPa) as might be attempted with series boosting devices. They

found benefits in efficiency, yet the peak cylinder pressure and thermal loading imparted on the pistons may

result in unacceptable durability. In a study of piston cooling effects, U. Nottingham and Ford published

experimental data from a light-duty diesel engine showing increases in piston temperature of over 150 °C

(max of 330 °C) between engine operation at 0.2 MPa and 1.2 MPa BMEP [21]. Design trends toward higher

peak cylinder pressure and power/torque density are evident as seen in Figure 9 [3,16,17]. Computational and

modelling methods have continually advanced along with sheer computational power [22]. The importance

16
of computational methods that enable engine developers to meet the challenges of simultaneous efficiency

improvement and emission reduction cannot be overstated. Models have now advanced to the point of

interfacing combustion processes and performance with materials response via conjugate heat transfer

solutions [23]. The relationship between efficiency and materials conditions and capabilities can be

calculated. Prediction of emissions as well as performance is feasible, with models of catalytic after treatment

systems.

Figure 9 - Historical progression (solid lines) and predicted future (dotted lines) peak cylinder pressure (squares) and
power density (triangles) in HDDE (Red lines [16], Blue Lines [3], Green Lines [17]).

Measures to increase power/torque density usually implement turbocharging or supercharging on the

engine. Turbocharging was introduced for diesel engines in 1905 and was in widespread commercial use in

the 1940s [22]. These measures of boosting enable more torque and power from the same engine size, so

certain losses like heat transfer (heavily dependent on surface area of the combustion chamber), decrease in

proportion to the power output. In diesel engines, boosting can enable adequate power output at lower

(leaner) fuel-air ratios, favorably impacting the cp/cv parameter and also reducing heat losses. The relationship

between increased peak pressure and engine efficiency is observed consistently in models and practice (in real

engines) as shown in Figure 5.

17
Reducing the heat transfer losses from the combustion process through the pistons and cylinder head

remains an opportunity and a difficult challenge [16]. Measures to reduce this loss include increased power

density (reducing surface area per unit of fuel burned) and strategic use of materials that reduce the heat

transfer ([8], [24]). In most diesel engines, engine oil sprayed on the back side of pistons is used to cool the

materials and keep the temperatures within safe operating limits. The piston cooling requires oil pumping, a

parasitic loss, and it increases heat rejection through the piston to manage its temperature. Experiments [21]

have shown that reducing the oil cooling to reduce the heat loss impacted emissions but effects on efficiency

were not significant over drive cycles. Additional discussion on the impacts of piston cooling is contained in

section 5.

An additional, considerably effective, technology for increasing engine efficiency is the use of waste

heat recovery systems, which in the recent SuperTruck project demonstrations were organic Rankine cycle

(ORC) bottoming cycles as well as some use of turbocompounding. These types of systems, able to improve

engine efficiency by 3-8%, are expected to see growing application in large on-highway trucks. Using an

ORC system, diesel engines in the SuperTruck project exceeded 50% brake thermal efficiency in on-road

conditions. The energy audit for the Cummins engine was shared in [16] and is reproduced below in Figure

10. For full effectiveness of these systems, exhaust thermal energy must be preserved through the ports,

manifolds, and pipes, wherein low-thermal-loss properties of the materials are highly important.

Figure 10 - Fuel energy distribution for Cummins 15L engine in class 8 truck-trailer rig running at 65 mph on level road
[16]. SuperTruck technology demonstration of 50.5% brake thermal efficiency. Reprinted with permission Copyright ©
2017 SAE International. Further distribution of this material is not permitted without prior permission from SAE.

18
2.3 DIESEL EMISSIONS CONTROL AND ALTERNATIVE FUELS FOR HEAVY DUTY

ENGINES

Diesel engines would not enjoy their prominent place in commerce without remarkable achievements

to reduce criteria emissions by over 98% since 1988 (in the HD engine sector). This phenomenal reduction in

NOx, particle matter (PM), hydrocarbons, and CO has come from developments in combustion, aftertreatment

catalysts and diesel particulate filters, and some crucial changes in diesel fuel, particularly reduced sulfur2.

Although reducing sulfur level was primarily driven to avoid harm to emission control devices [25], in

subsequent sections of this paper we will also see significant impact on combustion chamber materials. The

timeline of progress and key technologies are depicted in Figure 11 [16]. Rapid advancements in high

performance computing simulation during this time were pivotal in the emissions reduction R&D [22].

Figure 11 – Evolution of diesel engine and aftertreatment technologies to meet emissions regulations. This figure is
reproduced and modified from [16]. Reprinted with permission Copyright © 2017 SAE International. Further distribution
of this material is not permitted without prior permission from SAE.

2
The EPA first regulated diesel fuel sulfur in 1993, mandating a maximum of 500 ppm. In practice,
the diesel fuel sulfur content averaged about 350 ppm until the next regulation in 2006 that lowered the S
content to 15 ppm. Prior to 1993, ASTM had a specification of maximum 5000 ppm sulfur and various
sources indicated the typical level was 2000-3000 ppm.

19
The use of selective catalytic reduction (SCR), on MHDV diesel engines became widespread in 2010

to meet the 2010 standard that required tailpipe NOx emissions of 0.27 g kW-1 h-1 (0.2 g HP-1 h-1). The

introduction and high NOx-reduction effectiveness of SCR allowed calibrating the engine to higher efficiency

in spite of increased engine-out emissions of NOx, yielding a 5% reduction in fuel consumption of the 2010

engines compared with the 2007 engines. However, a trade-off required an addition of about 2% (per unit

diesel fuel consumed) of the urea solution diesel exhaust fluid (DEF) required by the SCR system (Cummins

FleetGuard Bulletin 2009). Significant further diesel emission reductions in the future are also possible. The

potential to further reduce NOx emissions to levels below 0.027 g kW-1 h-1 (0.02 g HP-1 h-1) has been

demonstrated in laboratory testing [26–28], which is approximately an order of magnitude lower than current

regulations.

Diesel engines also represent a suitable avenue for the utilization of renewable fuels [29]. In the

United States, bio-diesel is currently blended in fuel for on-highway diesels at a national average of a few

percent, with some current engine configurations allowing up to 20% bio-diesel. “Renewable diesel”, a

heavily hydro-processed bio-oil, is currently suitable as a 100% replacement for petroleum diesel and is

utilized in some locations. The use of renewable diesel and bio-diesel can result in significant reductions in

GHG [30]. The use of alternative fuels such as bio-diesels and renewable diesel has been reported to reduced

PM but also result in slight increases of NOx emissions, depending on fuel type, engine configuration, and

emissions equipment (e.g.,[30–32]). The different chemical compositions and properties of alternative fuels

relative to petroleum diesel may have materials implications, resulting in different behavior in terms of

component friction and wear, and different exhaust gas compositions and/or temperatures, which could

influence the degradation mechanisms of some materials discussed in this review [33]. For instance, general

wear characteristics of engines operating on different biodiesel blends have been evaluated by analyzing the

content of metal particles in the engine oil after testing, then comparing to the same engine configuration but

operating on standard diesel fuel. These analyses were summarized by Haseeb et al. [33] and in most cases it

was shown that the levels of Fe, Al, and Cr were lower in the lubricant of engines operating on biodiesel,

indicating lower wear of components containing these metals. The present review will focus on materials in

HDDE exposed to high temperatures and/or combustion gases associated with petroleum diesel combustion.

20
2.4 SUMMARY OF HEAVY DUTY DIESEL ENGINE MATERIALS AND

LIMITATIONS

Table 1 summarizes the typical materials used in some critical HDDE engine components exposed to

exhaust gases and/or high temperatures, including cylinder heads, pistons, intake valves, exhaust manifolds,

turbocharger housings, and exhaust valves. A wide range of different materials classes are utilized in high

temperature HDDE components, including gray cast iron, ductile cast iron, ferritic steel, austenitic steel, and

Ni-base super-alloys. The materials for HDDE are selected based on their cost and ability to meet the

demands of the particular component.

Table 1 also summarizes the limits of PCP and temperature, if available, for components

manufactured from specific materials, as well as some of the major life limiting mechanisms that are acting at

elevated temperatures on each component during operation. High temperature components in HDDE are

subjected to complex and severe thermal and/or mechanical loading as well as corrosive and oxidative

environments. Life limiting mechanisms acting on these components include but are not limited to thermal

fatigue, constrained thermal fatigue (CTF), high cycle fatigue (HCF), low cycle fatigue (LCF), wear, creep,

microstructural aging, oxidation, sulfidation, and corrosion. In general, these life limiting mechanisms can

interact severely and become increasingly deleterious to materials as temperatures rise. The current

temperature limitations of all components listed in

Table 1 are at or only slightly higher than their current operating temperatures. Significantly

increasing HDDE efficiency will undoubtedly require increasing PCP, which will result in higher component

21
operating temperatures. In a 2013 DOE report, HDDE targets and metrics for power density, brake thermal

efficiency, peak cylinder pressure, and engine weight were established. These metrics and targets were

established for a baseline 2010 15L 475 HP engine operating at 42% thermal efficiency and peak cylinder

pressure of 190 bar. The targets and metrics from the DOE report are reproduced in Table 2. Also shown in

Table 2 are the estimated/projected peak metal temperatures at the exhaust valve and turbo inlet (note: these

temperatures are typically similar and were provided as a single temperature for this analysis) that would

occur under high engine loads for these notional engine conditions. The expected peak temperature at the

turbocharger inlet corresponding to an engine operating with a brake thermal efficiency of 60% and PCP of

30 MPa is 900 °C, 140 °C higher than the temperature limitation of the current material used in turbocharger

housing application, high SiMo ductile cast iron. Correspondingly, an exhaust valve temperature of 900 °C is

greater than the safe operating temperature of all current materials utilized in HDDE exhaust valve

applications. Other components which are not listed in Table 2 will also experience significant increases in

operating temperatures. The PCP and peak component temperature increases projected in

22
Table 2 will impose severe demands on engine components and materials. Therefore, many of the

materials that are currently used for HDDE components will need to be replaced by more durable materials in

order to enable these efficiency increases. As such, the development and/or implementation of new advanced

economically feasible materials for nearly all these components is required in order to enable higher cylinder

pressures and operating temperatures, along with continued increases in HDDE efficiency.

23
3 CAST IRON ENGINE BLOCKS AND CYLINDER HEADS

3.1 CYLINDER HEADS – DESIGN AND MANUFACTURING

Cylinder heads on HDDE in the 10-15 L are typically manufactured from gray cast iron. The

cylinder heads are complex cast structures which include numerous ports (intake and exhaust), coolant

passages, water jackets, and must undergo significant machining operations. An example of a recently

designed gray cast iron HDDE head is shown in Figure 12 [3]. Numerous considerations must go into

cylinder head design, including cost, castability, machinability, and complex thermal and structural analysis

[3,34,35]. Effective cooling is critical to keep the cylinder head temperatures below allowable limits,

particularly at the fire deck and at the exhaust-exhaust and exhaust-intake valve bridges, where failure due to

fatigue cracking is most likely to occur [35]. The fire deck of the heads in modern HDDE can reach

temperatures up to ~ 375 °C during operation [3] and significant temperature gradients are present owing to

hot exhaust gases exiting the exhaust ports, cooler air entering through the intake ports, and different cooling

characteristics at different locations on the head. Figure 13 shows the thermal gradients determined by

computational fluid dynamics for recently designed gray cast iron heads assuming 600 hp operation [3].

Temperatures are greatest at the exhaust-exhaust valve bridge. The high temperatures along with high

frequency cyclic stresses due to repeated combustion events, impose both low cycle thermo-mechanical

fatigue (TMF) (due to the engine start, operate, and stop cycles) along with superimposed HCF loading.

Computational fluid dynamics (CFD) and finite element analysis as well as other methods have recently been

used to predict fatigue lifetimes of cylinder heads [3,34–36]. At current PCP and temperatures, gray cast iron

in current cylinder head designs is near its mechanical limits. As peak cylinder pressures and temperatures are

increased in future HDDE in order to obtain enhanced efficiencies, cylinder head designs and/or materials will

need to be optimized.

24
Figure 12 – Top side of a cast and machined cylinder head manufactured from gray cast iron and designed for 25 MPa
peak cylinder pressure [3]. Reprinted with permission Copyright © 2017 SAE International. Further distribution of this
material is not permitted without prior permission from SAE.

Figure 13 – Temperature distribution (in °C) of the metal at the fire deck in HDDE gray cast iron heads based on CFD
simulations[3]. Reprinted with permission Copyright © 2017 SAE International. Further distribution of this material is not
permitted without prior permission from SAE.

Several approaches to improve the durability of next generation cylinder heads are being considered,

including re-design of existing cylinder head architecture [3] and/or the incorporation of materials with

improved high temperature capability, such as compacted graphite iron (CGI) [35–41]. Along the lines of the

former approach, Megel et al. [3] performed a parametric study in order to optimize the architecture of

cylinder heads with the goals to manufacture the heads from typical alloyed gray cast iron and enable

operation at 25 MPa PCP. The choice of alloyed gray cast iron rather than CGI was made due to the higher

thermal conductivity of gray cast iron, its wide industrial use, and reduced impact on existing machining lines.

25
The newly designed heads were cast using a hybrid ceramic/sand core process and common issues such as

shrinkage porosity, carbide formation, and high residual stresses were not significant. Successful casting and

benchtop fatigue testing were achieved in preparation for engine testing to validate the concept.

Manufacturing cylinder heads from CGI is an alternative approach to achieving improved durability

at high temperatures [35–41]. It has long been known that CGI exhibits improved mechanical behavior at

room temperature (RT) and elevated temperatures as well as different physical properties compared to gray

iron, in part due to different graphite morphology. This includes better fatigue strength, higher stiffness, and

lower thermal conductivity. The former contributes to improved durability at higher temperatures. Lower

thermal conductivity can result in higher component temperatures which may be detrimental to longevity but

also reduce heat loss. Higher stiffness can be beneficial in terms of structural rigidity of the heads.

3.2 ENGINE BLOCK – DESIGN AND MANUFACTURING

The cylinder block is the largest and heaviest component and the backbone of the engine, comprising

the cylinders, coolant passages, exhaust passages, etc. The crankshaft and camshaft are integrated into the

block and other components such as cylinder heads, exhaust manifolds, turbochargers, water pump, oil pump,

etc., are bolted directly to the block or to components connected directly to the block. A HDDE in the 10-15 L

range typically has six cylinders in-line. Cylinder liners serve as the inner wall of the cylinder and form a

sliding surface for the pistons and rings. The engine block is exposed to low cycle TMF due to the engine

start-operate-stop cycles as well as high cycle mechanical fatigue due to repeated combustion events. Static

strength and fatigue life at elevated temperatures is an important factor in selecting materials for these

components [42][39] as is wear resistance in the case of the cylinder liners [43,44]. It should be noted

however, that that the thermal loading on the engine block is not as severe as the cylinder heads.

Blocks and liners for HDDE in the 10-15L are typically manufactured from gray cast iron [45] or in

some cases, compacted graphite iron (CGI) [38] [37] [39]. Gray cast iron is widely used in automotive engine

blocks due to a combination of excellent castability and machinability, relatively low cost, and good

mechanical and thermal properties [46]. CGI exhibits better static strength, fatigue strength, higher stiffness,

and lower thermal conductivity as well as different physical properties compared to gray iron, in part due to

26
different graphite morphology. The higher fatigue strength of CGI compared to gray cast iron contributes to

improved durability and longevity at higher temperatures while the higher stiffness can be beneficial in engine

blocks to reduce bore distortion and ring tension during operation, resulting in reduced frictional losses and

improved efficiency and wear characteristics. The higher strength of CGI also enables the use of smaller and

thinner castings in some cases. Weight savings by switching from gray cast iron to CGI in blocks in the 10-

15L displacement range is reported to range from 10-15% [37]. CGI engine blocks are present in some lighter

duty applications (e.g., Cummins 5.0 L V-8 diesel and 6.7 L B engine) as well. Nonetheless, CGI can be more

difficult to machine [47], is more complex to cast, and is more expensive compared to gray cast iron.

Machinability is a major cost driver for manufacturing the block and heads and therefore must be

considered in the selection of cast irons. In general, CGI is significantly more difficult to machine compared

to gray cast iron. CGI typically has ~75% greater strength and 40% higher elastic modulus. During low speed

cutting (150-250 m min-1) with conventional carbide inserts or high speed carbide milling (400-800 m min-1)

with polycrystalline cubic boron nitride (PBCN), the CGI tool life is commensurate with the increases in

mechanical properties according to Dawson et al. [47]. However, the authors also noted that for high speed

(>700 m min-1) continuous cutting operations such as turning and cylinder boring operations, which are

present in modern high speed and high volume transfer lines, PBCN tool lives can be 10-20 times less for

CGI than gray iron. Therefore, the authors investigated the influence of metallurgical variables, such as

pearlite content, graphite shape, alloying elements (Si, Cr, Ti, and S) on the machinability of gray and CGI. A

major finding of the work was that the presence of sulfur contributes strongly to increased gray cast iron tool

life during high speed continuous cutting operations by forming a MnS protective layer on the cutting edge of

inserts. MnS is soft and pliable. In contrast, compacted graphite particles are only stable at low sulfur content

and CGI contains an order of magnitude less sulfur (0.005-0.025 wt.%) than gray cast iron (0.08-0.12 wt.%)

due to stricter processing requirements. Mg is added to liquid iron in the production of CGI to consume S and

oxygen. Mg is a stronger sulfide former than Mn, forming hard MgS and also Mg silicates. Figure 14 shows

flank wear vs. cutting length of PBCN inserts after turning at 800 m min -1 on a series of gray cast irons with

different sulfur content and a CGI. The wear of the PBCN inserts is significantly increased with decreasing

sulfur content for the gray cast irons and is highest for the CGI [47]. Other findings showed that additions of

27
Mo and Cr reduced the tool life of PCBN by a factor of 10 and that in the range of 70-100% pearlite, lower

pearlite levels improve turning operations but impair milling. In general, the authors suggested that in order to

improve the machinability of CGI, efforts should be focused primarily on developing improved machining

processes rather than metallurgical development with the goal of improving the machinability of CGI.

Figure 14 – Flank wear vs. cutting length for PBCN inserts during turning at speeds of 800 m min -1 [47]. Copyright ©
2017 SAE International. Further distribution of this material is not permitted without prior permission from SAE.

3.3 CAST IRON EVOLUTION

Gray cast iron that is used in HDDE heads typically consists of a ferrite-pearlite or pearlite

microstructure with lamellar or “flake-like” graphite particles. The composition ranges for select gray cast

irons and CGIs from the literature are shown in

28
Table 3 along with their RT UTS values. Significant research and development on gray cast iron has

been conducted over many decades (e.g., [46,48–55]). Important materials selection criteria for diesel engine

heads are thermal conductivity and resistance to high cycle mechanical fatigue, thermal cycling, and CTF.

High cycle mechanical fatigue occurs at operating temperatures from repeated combustion pressures acting on

the fire deck of the cylinder head. In addition, low cycle thermal fatigue can occur due to transients such as

the start-operate-shut down cycle of the engine. During this cycle, compressive stresses develop at the valve

bridges as temperature increase and are held for long periods of time, causing creep and stress relaxation to

occur in the material. As a result, when the engine is shut off and the heads cool, significant tensile stresses

can develop in the heads, resulting in plastic deformation, as well as the potential for crack opening and

propagation [34,56]. In many cases, as creep and stress relaxation become greater at high temperatures, larger

tensile stresses develop during cooling. Therefore, additional alloying to gray cast iron for cylinder head

applications is typical to increase high temperature compressive yield strength and improve resistance to

stress relaxation and creep.

It has been well-documented that additions of Mo increase the elevated temperature strength [49] and

thermal fatigue resistance of gray cast iron [48,53]. Wallace and Turnbull [53] investigated the influence of

Mo and Mo+Cr additions on the elevated temperature tensile and creep strength of gray cast iron. The authors

reported that tensile strength at 650 °C was improved from 69 to 207 MPa by the addition of 2 wt.% Mo. Mo

promotes elevated temperature strength in cast irons through solid solution strengthening, precipitation of fine

Mo carbides [57], and by refinement of the pearlite, owing to its role as a hardenability promoting element

[53]. When 0.6 wt.% Cr was synergistically added with Mo, additional increases in tensile strength at 650 °C

were reported but only for Mo content below 1.6 wt.%. With 0.6 wt.% Cr, and increasing levels of Mo above

~1.6 wt.%, free cementite (cementite that is not contained in pearlite) is formed, which embrittles the

material. During creep testing at 650 °C, Wallace and Turnbull [53] also noted that the pearlite in alloys with

0.6 wt.% Cr was significantly more resistant to spheroidization, which contributed to the improved creep

performance of the Cr containing gray cast irons. It should also be noted that too much Mo can induce micro-

shrinkage and be detrimental to mechanical properties. In 1982, Janowak and Gundlach [54] published a

comprehensive review on the beneficial and detrimental effects of alloying additions on processing,

29
microstructure, and properties of gray cast iron [54]. The authors noted that synergistic alloying of pearlite

promoting elements (Sn, Sb, Cu, and Cr) with elements that strongly promote pearlite refinement (V and Mo)

can result in greater strength enhancement with lower overall alloying amounts. More complex alloying, such

as combinations of Ni+Mo+Cr have been applied where Ni offsets the chilling tendency of Cr while refining

pearlite [54]. Compositions of some grade 250 and 300 gray cast irons containing synergistic alloying

additions of Cu+Sn+Cr and Cu+Sn+Cr+Mo, respectively, are reported in

30
Table 3. Higher alloying content is typically required in cylinder head applications, as compared to

the engine block due to higher thermo-mechanical loading.

More recently, research efforts have been focused on developing and evaluating CGI for cylinder

head and engine block applications [35–41,47,56,58–62]. CGI in HDDE heads typically exhibits a ferrite-

pearlite, or pearlite microstructure, with vermicular graphite particles. The microstructure of compacted

graphite iron with pearlitic matrix is shown in Figure 15. The graphite morphology of gray cast iron and CGI

exhibits significant differences. An SEM micrograph in Figure 16 shows the three-dimensional graphite

morphology of gray, CGI, and also ductile cast iron after deep etching for comparison. As observed in Figure

16, the graphite/matrix interfacial area will be much larger in gray cast iron due to the flake like graphite

morphology. As such, the differences in graphite morphology result in important differences in mechanical

and physical properties between gray cast iron and CGI, including static and fatigue strength, stiffness, and

thermal conductivity.

Figure 15 – Microstructure of a pearlitic compacted graphite iron. Figure reproduced from [40].

31
Figure 16 – Scanning electron micrograph of graphite morphology in gray, CGI, and ductile cast iron after deep etching
[47]. Reprinted with permission Copyright © 2017 SAE International. Further distribution of this material is not permitted
without prior permission from SAE.

The elastic stiffness and damping capacity are important properties in cylinder heads and engine

blocks. Higher stiffness can support reduced cylinder head and engine bore distortion. Cylinder head

distortion can result in improper valve seating while bore distortion can result in greater friction and ring

tension [63]. In contrast, however, higher stiffness also leads to reduced damping capacity, particularly in the

engine block. Higher damping capacity is often desirable from a noise, vibration, and harshness (NVH)

perspective. Measurements of the elastic modulus and damping capacity of a series of ductile, compacted, and

gray cast irons are presented in Figure 17 [63]. The stiffness of the cast iron is highly dependent on the

graphite morphology and strength, with stiffness decreasing from ductile, to CGI, to gray cast iron. The

damping capacity exhibits the opposite trend. In gray cast iron, the larger graphite-matrix interface area and

greater discontinuity of the matrix leads to lower stiffness values compared to CGI and ductile cast iron [38].

Figure 17 – Elastic modulus and damping capacity for a series of ductile (DI), CGI, and gray cast irons (GI) [63].
Reprinted with permission Copyright © 2017 SAE International. Further distribution of this material is not permitted
without prior permission from SAE.

32
Thermal conductivity of the water cooled cylinder heads plays an important role in the maximum

temperatures on the fire deck. The thermal conductivity of cast iron is dependent on the graphite

(morphology, orientation, and amount), the matrix (with ferrite having greater conductivity than pearlite), the

interface between the graphite and the matrix, alloying content (greater alloying results in more phonon

scattering sites and reduced thermal conductivity), section size (with gray cast iron more sensitive to section

size), and temperature [34,37,38,62,64–66]. In general, gray cast iron grades used in cylinder head

applications will have higher thermal conductivity than CGI grades for those applications [34], due to

differences in the graphite morphology, but not in all cases [64]. A schematic of heat transfer through steel,

ductile cast iron, and gray cast iron is shown in Figure 18a [64]. The thermal conductivity along the (0001)

basal planes of graphite has been reported to range from 293 to 2000 W m-1 K-1 compared to a range of ~10-

84 W m-1 K-1 along the prism planes [62,65]. In gray cast iron, the graphite grows along the a-direction of the

hexagonal unit cell which results in extended paths of high heat flow with minimal matrix interruption [65].

In compacted or spheroidal graphite, enhanced growth in the c-direction of the unit cell occurs at the expense

of the a-direction, causing thermal conduction to take a more tortuous path within the graphite particle and

longer pathways in the lower conductivity matrix occur. In Figure 18b, the thermal conductivities of cast irons

commonly used in cylinder head applications are shown (grades 250 and 300 gray cast iron as well as grades

350 and 450 CGI [62]). The matrix of the grade 250 and 300 gray irons and the grade 450 CGI are essentially

pearlitic, while the CGI 350 sample contains 48 vol. % ferrite. The thermal conductivity of grade 300 is

reduced compared to grade 250 gray cast iron, due to higher alloying content whereas CGI 450 is lower than

CGI 350 due to greater pearlite fraction, increased alloying content, and greater graphite nodularity. The

differences in thermal conductivity between gray and CGI become less as temperature increases. Although

not shown in Figure 18b, CGIs (and ductile cast irons) typically exhibit a significant decrease in thermal

conductivity with decreasing temperature below 100 °C in contrast to gray cast iron, as reported by other

authors [64,65]. It has been suggested that this is caused by lower thermal conductivity of the graphite-matrix

interface in cast irons containing vermicular or spheroidal graphite particles, potentially due to irregularities

and/or roughness at the interface [67], and that this component of thermal conductivity increases with

increasing temperature in CGI and ductile cast irons as improved contact results from thermal expansion.

33
Figure 18 – (a) Schematic of heat transfer through steel (left), ductile (middle) and gray cast iron (right) and (b) thermal
conductivity of gray cast iron and CGI as a function of temperature. (a) is modified from [68] and (b) is developed from
data in [62].

Comparisons of the mechanical properties between gray cast iron and CGI can be difficult due to the

many different compositions and strength levels for each, as shown in

34
Table 3. Ultimate tensile strength (UTS) ranges of gray cast irons and CGIs with pearlitic matrixes

typically range from 230-350 MPa and 410-580 MPa, respectively [38]. CGIs with ferritic matrixes exhibit

UTS in the range of 330-400 MPa. Figure 19 shows the RT unloading and loading stress strain curves in both

tension and compression for CGI (grade 450) and gray cast iron (grade 250) [34]. Both materials exhibit

significant anisotropy in their response to tension and compression loading, with the difference particularly

large for the gray cast iron. The CGI exhibits superior strength relative to the gray cast iron. The differences

in strength are primarily due to the graphite morphology. In gray cast iron, the sharp edges of the graphite

flakes contribute to crack nucleation and the flake surfaces provide cleavage planes for cracks to propagate,

resulting in much lower tensile strengths (when the cracks are open) [38]. Consequently, increasing the C

equivalent by 0.1% (promoting greater amounts of graphite and more crack initiation/propagation sites) in

class 30 gray iron reduces the tensile strength by 10% [38]. In contrast, the vermicular graphite particles in

CGI are more blunt, and localized stress concentrations are less severe near these particles compared to flake

graphite in gray cast iron. This results in fewer crack initiation sites and more difficult crack propagation. The

temperature dependence of the ultimate tensile strength for a series of gray cast irons and CGIs are show in

Figure 20. As is typical with materials containing substantial amounts of ferrite, with increasing temperature

above ~450 °C, strength decreases rapidly for both the gray cast irons and CGIs. In general, CGIs typically

maintain a strength advantage over gray cast iron at elevated temperatures, predominately due to graphite

morphology and the same factors that influence differences at RT. However, the temperature dependent

strength of gray cast iron and CGI is significantly influenced by alloying and microstructure, as evidenced by

Figure 20. For instance, a highly alloyed gray cast iron with nominal composition of 3.2C-2.1Si-0.73Mn-

0.7Cr-0.8Mo [53] exhibits higher UTS compared to a leaner alloyed 3.8C-2.4Si-0.4Mn-0.01Cu-0.1Sn-0.08Ni-

0.03Cr CGI over the temperature ranges shown in Figure 20. Furthermore, the highly alloyed gray cast iron

exhibits tensile strengths at temperatures greater than 450 °C that are equivalent or superior to the UTS values

of the other gray cast irons or CGIs shown in Figure 20.

35
Figure 19 – Unloading and loading curves for grades CGI (grade 450) and gray cast iron (grade 250) [34]. Copyright ©
2017 SAE International. Further distribution of this material is not permitted without prior permission from SAE.

Figure 20 – Temperature dependent ultimate tensile strengths for a series of gray cast irons (un-filled symbols) and CGIs
(filled symbols). These data are from [51,53,55,61,69].

Gray cast iron and CGI also exhibit different responses under TMF testing. TMF, when the thermal

loading is applied 180 degrees out of phase with the strain loading, is intended to approximate the temperature

and mechanical strain loading on components resulting from the engine start-operate-shut down cycle.

Norman et al. [69] performed 180 degree out of phase TMF testing on a non-commercial grade 250 alloyed

36
gray cast iron, a grade 400 commercial CGI (EN-GJV-400), and ductile iron grade SiMo5-1. During the tests,

the temperature was cycled from 100-400 °C or 100-500 °C and various TMF strain ranges (ΔεTMF) were

applied. In addition, the authors also performed tests in which HCF, with a strain range (ΔεHCF), was

superimposed on the low cycle TMF strain, yielding a total mechanical strain range of ΔεMech. The schematic

of these strains is shown in Figure 21. The frequency of 15 Hz for the HCF strain was chosen to be similar to

the rotational speed of the crankshaft. The cycles to failure, both as a function of ΔεTMF (with no applied

ΔεHCF) and maximum stress at half-life, are shown in Figure 22a and b, respectively [69]. The CGI exhibits a

small improvement over the gray cast iron under TMF conditions. However, the authors noted that the CGI is

significantly stronger than the gray cast iron and greater stresses develop in order to meet the prescribed

strain, as evidenced by the significantly larger maximum stresses (Figure 22b). Therefore, in actual operating

conditions (rather than ideal test conditions), it would be expected that the CGI material would experience

lower actual deformation compared to the gray cast iron, since the surrounding material is compliant rather

than rigid [69].

Figure 22c, d, and e show the number of cycles to failure as a function of ΔεHCF for grade 250 alloyed

gray cast iron, EN-GJV-400, and SiMo5-1, respectively. For each material, the influence of maximum

temperature (400 or 500 °C) of the cycle and ΔεMech (~0.43 or 0.59) is also shown. As shown in Figure 20, the

static strength of all the cast irons begins to decrease significantly above 450 °C. The results in Figure 22c, d,

and e also show a dramatic decrease in the number of TMF cycles to failure (reduced typically by a factor of

3-5 for values of ΔεHCF below ~0.075) as the maximum TMF cycle temperature is increased from 400 to 500

°C. As the maximum temperature of cylinder heads is currently near 400 °C in HDDE, expected increases in

temperatures above 400 °C in future diesel engines could have significant and detrimental influences on

fatigue life. As shown in Figure 22c, d, and e, ΔεHCF can be increased (while ΔεMech remains constant) to a

material dependent value, below which there is no significant detriment to the fatigue life. This value is

largest in the SiMo5-1 ductile cast iron, then the CGI, then the grade 250 gray cast iron. However, increasing

ΔεHCF above these threshold values results in large reductions in fatigue life. As peak cylinder pressures

increase in future HDDE, this may also increase HCF strain during operation and should be considered.

37
Park et al. [50] evaluated the elevated temperature fatigue properties of a series of gray and CGI. The

authors reported that CGI exhibited a significantly greater number of cycles to failure during CTF testing

compared to gray cast iron for equivalent Mo content, where samples were cycled from 100-500 °C or 100-

540 °C. In addition, the disparity in the number of cycles to failure increased when the maximum cycle

temperature was raised from 500 to 540 °C, indicating the CTF performance of the gray cast iron is more

adversely affected by increasing maximum cycle temperature. The authors reported a strong relationship

between CTF during cycling from 100-540 °C and both the tensile strength (at RT) and Mo content,

proposing the following relationship for the gray cast irons and CGIs studied in their work:

(2)

where the UTS is in ksi and N is the number of thermal cycles to failure. This relationship reflects

the importance of Mo on preventing creep and stress relaxation at elevated temperatures and therefore

reducing the maximum tensile stresses which occur during cooling that tend to open and propagate cracks.

The UTS term in Equation (2) reflects the materials ability to resist tensile loading during cooling under CTF.

Gray cast irons typically have much lower tensile strengths than CGI even at equivalent hardness [50], due to

the flake graphite morphology which enhances crack initiation and propagation during tensile loading. A

more recent study on the influence of Mo content of the CTF behavior of CGI was performed by Diaconu et

al. [59]. In the study by Diaconu et al. [59], CGI samples were cycled from 110 to 600 °C, where the CGI

samples contained between 0 and 1 wt.% Mo. Additions of Mo significantly enhanced the thermal fatigue

behavior of the CGI, which the authors attributed to refinement of the pearlite interlamellar spacing, an

increase in the UTS, and improved resistance to stress relaxation at 600 °C, which was identified in a follow

on study by the same authors [60]. The authors did however note that Mo alloying is not as effective at

improving the CTF performance in CGI compared to gray cast iron. As noted previously (see Figure 18)

however, gray cast irons typically exhibit greater thermal conductivity compared to CGI, which may result in

lower metal temperatures during actual operating conditions where the external surface of the cylinder head is

exposed to ambient temperatures. Therefore, Park et al. [50] also conducted a “finned disc thermal shock

test”, intended to account for differences in the thermal conductivity of the materials. In this test, samples

were thermally cycled by transferring specimens of gray cast iron and CGI between a high and low

38
temperature fluidized beds. Samples were placed in a high temperature bed (955 °C) for 5 s than cooled for

120 s in the low temperature bed set at 95 °C. Interestingly, under the “finned disc thermal shock test”

conditions, the gray cast irons exhibited a greater number of cycles before observation of the first major crack

and fewer cracks after a total of 2000 cycles. These results contradict those for CTF. The authors suggested

that the lower thermal conductivity of CGI results in steeper thermal gradients and higher maximum

temperatures, which facilitated cracking and caused poorer fatigue performance in the CGI samples under

these conditions. These results emphasize that thermal fatigue testing conditions have a strong influence on

material performance and that testing conditions of cast irons should replicate actual service conditions to the

maximum extent possible. In addition, because of the complexities, cost, and time required in replicating

actual service conditions in laboratory tests, computational fluid dynamics (CFD) and finite element analysis

as well as other method have recently been used to predict fatigue lifetimes of cylinder heads [3,23,34–

36,70,71].

Figure 21 – Schematic of simultaneously applied TMF and HCF strain in the study by Norman et al. [69]. This figure is
reproduced from that study.

39
Figure 22 – (a) TMF strain range (ΔεTMF) and (b) maximum stress at half-life as a function of cycles to failure for a non-
commercial grade 250 alloyed gray cast iron, a grade 400 commercial CGI (EN-GJV-400), and ductile iron grade SiMo5-
1 during 180 degree out of phase TMF testing with a temperature cycle from 100-500 °C. (c), (d), (e) Number of cycles to
failure as function of HCF strain range (ΔεHCF) during TMF+HCF testing for grade 250 gray cast iron, the CGI, and
SiMo5-1, respectively, for different values of total mechanical strain range (ΔεTMF+ΔεHCF) and temperature range of the
TMF cycle (100-400 °C or 100-500 °C). This figure is reproduced from [69].

3.4 LIFE LIMITING MECHANISMS AT ELEVATED TEMPERATURES DURING

OPERATION

A recent study by Norman et al. [69] investigated the damage mechanisms in gray cast iron (EN-

GJL-250), CGI (EN-GJV-400), and also a ductile cast iron SiMo5.1 under TMF. The temperature cycle was

varied from 100 to 500 °C [69]. The initial microstructures of these materials prior to TMF testing are shown

in Figure 23a, b, and c, respectively. The graphite morphology is lamellar in the gray iron, vermicular in the

CGI, and predominately spheroidal in SiMo5.1. Shrinkage pores are also observed in the microstructure of

SiMo5.1 in Figure 23c. The microstructures of the specimens after TMF testing for approximately half of the

lifetime of the specimens are shown in Figure 24. In Figure 24a and b, representative cracks emanating from

the tips of the lamellar and vermicular graphite particles are shown and were present in the bulk and near the

surface (the latter not shown). In contrast however, cracks were only observed to emanate from

microshrinkage pores in the SiMo5.1 alloy and not the spheroidal graphite particles, which will be discussed

40
further in section 4.4. The authors concluded that damage accumulation and failure in both EN-GJL-250 and

EN-GJV-400 during TMF in the temperature range from 100-500 °C occurs by microcrack initiation at the

graphite matrix boundary throughout the bulk material, crack propagation, and finally, coalescence and the

formation of a dominant macroscopic crack which ultimately leads to failure.

Figure 23 – Microstructures of (a) gray cast iron EN-GJL-250, (b) CGI grade EN-GJV-400, and (c) SiMo5.1. This figure
is reproduced from [69].

Figure 24 - Microstructures of (a) gray cast iron EN-GJL-250, (b) CGI grade EN-GJV-400, and (c) SiMo5.1 after
interrupting the samples at approximately half the fatigue life during a TMF test with temperature range of 100-500 °C
and TMF strain range of 0.35-0.37%. The arrows in the micrographs indicate cracks emanating from graphite particles in
(a) and (b) and microshrinkage pores in (c). This figure is reproduced from [69].

Ghodrat et al. have also investigated the nature of TMF behavior [41] as well as the influence of

holding time on TMF behavior [40] of a CGI. In both studies, the temperature was cycled between 50 and

420 °C and smooth and notched specimens were employed. The specimens were fully constrained. Specimens

that were pre-cycled, notched, and then TMF tested exhibited similar fatigue life to notched specimens of

virgin material, suggesting that the TMF crack growth rate is not dominated by accumulated plasticity in the

bulk but rather the plasticity in the crack tip. Therefore, notched specimens were found to be effective in

evaluating the TMF behavior. Figure 25 shows an optical micrograph of a notched specimen of CGI after

41
some degree of TMF loading. The graphite particles are the preferred sites for crack initiation both in the

notched specimens as shown in Figure 25 and smooth specimens (not shown). The authors noted that the

graphite particles act as internal notches and that crack growth begins almost immediately in the beginning of

TMF cycles. The authors concluded that TMF in CGI is governed by crack growth rather than crack initiation

and that crack growth can be approximated by the Paris model:

(3)

where is the crack growth rate (m cycle-1), m and C are material dependent parameters, and is the stress

intensity factor range.

The same authors investigated the influence of holding time at high temperature (420 °C in this case)

on the TMF behavior [40]. Holding time may be considered analogous to the operating time of an engine

during one start stop cycle. The authors found that increasing the holding time from 30 to 1800 s dramatically

reduces the fatigue life. Longer hold times result in greater stress relaxation during the high temperature

compression portion of the cycle. The stress relaxation is clearly evident in the stress-strain hysteresis loops

shown in Figure 26, where a significant reduction in the compressive stress occurs during the high

temperature hold. Furthermore, greater stress relaxation leads to larger maximum stresses (in tension) during

cooling of the specimen. Cracks are open in tension and larger stress equates to faster crack growth rates as

shown in Equation (3). Alloying strategies to resist creep and stress relaxation in alloyed gray cast iron were

discussed, such as Mo additions, and these have been applied to some extent in CGI alloys in order to further

increase fatigue strength [59,60].

42
Figure 25 – Cross section of a TMF specimen of CGI with 0.2 mm notch after TMF loading showing cracks emanating
from the notch tip and graphite particles. This figure is reproduced from [40].

Figure 26 – TMF stress-strain hysteresis loops for CGI specimen in total constraint during temperature cycling from 50 to
420 °C corresponding to TMF testing with a hold time of 1800 s at the elevated temperature. Maximum stresses in tension
tend to increase with increasing cycles before decreasing prior to failure. The mechanical strain range is 0.6 %. This figure
is reproduced from [40].

3.5 PATHWAYS FOR FUTURE MATERIALS DEVELOPMENT

Alloyed gray cast iron has historically been the material of choice for cylinder heads and engine

blocks in HDDE in the 10-15 L displacement range. However, significant increases in PCP and temperatures

are projected in the coming decades in order to meet efficiency and emissions targets (see

43
Table 2). In addition, significant further engine down-sizing is anticipated to be necessary to meet

these efficiency targets. The increased mechanical and thermal loads in the next 5 to 10 years may require a

re-design of alloyed gray cast iron heads [3] and/or new cylinder head materials. While the thermo-

mechanical loading on engine blocks is not as severe as on cylinder heads, stronger materials than current

gray cast iron could play an important role in engine weight reduction and more importantly for semi-truck

applications, down-sizing. CGI is currently the primary candidate material for replacing alloyed gray cast iron

in HDDE cylinder heads and engine blocks. CGI exhibits significantly improved elevated temperature

mechanical properties, such as static strength, fatigue strength, and stiffness, compared to alloyed gray cast

iron but its significantly higher processing costs and reduced machinability are factors currently preventing

widespread implementation. Approaches to improve the machinability are generally focused on alloy

development of CGI, development and optimization of insert materials for machining and cutting tools, and

improving the machining processes [47,72,73]. Alloy development of CGI for the purpose of improving

machining has typically resulted in improvements to machinability but with tradeoffs in mechanical properties

[47]. Consequently, recent research has also focused on developing new materials for cutting tools as well as

novel machining processes (e.g., laser assisted machining [73], modulation assisted machining [74], etc.).

Further research should be devoted to advanced processes and tools to more effectively machine CGI and at

lower cost.

As noted in

44
Table 2, as PCP is projected to rise to as high as 300 Bar, component temperatures will also increase

substantially. The low and high cycle thermo-mechanical fatigue resistance of current CGI grades (see Figure

22) is unlikely to be suitable under these long term projected increases in thermal and mechanical loading.

Therefore, longer term research should evaluate CGI for these projected increases in thermo-mechanical

loading in the coming decades and investigate and/or develop the next generation of cylinder head and engine

block materials beyond gray cast iron and CGI as required. These next generation materials should exhibit

significantly improved elevated temperature mechanical properties compared to current CGI grades, such as

static and dynamic strength, stiffness to be able to withstand more demanding environments and minimize

distortion, while being affordable, machinable, and easily manufacturable. These future materials would also

likely benefit from improved machining processes and tools. In addition, research should aim to decouple the

inverse relationship between thermal conductivity and strength in current cast irons, which can partially

negate the benefits of higher strength. For instance, covetic cast iron has shown initial promise in increasing

the thermal conductivity of baseline cast irons [75].

Because localized high temperatures at the fire deck surface contribute strongly to the failure of

cylinder heads, and particularly at the valve bridges, novel technologies to reduce these temperatures and/or

reinforce the cylinder heads at these locations should be evaluated. Thermal barrier coatings (TBC) are

effective at reducing surface temperatures of the base metal in order to extend the operating range of current

materials [76–80]. Typically, a TBC consists of an outer ceramic “top” coating with a low thermal

conductivity such as yttria stabilized zirconia, mullite, and gadolinium zirconate, which are under

investigation for diesel engine components [76–80] and an inner metallic “bond” coating, which provides

oxidation resistance and assists in adhesion of the ceramic layer to the metallic substrate. These coatings must

have excellent adhesion for the life of the component to eliminate the risk of spallation leading to significant

damage to the engine or downstream components. Therefore, methods to improve TBC adhesion, i.e.

resistance to spallation, should be investigated.

45
4 TURBOCHARGER AND EXHAUST MANIFOLD HOUSINGS

4.1 DESIGN AND MANUFACTURING

In HDDE, hot exhaust gases under pressure are expelled from the cylinders during the exhaust stroke

and are collected by the exhaust manifold. A portion of these gases flow through the high-pressure exhaust

gas recirculation loop and back into the intake while the remaining exhaust gases flow into the hot side of the

turbocharger. The hot gases flow through the turbocharger, spinning the turbine wheel, which in turn drives

the compressor wheel on the other end of the turbocharger shaft. The compressor wheel increases the

pressure of the intake air. The turbocharger converts usable energy from the hot pressurized exhaust gases to

work, so upon exiting the turbocharger, the exhaust gases are usually at significantly lower temperatures and

near ambient pressures (Figure 7). Figure 27 shows a schematic of the gas flow on the hot and cold sides of a

turbocharger. Figure 28 is a photograph of the exhaust side of a Cummins X15 performance series engine

installed in a Peterbilt 389 showing both the exhaust manifold and turbocharger. Upon exiting the

turbocharger, the exhaust gases then go through a series of emissions control equipment such as the diesel

particulate filter (DPF), selective catalytic reduction (SCR), etc. Figure 29 shows a schematic flow diagram of

a typical exhaust system in a HD on road diesel engine designed to meet EPA emissions criteria as of 2010.

Further information about emissions reduction equipment is provided in reference [7].

Figure 27 – Turbocharger schematic showing the compressor and turbine sides and the relative flows of hot gases through
the turbine [81].

46
Figure 28 – Photograph looking at the exhaust side of a Cummins X15 performance series engine installed in a Peterbilt
389 tractor trailer. Shown in the photograph is the exhaust manifold (unpainted) and turbocharger. The left (unpainted
side) of the turbocharger receives hot gases from the exhaust manifold. Image reused from [82].

Figure 29 – Schematic diagram of a typical emissions control system and equipment in on road HDDE for meeting EPA
regulations as of 2010. This figure is reproduced from [7].

The use of cast alloys to manufacture exhaust manifolds and turbocharger housings are currently the

most economical manufacturing route to produce these components due to their complex shapes. The exhaust

manifold is typically bolted to the heads and the turbocharger housing is attached to the exhaust manifold as

shown in Figure 28. These components are exposed to complex and unique thermal loads as well as hot

exhaust gases on their internal surfaces and ambient engine compartment air on their outer surfaces during

normal operation. Mechanical loads on the housing components include but are not limited to the following 1)

static loads due to component weight, 2) high cycle vibrations due to engine operation, 3) stresses due to

thermal gradients, and 4) low cycle thermally induced loads related to thermal expansion (or the constraint

thereof) when the components undergo large temperature swings such as what occurs during the start,

operation, shutdown cycle of the engine. In addition, due to the high operating temperature and nature of

47
diesel exhaust gases, oxidation of these materials is also an important consideration. Historically, ductile cast

iron has been the material of choice for the exhaust manifolds and turbocharger housings of HDDE and

provided adequate durability under the aforementioned operating conditions. However, high SiMo ductile cast

irons, which are currently the predominate materials used in turbocharger housings and exhaust manifolds, are

at their temperature limits for these applications and replacement materials are actively being investigated or

developed [83–86].

4.2 CURRENT EXHAUST MANIFOLD AND TURBOCHARGER HOUSING

MATERIALS

Numerous types of cast iron have been used in diesel engine exhaust manifolds and turbocharger

housings. The type of cast iron used often depends on the type of diesel engine, whether or not the manifold is

actively cooled, the operating temperature, and the mechanical properties and corrosion/oxidation resistance

required of the materials at those temperatures. In general, the maximum operating temperatures for cast

irons, considering both elevated temperature strength and oxidation resistance, increases from gray, to CGI, to

ductile, to Ni-resist [87]. Within those classes of cast iron, alloying and processing has historically been

utilized to optimize the elevated temperature properties. Modern HDDEs in the displacement range of 10-15L

for on road applications have traditionally used ductile cast iron with additions of Si (up to 4.2 wt.% [85])

and/or Mo (up to 0.9 wt.% [85]) for exhaust manifolds and turbocharger housings. The microstructure of

these cast irons is characterized by a ferritic matrix with the graphite particles taking on a spheroidal

morphology, as shown in the optical micrograph in Figure 30. In ductile cast irons for manifold and

turbocharger housing applications, pearlite in the matrix is sometimes considered detrimental as it can

undergo spheroidization and/or decomposition during the thermal cycling of the components. Pearlite

decomposition can cause expansion and increased thermal stresses during service [48]. For these reasons,

ductile cast irons are given a ferritizing anneal, typically above 900 °C, to stabilize the casting against any

significant permanent growth in service [46].

48
Figure 30 – Optical micrograph of the microstructure of a high SiMo ductile cast iron.

Ductile cast irons with Si and Mo additions designated for high temperatures were developed in the

late 1960s by the Climax Molybdenum company and commercialized in the 1970s [57,88]. A series of alloys

were developed containing 4 wt.% Si, 2.9-3.3 wt.% C, Mo additions of up to 2 wt.% [57], and smaller

additions of Al, B, Co, and/or Cr. The Si content of 4 wt.% was chosen to significantly raise the A 1

transformation temperature and to promote oxidation resistance while Mo was added to raise the creep

strength. These alloys contained ~0.3 wt.% Mn, which was less than the corresponding amount for typical

ductile irons at that time, and reductions in the Mn content helped to further increase the A1 temperature. The

authors reported that Mo significantly increased the tensile strength and 100 h rupture strengths at 650 and

815 °C, in part due to the stabilization of fine carbides in the matrix by Mo. As previously discussed, the

decomposition of carbides to graphite during service can result in thermal growth of the component and

adverse effects. A more recent study investigated the thermal stability of Mo-carbides in a high SiMo grade of

ductile cast iron [89]. The authors compared the thermal stability of Mo-carbides in the high SiMo (3.35C-

4.0Si-0.6Mo-0.45Mn-0.04Cu wt.%) ductile iron to the pearlite phase in a high Si (3.8C-3.0Si-0.22Mn-0.04Cu

wt.%) ductile iron [89]. The authors reported that Mo segregates to cell boundaries during solidification and

results in the precipitation of M6C and Fe2MoC carbides at the cell boundaries. Dilatometry results from this

study, shown in Figure 31, indicate that both alloys undergo similar thermal expansion up to 700 °C.

However, at the hold temperature of 725 °C, the high Si grade containing pearlite undergoes anomalous

49
expansion related to the decomposition of pearlite to ferrite and graphite. The authors indicated that the Mo-

carbides were stable even at temperatures up to 925 °C, which is above the A3 temperature.

1.2

1
3.8C-3.0Si-0.22Mn-0.04Cu

0.8
Expansion (%)

0.6

0.4

3.35C-4.0Si-0.6Mo-0.22Mn-0.04Cu
0.2

0
0 200 400 600 800
Temperature ( C)

Figure 31 – Expansion as a function of temperature obtained by dilatometry for high SiMo and high Si grades of ductile
cast iron [89]. The 3.0 Si ductile cast iron with no Mo additions undergoes anomalous expansion at temperatures above
700 °C owing to the decomposition of pearlite to graphite and ferrite. This data is reproduced from [89].

Two ductile irons are generally used in exhaust manifolds and turbocharger housings in modern 10-

15L HDDE for on-highway use, designated high SiMo and SiMo5.1. The actual compositions of high SiMo

and SiMo1.5 grades that were used in two recent investigations of the high temperature mechanical properties

are shown in Table 4 [85,90]. The SiMo5.1 grade provides a slight upgrade to the elevated temperature

mechanical properties and corrosion resistance compared to the high SiMo grade, owing to greater levels of Si

and Mo. Current manifolds and turbocharger housings manufactured from high SiMo are limited to operating

temperatures of ~760 °C in some 15L applications in order to enable the required longevity. In addition, metal

temperatures of ~650 °C and ~760 °C have been recorded on the exhaust manifold and on the turbo housing

inlet, respectively, in recent production 15 L engines. Therefore, at least in turbocharger housings, high SiMo

alloys are currently at their temperature limits for these applications. Future engine efficiency targets require

increases in PCP which will result in further increases in exhaust gas temperatures. In a 2013 Department of

Energy (DOE) report summarizing industry feedback from a materials workshop, peak metal temperatures on

50
the turbocharger inlet of approximately 700 °C expected under high engine load were reported for a baseline

2010 15L 475 HP engine at 42% thermal efficiency and peak cylinder pressure of 19 MPa (see

51
Table 2). As previously mentioned, approximated peak metal temperatures at the turbo inlet in some

current production 15 L engines can reach 760 °C. In the same DOE report, approximate peak metal

temperatures at the turbo inlet was projected to increase to 800 °C by 2025, 850 °C by 2030, and up to 875 °C

by 2040. The results from this report are reproduced in

52
Table 2 and assume corresponding and significant increases in PCP, engine power density, and

thermal efficiency. These projected temperature increases will exceed the current design limits of manifolds

and turbocharger housings manufactured from SiMo and SiMo5.1 for these applications. Thus, the evaluation

and/or development of new candidate cast materials for exhaust manifolds and turbocharger housings are

ongoing and aspects of these activities will be discussed next.

53
4.3 CANDIDATE MATERIALS FOR FUTURE EXHAUST MANIFOLD AND

TURBOCHARGER HOUSINGS

Various materials related strategies are being pursued in order to increase the operating temperature and

durability of next generation exhaust manifolds and turbocharger housings, including 1) alloying modifications to

existing high SiMo and SiMo5.1 grades of ductile cast irons, 2) the evaluation of existing materials as candidates for

these applications, 3) the development of new materials specifically for these applications, and 4) the development

of TBCs for ductile cast irons to enable them to operate at higher temperatures than would otherwise be possible. As

a result of the complex mechanical and thermal loads and oxidation that exhaust manifolds and turbocharger

housings are exposed to during operation, their materials are typically evaluated by a series of different tests. These

include several tests to evaluate high temperature strength, including static tensile testing, low cycle TMF, HCF, and

CTF. In addition, oxidation testing is performed at different temperatures and under different atmospheric

environments. The most realistic experiments replicate diesel engine combustion gases and include water vapor,

which has a significant additional detrimental effect on Fe-based alloys in this temperature range [91,92]. The

results of some recent investigations of these properties will be discussed in the forthcoming sections in order to

compare candidate materials and elucidate metallurgical parameters contributing to their performance. The actual

chemical compositions taken from specific studies of a series of alloys that are currently used in HDDE exhaust

manifolds and turbocharger housings or are under consideration for these applications are listed in Table 4. The high

SiMo and SiMo5.1 alloys were briefly discussed in section 4.2. SiMo1000 is a ferritic cast iron with alloying

additions of Si, Mo, Al, and Ni and the microstructure consists of ferrite, vermicular and spheroidal graphite,

pearlite, and pseudo pearlite. The composition of the carbide in the pseudo pearlite was measured to be ~Fe3AlCx

(x=0.5-0.7) [93]. Alloy 1.4509 is a low C 18 wt.% Cr cast ferritic steel. Ni-resist D5S is an austenitic cast iron with

33 wt.% Ni content that is typically used for exhaust manifolds of gasoline engines where exhaust gas temperatures

are higher [94]. Austenitic alloys HF and A3N are nominally 20Cr/10Ni austenitic steels of similar composition with

the exception that A3N contains Nb, W, and Co additions. CF8C-Plus is a 19Cr/12Ni austenitic steel recently

developed for exhaust manifolds and similar applications [90,95,96]. Unlike the other austenitic alloys considered

here for exhaust manifolds, CF8C-Plus is alloyed with N, which results in unique precipitation of M(C,N) and the

tetragonal Z-phase carbonitride precipitates. Austenitic alloy CN-12 Plus [97] contains nominally 25Cr/16Ni and

54
significantly higher C+N additions (~0.8 wt.%) than the other alloys examined here. Austenitic alloys HK 30, HK-

Nb, and TMA 4705 are nominally 25Cr/20Ni austenitic steels with similar composition but HK-Nb and TMA 4705

contain significant quantities of Nb [83,85]. Alloys HP and TMA 6501 are nominally 25Cr/35Ni austenitic steels.

CAFA is a developmental 14Cr/25Ni/3.5Al cast alumina forming austenitic steel (AFA). These steels are a new

class of material which form a highly protective and stable Al 2O3 oxide layer, making them well suited for high

temperature applications in harsh environments [98].

4.3.1 Thermal Properties

The thermal conductivity and thermal expansion are shown for a series of candidate ferritic and austenitic

exhaust manifold and turbocharger housing materials in Figure 32a and b [85,93,99], respectively. Since the

measurements reported in Figure 32a and b are from multiple studies, some systematic variation may be present.

High thermal conductivity can result in lower exhaust manifold and turbocharger housing temperatures during in

service operation, which can positively affect the components longevity. In general, the thermal conductivity tends

to increase with increasing temperature. The thermal conductivity of the ferritic materials, ductile cast irons SiMo5.1

and SiMo1000, as well as cast steel 1.4509, is greater than the austenitic materials up to ~600 °C. However, above

600 °C, the ferritic materials exhibit similar thermal conductivities compared to alloys A3N, CF8C-Plus, HK-Nb,

and HF. In contrast, Ni-resist D5S and HK 30 exhibit the lowest thermal conductivity. The authors [85] suggested

the relatively low thermal conductivity of HK 30 is due to two factors 1) a high alloying content relative to the other

austenitic steels (A3N, CF8C-Plus, HK-Nb, and HF) which results in increased phonon scattering and 2) reduced

carbide precipitation compared to the modified variant (HK-Nb), which results in more C and alloying elements in

solution and less free volume for thermal conduction. Similarly, increased phonon scattering in Ni-resist D5S due to

the relatively high alloying content may also be contributing to the poor thermal conductivity of this material.

Therefore, from an exhaust manifold alloy design perspective, precipitation strengthening of austenitic steels can

have multiple positive effects, such as enhancing thermal conductivity (by virtue of removing alloying elements

from solution) while enhancing LCF strength at high temperatures, which will be discussed in more detail in section

4.3.3.

As the temperature of the exhaust manifold increases during start-up, the material undergoes thermal

expansion. Points of constraint, such as bolted connections, resist this expansion and mechanical compressive loads

can develop resulting in elastic and potentially plastic deformation [85]. The thermal expansion coefficient of the

55
material directly influences the magnitude of the loads that develop during constrained thermal expansion. In

general, the ferritic materials exhibit lower thermal expansion coefficients than the austenitic materials, suggesting

that greater loads may develop in the austenitic materials during operation. Interestingly however, both Ni-resist

D5S and HK30 exhibit remarkably low thermal expansion coefficients at temperatures below 50 °C, which is

reported due to the Invar effect, which diminishes with increasing temperature and disappears above the Curie

temperature [93].

a) 30 b) 2.00

1.75
Thermal Conductivity (W m-1 K-1)

Thermal Expansion (, K-1 x 10-5)


25

1.50
20

1.25

15 SiMo1000 HF
1.4509 CF8C plus
SiMo5.1 1.00 A3N
A3N HK-Nb
CF8C plus HK30
10 HK-Nb Ni-resist D5S
0.75
HF SiMo1000
HK30 SiMo5.1
Ni-resist D5S 1.4509
5 0.50
0 200 400 600 800 1000 0 200 400 600 800 1000
Temperature (°C) Temperature (°C)

Figure 32 – (a) Thermal conductivity and (b) thermal expansion as a function of temperature for a series of candidate manifold
materials [85,93,99]. In (b), the thermal expansion data obtained from [85] corresponds to the mean thermal expansion over a
range of +/- 25 °C from a specific temperature.

4.3.2 Static Strength

The temperature dependence of the yield strengths of some of the alloys listed in Table 4 are shown in

Figure 33. The SiMo5.1 alloy exhibits greater yield strength below ~500 °C in comparison to the other alloys.

However, with increasing temperature above ~250 °C, the yield strength of SiMo5.1 decreases dramatically up to

800 °C. At 800 °C the yield strength is only 20 MPa which is similar to alloy 1.4509, another BCC alloy, but

significantly less than the austenitic alloys. The low strength at these high temperatures is a one reason why SiMo5.1

is generally limited to HDDE exhaust system applications where the temperature of the exhaust manifold will not

exceed 750-760 °C [85]. Ekström and Jonsson [85] noted that alloy 1.4509 exhibited poor mechanical properties in

general due to excessively large grain size. The authors attributed the large grain size to the direct liquid to ferrite

transformation which avoids the grain size reduction of ferrite that can occur during the austenite to ferrite

56
transformation other materials undergo. The austenitic alloys exhibit less temperature dependence than the BCC

alloys and superior yield strength above 500 °C. CN-12, with relatively high levels of C+N, exhibits the highest

strength above 600 °C. The yield strength of austenitic steels A3N and HK-Nb are superior to that of alloys HF and

HK30 over all temperatures, in part due to the precipitation of NbC carbides. Interestingly, the yield strength of

CF8C-Plus remained constant at ~114 MPa from 700-900 °C. This behavior is unique among the tested alloys. A

lack of reduction in yield strength in this temperature range is thought to be related to the highly stable carbonitirides

[96]. The thermal stability of these carbonitrides is often superior to that of carbides. An example of the size and

distribution of intragranular Nb carbides in CF8C-plus after creep rupture testing is shown in Figure 34, where both

cubic Nb(C,N) and the tetragonal Z-phase precipitates are observed, both of which are enriched with N and stable up

to temperatures of ~900 °C and 1350 °C, respectively. Coincidentally, the contribution of the dispersed carbonitrides

to the yield strength at 750 C was estimated, based on Orowan’s theory for the stress required to bypass an

impenetrable particle, and their size and distribution, to also be 114 MPa (the same as the yield strength at these

temperatures). This indicates that the Nb(C,N) and tetragonal Z-phase carbonitride precipitates are the predominant

contributor to the yield strength in this temperature regime and further reflects their high thermal stability.

Figure 33 – Evolution of yield strengths (YS) with increasing temperature for SiMo5.1 and candidate materials to replace
SiMo5.1 in HDDE exhaust manifolds and turbocharger housings. Data obtained from [85,90,100].

57
Figure 34 – TEM images of the microstructure of austenitic steel CF8C-Plus after creep-rupture testing for 684 h at 750 °C and
100 MPa: (a) cuboidal precipitates, with diameters of ~50 nm, are dispersed throughout the matrix; (b) A higher magnification
image showing a second precipitate phase, with diameters of ~16 nm; (c) SAD patterns recorded from the cluster of precipitates
encircled in (b) indicate that these are tetragonal Z-phase; (d) SAD pattern from the larger precipitate arrowed in (b) indicates this
is cubic Nb(C,N) phase. This data is reproduced from [96].

4.3.3 Fatigue Strength

Constrained thermal fatigue is a common cause of exhaust manifold failure and is an important design

criterion [86]. In 2014, Shyam et al. [86] compared the CTF performance of a series of cast irons (a SiMo grade, a

high SiMo grade, and Ni-resist D5S) and cast steels (CF8C-Plus, TMA 4705, and TMA 6301). The specimens were

rigidly constrained in displacement control (fixed grip) within a servo-hydraulic testing machine. Temperature

cycling was performed to induce stresses in the material. The maximum temperature of the cycle was 800 °C and the

minimum temperature was varied. Although the endpoints of the specimen are fixed, strains within the specimen do

still occur and are measured by extensometer to obtain the stress strain hysteresis loops. The temperature range (ΔT)

of the test along with the number of cycles to failure (Nf) is shown in Figure 35a. A power law relationship exists

between the magnitude of ΔT and Nf for each material. The materials can be ranked according to their performance

as follows (best to worst): CF8C-Plus > TMA6301 > TMA4705 >> D5S > High SiMo ~ SiMo cast iron. As

temperatures continue to increase in future HDD engines, the temperature range (ΔT) over which materials will be

exposed to will also increase, potentially enhancing the detrimental effects of CTF. Figure 35b shows the plot of the

58
inelastic strain, which includes contributions from both plastic and creep strain, as a function of Nf for each material.

The modified Coffin-Mason fits for each material and the common fit including all data in Figure 35b is shown and

provides a means to predict the fatigue life under CTF conditions from a single parameter, the inelastic strain.

Figure 35 – Constrained thermal fatigue data for cast irons (SiMo, high SiMo, and Ni-resist D5S) and cast steels (CF8C-Plus,
TMA 4705, and TMA 6301). (a) ΔT vs cycles to failure. The maximum temperature in all tests = 800 °C, the heating/cooling
rates are 1 °C s-1, and the hold time is 60 s. (b) Inelastic strain vs. fatigue life the data in (a). The Coffin Mason equation is fit to
all the data. This data in this figure reproduced from [86].

The fatigue life of alloys SiMo5.1, Ni-resist D5S, HF, A3N, HK30, and HK-Nb were also evaluated by

Ekström and Jonsson [85]. The authors performed strain-controlled fatigue tests at various different temperatures.

Figure 36a and b show the S-N curves obtained at 500 and 700 °C, respectively. The fatigue strength of SiMo5.1 is

superior to that of the austenitic steels at 500 °C but significantly less than that of the austenitic steels at 700 °C.

Figure 36c and d show the percentage of the total strain during LCF testing that is elastic and plastic as a function of

test temperature for SiMo5.1 and alloy A3N, respectively. The data in Figure 36c and d were recorded for a strain

amplitude of 0.25%. In Figure 36c, it can be observed that during the LCF testing, elastic strain dominates at 500 °C

but plastic strain dominates during testing at 700 °C for SiMo5.1. This transition from elastic to plastic strain

domination is consistent with a marked reduction in yield strength in SiMo5.1 with increasing temperature from 500

to 700 °C. In contrast, the proportion of elastic and plastic strains during fatigue testing is relatively stable as a

function of temperature up to 900 °C for the austenitic steels. An example of this is shown in Figure 36d for alloy

A3N. The test temperature of 900 °C is relevant in that it corresponds to a projected metal temperature of future

HDDE exhaust manifolds (see

59
Table 2). At 900 °C, the Nb precipitation strengthened austenitic steels A3N and HK-Nb, which are Nb

modified versions of HF and HK30, exhibit the best fatigue performance.

Figure 36– (a),(b) Stress amplitude vs. fatigue lifetime for alloys SiMo5.1, Ni-resist D5S, HF, A3N, HK30, and HK-Nb at
temperatures of (a) 500 °C and (b) 700 °C. (c),(d) The percentage of elastic and plastic strain of the total strain as a function of
temperature during LCF testing for (c) SiMo5.1 and (d) alloy A3N. The inset stress strain hysteresis loops in (c) are for the LCF
tests at 25 and 750 °C. The strain amplitude corresponding to the data in (c) and (d) is 0.25%. Fatigue tests were strain controlled
and fully reversed (R=-1) in air with a constant strain rate of 10-3 s-1. Stress amplitudes wear taken at half-life they were
cyclically stable. This figure is modified and reproduced from [85].

In the 2017 study by Norman et al. [101], the fatigue behavior of SiMo5.1 and SiMo1000 [93] were

investigated under two conditions: 1) TMF and 2) TMF combined with HCF. TMF represents the start, operate, and

shut down cycle of everyday engine use as components are heated and then cooled. The superimposed HCF is

intended to mimic engine cyclic structural loading due to combustion events every two revolutions per cylinder. In

the study, the temperature was varied from 300 to 750 °C and the HCF frequency was set at 15 Hz to be similar to

the rotational speed of the crank shaft. Figure 37a is a schematic of the employed thermal and mechanical loadings

employed in the study, where ΔεTMF and ΔεHCF represent the mechanical strain due TMF and HCF, respectively, and

60
Δεmech represents the sum of the two aforementioned quantities. Figure 37b is the hysteresis loop of the first cycle of

a TMF test (without HCF) for the SiMo5.1 SC (SC designates slower cooling during casting in a thicker mold),

SiMo5.1-RC (RC designates rapid cooling during casting in a thinner mold), and SiMo1000. During heating and

mechanical compression of the sample, the compressive stress increases to a maximum in all samples (which occurs

in the range from -0.1% to -0.15% strain). With additional compression and heating of the sample to -0.2% strain

and 700 °C, respectively, the materials undergo softening and or/creep, as evidenced by the reduction in

compressive stress. The softening is more pronounced in the SiMo5.1 grades than the SiMo1000, as indicated by the

black arrow in Figure 37b. The authors attributed the better resistance to softening of SiMo1000 to its better high

temperature strength [101]. As a result of its better resistance to softening at elevated temperatures, the maximum

stress (which is in tension) that develops in SiMo1000 during subsequent cooling to 300 °C and tensile strain to 0%

is lower for a given total strain range compared to SiMo5.1. In turn, SiMo1000 exhibits a larger number of cycles to

failure for a given strain range relative to SiMo5.1, as show in Figure 37c.

The authors also imposed a HCF strain simultaneously with a TMF strain [101]. The HCF was increased

with a commensurate decrease in TMF strain in order to keep the total mechanical strain constant. Figure 37d shows

the evolution of the maximum stress in the SiMo5.1 SC sample for a TMF test, a TMF test with applied HCF strain

of 0.08, and a TMF test with applied HCF strain of 0.12%. The total mechanical strain range was fixed at 0.25%.

Interestingly, increasing the HCF strain to 0.08 and again to 0.12% results in a substantial decrease in the maximum

stresses. The authors attributed these phenomena to a reduction in dynamic recovery that would otherwise occur

near 750 °C in the absence of the HCF load. As such, a slightly larger compressive stress occurs during elevated

temperatures but a lower maximum stress occurs upon cooling to 300 °C as the specimen is brought in to tension. In

the test with an applied HCF strain of 0.08%, the lower maximum stress results in a more than doubling of the

number of cycles to failure. In contrast, when the HCF strain is increase to 0.12%, the number of cycles to failure

drops to about half relative to the test with no HCF strain applied. Based on microstructural observations, the

authors reported that controlling mechanism responsible for the fatigue life in the TMF specimens are

environmentally assisted surface cracks. The cracks generally form from oxide intrusions which, over the course of

testing, grow to become macroscopically large cracks ultimately responsible for final failure [101]. However, when

the TMF was combined with HCF, the authors also reported the formation of microcracks emanating from graphite

nodules in the bulk that were not present in the samples where only TMF was applied. The authors concluded that

61
the formation and increased growth of these microcracks were primarily responsible for the significant decrease in

fatigue life when the HCF strain was increased from 0.08 to 0.12%. The damage mechanism associated with TMF

and TMF-HCF are discussed further in section 4.4.

Figure 37 – TMF testing results for SiMo5.1 slowly cooled (SC), SiMo5.1 rapidly cooled (RC), and SiMo1000 (a) Example of
the employed temperature cycle and strain due to TMF and HCF in the study by Norman et al. [101]. (b) Hysteresis loop of the
first cycle of a TMF test with no applied HCF strain. The black arrow indicates greater stress relaxation at elevated temperature
in the SiMo5.1 alloys compared to SiMo1000. (c) Cycles to failure as a function of mechanical strain range for TMF testing (no
HCF strain applied). (d) Maximum stress values vs. number of cycles for a test with a total mechanical strain range of 0.25% test
with no HCF and the application of HCF strains of 0.08% and 0.12%. This figure is modified and reproduced from the results of
Norman et al. [101].

4.3.4 Creep Strength

Creep can occur in manifolds and turbocharger housings due to stresses resulting from the component

weight and/or constrained thermal expansion acting over long periods of time at elevated temperatures [86]. Creep

rupture data for several ferritic (high SiMo and SiMo1000) and austenitic (Ni resist, CF8C, HK-Nb, and CF8C-Plus)

alloys that are used in HDDE exhaust manifolds or turbocharger housings or are being considered for these

applications are shown in Figure 38 ([90,93,99]). Among the alloys shown in Figure 38, the creep resistance from

greatest to least is as follows: CN-12 Plus=CF8C-Plus>HK-Nb>CF8C>Ni-resist>SiMio1000>High SiMo. It should

62
be noted that the ranking of these materials in terms of creep strength is also similar to the ranking of the materials

in terms of CTF performance, illustrating the importance of creep deformation during CTF. Composition,

microstructure, and crystal structure all influence the creep resistance of an alloy. Diffusion of atoms is one

mechanism which can enhance creep. Austenitic materials often have significantly lower rates of self-diffusion

(lower self-diffusion coefficients), compared to ferritic materials at the same temperature, and often have better

creep resistance [102]. This trend is reflected in Figure 38, with both the high SiMo and SiMo1000 cast irons

exhibiting lower creep rupture stresses for a given value of the Larson Miller Parameter (LMP) than the austenitic

materials. The creep resistance of the SiMo1000 is modestly improved relative to the high SiMo cast iron, related in

part to the superior high temperature strength of SiMo1000 [93], which is likely a result of solid solution hardening

by Al at the elevated temperature and other factors [93,103]. The CF8C-Plus and CN-12 Plus alloys exhibits the best

creep resistance of the alloys in Figure 38, likely due in part to unique precipitate structure. As discussed in sections

4.3 and 4.3.2 and shown in Figure 34, the microstructure of the CF8C-Plus alloy contains a dispersion of nano-sized

Nb-carbonitrides that have high thermal stability and resistance to coarsening, which significantly enhances the

creep strength. However, the austenitic materials also tend to have greater coefficients of thermal expansion and

reduced thermal conductivity. Therefore, under the same service conditions, austenitic exhaust manifolds could in

theory develop greater stresses due to constrained thermal expansion and/or operate at higher metal temperatures.

These factors are not accounted for in the laboratory creep tests.

63
Figure 38 – Plot of creep rupture stress vs. Larson Miller Parameter (LMP). Data were obtained from [90,93,99] and internal
ORNL data.

4.3.5 Oxidation Resistance

Resistance to oxidation is an important criterion for exhaust components in HDD engines. Not only can

long term oxidation at elevated temperatures cause significant metal section loss in some materials but spallation of

the oxide may also result in damage to rotating downstream components, such as high RPM turbine wheels [83].

Oxide intrusion at the surface has also been shown to facilitate the growth of surface cracks during TMF testing of

ductile cast irons SiMo5.1 and SiMo1000 [101], as detailed in section 4.4. These surface cracks were show to

govern fatigue life during TMF testing in the temperature range of 300-750 °C [101]. Ductile cast irons such as high

SiMo and SiMo5.1 used as exhaust manifolds and turbocharger housings are very near their limits in terms of

strength and oxidation resistance. Exhaust gas temperatures and the peak metal temperatures of exhaust system

components in HDDE for on highway use are projected to increase significantly in the coming decades (see

64
Table 2). These increased temperatures will exponentially increase the kinetics of oxidation. As a result,

there have been numerous research efforts to develop new materials with improved oxidation resistance and/or

evaluate the oxidation resistance of existing materials for their suitability in exhaust manifolds for future HDD

engines (e.g., [83,93,104,105]).

In a 2013 study, Ekström et al.[104] evaluated the influence of Cr and Ni additions on the oxidation

resistance of a base SiMo5.1 alloy. Two series of oxidation tests were performed, one in air and another in the

exhaust gases from a diesel test engine. Oxidation tests in air were performed in a furnace at 700 and 800 °C for

times from 6 to 300 h. Two different diesel engine oxidation tests were conducted. In both diesel engine tests,

screws of the material to be tested were mounted in the exhaust manifold, exposing one end of the screw to the

exhaust gases. In one test, the exhaust gases were held isothermally at temperatures between 700 and 740 °C and in

the other test the exhaust gas temperatures were cycled between 240 and 730 °C. The exhaust gases were estimated

to have a velocity of more than 100 m s-1. The composition of the exhaust gases was measured to contain primarily 4

mol.% O2, 4 mol% H20, 9 mol% CO2, and balance N2, although smaller amounts CO, NOx, SOx, and HC (unburned

fuel) were also present. The authors reported that the microstructure of the baseline SiMo5.1 consisted of graphite

nodules and M6C carbides (M=Fe, Mo, Si) distributed in the ferritic matrix. The addition of Cr resulted in increased

amounts of pearlite and carbides, the latter being M6C and a Cr-rich carbide which the authors speculated was the

M7C3 type. Ni did not have a significant influence on carbide or pearlite content and was primarily dissolved in the

ferrite matrix. The SiMo5.1 alloy with additions 0.5 wt.% Cr exhibited the best oxidation resistance in air at 800 °C,

showing roughly a factor of 4 less mass gain after 325 h compared to the same alloy but with 1 wt.% Ni added in

lieu of Cr, and was the only alloy that showed stabilized oxide growth during testing in air at 800 °C. In general, the

SiMo5.1 alloy and alloys with added Cr and/or Ni all formed complex multi-layer oxide structures after oxidation

testing in air at 700 and 800 °C. In the alloy with added Cr, XRD results indicated the presence of Cr2FeO4 among

the phases in the different oxide layers. Furthermore, based on thermodynamic calculations and SEM-EDS analysis,

the authors speculated that a relatively thin layer of SiO2 and Cr2O3 was forming at the oxide/metal interface which

resulted in the superior oxidation resistance of the alloy with 0.5 wt.% Cr. In contrast, the addition of 0.5 wt.% Cr to

the SiMo5.1 alloy resulted in no significant increase in oxidation resistance when the samples were tested in exhaust

gases. Furthermore, Cr was not detected at the oxide/metal interface or in the oxide scale itself. The authors

suggested that water vapor assisted evaporation of Cr may be occurring. Similarly, in a 2002 study by Asteman et

65
al.[106], the authors studied the effect of gas velocity on the oxidation of 310 steel. The gases used in the study were

dry O2, O2+10% H2O, and O2+40% H2O. The authors reported that a critical gas velocity exists, above which

significant water vapor assisted evaporation of Cr occurs, which interferes with the formation of the -(Cr,Fe)2O3

protective oxide layer. The critical gas velocity was lowered with increased water vapor content in the oxygen.

These studies by Ekström et al.[104] and Asteman et al.[106] illustrate how variables such as gas composition and

flow rate can significantly influence oxidation behavior.

Ekström [84] also evaluated oxidation resistance of alloys SiMo5.1, HF, A3N, HK30, HK-Nb, 1.4509, and

Ni-resist D5S both in air and 0.6%O2-1.2%H2O-Ar (bal.). Shown in Figure 39a and b are the mass gain curves of

the aforementioned alloys as a function of exposure time to air and 0.6%O2-1.2%H2O-Ar, respectively, at 900 °C,

except for SiMo5.1 which was held at 800 °C. In air and 0.6%O2-1.2%H2O-Ar, alloys HK30, 1,4509, and HK-Nb

all exhibit rather stable mass gain after 325 h of exposure. In contrast, alloys SiMo5.1, HF, and A3N have

considerably higher mass gain rates in air but even larger mass loss rates in 0.6%O2-1.2%H2O-Ar. The authors noted

that in the presence of water vapor, the lower Cr/Ni (20Cr/10Ni) alloys HF and A3N exhibited increased formation

of poorly protective Fe-oxide compared to alloys HK-30 and HK-Nb (~25Cr/20 Ni). The authors suggested this is

likely due to the evaporation of Cr due to the formation of CrO2(OH)2. Interestingly, alloy 1.4509 exhibited

excellent oxidation resistance despite a slightly lower bulk Cr content than HF and A3N. The authors suggested this

may be a result of less Cr tied up as carbides, due to the significantly lower C content of the alloy, and more Cr

available in the matrix to form a protective Cr-oxide layer. Additional oxidation mass gain data for SiMo5.1 during

isothermal holding at 700 [93] and 750 °C[101] are shown in Figure 39a to complement the data at 800 °C [84].

Interestingly, the oxide mass gain appears to become greater with decreasing temperature from 800 to 700 °C for

SiMo5.1, a trend that has been observed elsewhere and attributed to a continuous SiO2 layer which forms at the

oxide-metal interface at 800 °C but not 700 °C during the test. The oxidation mass gain for SiMo1000 in air at 700

[93] and 750 °C [101] is also shown in Figure 39a, and is significantly less than that for SiMo1000 at each

temperature. SiMo1000 also displayed slightly less mass gain relative to Ni-resist D5S during oxidation testing in air

at 700 °C for up to 300 h [93]. The mass gain of SiMo5.1 and SiMo1000 were also measured isothermally at 800

°C in synthetic exhaust gas with composition 5%O2-10%CO2-5%H2O-1ppmSO2-N2 (bal.) [105]. While a general

increase in mass gain was observed in SiMo5.1, SiMo1000 exhibited a relatively rapid mass decrease up to 160 h.

SEM EDS analysis indicated that significant decarburization occurred at the surface of SiMo1000 but not in

66
SiMo5.1, which the authors attributed to greater amounts of lamellar type graphite in SiMo1000, which is more

susceptible to decarburization than the spheroidal type.

Figure 39 – Mass gain as a function of exposure for isothermal oxidation tests in (a) air and (b) in synthetic exhaust gases of
composition 0.6%O2-1.2%H2O-Ar (bal.)[84] or 5%O2-10%CO2-5%H2O-1ppmSO2-N2 (bal.), as indicated. The data in these
figures are reproduced from [84,93,101,105].

Since most cyclic oxidation studies have focused on short term oxidation behavior (<500 h), Brady et al.

[83] evaluated the long-term oxidation behavior (>5000 h) of a series of candidate cast irons and cast stainless steels

for use in diesel exhaust systems using 100 h cycles. The long-term oxidation tests are more representative of

service lifetimes but do not include the more rapid thermal cycling (i.e. periodic cooling to room temperature) that

would occur during normal operation. Exhaust system hot lifetimes can exceed 15,000 h (assumes 1 million mile

engine life at 65 miles per hour average peak cruise speed) [98]. The materials studied were a SiMo cast iron similar

in composition to SiMo5.1, Ni resist D5S, CF8C-Plus, HK, TMA 4705, HP, TMA 6301, and CAFA 4. The latter

alloy is a cast alumina forming austenitic stainless steel. In many environments Al 2O3 oxide scales provide superior

protection relative to Cr 2O3 due to higher thermodynamic stability (particularly in the presence of H2O [107]) and

slower growth rates, which can be ~1-2 orders of magnitude slower than for Cr 2O3 [98]. The oxidation tests were

conducted at 650, 700, 750, and 800 °C by flowing air with 10% H2O across the face of the tested samples. Figure

40 shows the specific mass change as a function of time at 800 °C, recorded every 100 h, up to 5000 h. Peak

turbocharger housing temperatures are projected to increase to 800 °C by 2025, which is significantly greater than

the temperature limit of 760 °C specified for some SiMo cast irons in HDDE turbocharger housing applications. The

67
SiMo alloy exhibits specific mass gain of ~20 mg cm-2 after 3000 h at 500 °C, after which the rate of mass gain

significantly increases, with total mass gain of ~100 mg cm-2 after 5000 h. Microstructural characterization

suggested that the increase in mass gain after 3000 h may be attributed to the breakdown of the inner SiO2 rich oxide

layer that forms a protective layer at the oxide/alloy interface. The SiO2 rich oxide layer was confirmed to be present

at the oxide/alloy interface after 1000 h at 800 °C. However, the SEM image of the cross section of the SiMo cast

iron after 5,000 h at 800 °C in Figure 41 shows an oxide layer several hundreds of microns thick, significant base

alloy loss (original specimen thicknesses were over 1 mm), and remnants of the protective SiO2 rich layer in the mid

thickness of the oxide layer. While the oxide layer on the SiMo cast iron remained relatively adherent at 800 °C, the

large increase in thickness of the oxide scale can in some cases cause interference with the blades on a turbowheel.

In addition, these thick oxide layers, due to their lower thermal conductivity than the base metal [108], could act as

thermal barrier coatings and reduce the temperature of the base metal during operation as the oxide layer increases

in thickness over time. The tendency for or resistance to oxide spallation is a complex phenomenon with interplay

between numerous factors, including the thermal expansion coefficients, thermomechanical behavior of the oxide

phases and underlying substrate, and cycle frequency during testing [98,109,110]. The influence of cyclic frequency

on oxide spallation during oxidation testing of a series of Fe and Ni based alumina forming alloys was studied by

Pint et al. [110]. The authors reported that increasing the cycle frequency, from 100 h between cycles to 1 h between

cycles, can have significant positive or negative effects on oxide spallation depending on the class of material and

the significance of various competing factors.

The higher alloyed chromia-forming alloys with 25Cr and 20-35 wt.% Ni (HK, TMA 4705, TMA 6301,

and HP) all exhibited good oxidation behavior and only modest mass loss and limited Cr loss due to oxy-hydroxide

volatilization under the test conditions [83]. However, the CF8C-Plus alloy with lower Cr and Ni (19Cr/12N)

exhibited extensive oxide scale spallation (see Figure 40). Consistent with these results, an oxidation study of model

Fe-Cr-Ni austenitic steels with Cr contents in the range of 16 to 20 wt.% and Ni contents in the range of 15 to 30

wt.% also showed that higher levels of both Cr and Ni delayed the onset of accelerated attack and spallation during

oxidation testing at 800 °C in air with 10% water vapor. While the beneficial effects of increasing Cr on oxidation

resistance are well known, the mechanism(s) by which Ni additions improve oxidation resistance are less clear. The

authors suggested the most likely explanation as to why Ni additions improve the oxidation resistance is by

increasing the Cr activity and/or decreasing the Fe activity at the alloy/oxide interface or by lowering the oxygen

68
solubility/diffusivity in the alloy, which would have a similar effect as raising the Cr activity. However, it should be

noted that CF8C-Plus was deployed as a regeneration system housing material for ceramic diesel particulate filters

[90]. In this application where the material was exposed to rapid thermal cycling and temperatures that exceed

850°C, the CF8C-Plus alloy exhibited no oxidation issues, in contrast to the laboratory results [83]. Brady et al. [83]

suggested possible reasons for the discrepancy, including the potential for lower water vapor levels in the actual

diesel exhaust gases, which can range from 2-12% depending on numerous operational factors, as well as lower

effective metal time temperature profiles than suggested by the exhaust gas temperatures. An alternative explanation

is that the high temperatures occurred when the filter was being cleaned, which could have resulted in a much more

reducing environment than the laboratory tests conducted in wet air. This observation illustrates that accurate data

regarding component temperature and exhaust gas composition can be important to inform alloy development and

materials selection for these components. The development of Al-modified CF8C-Plus alloys is currently underway

to improve oxidation resistance. The alumina forming austenitic steel designated CAFA 4 clearly possessed the best

oxidation resistance under these test conditions, exhibiting a relatively small positive mass gain after 5000 h at

temperatures from 700 to 800 °C. A cross section of the CAFA 4 alloy after 5000 h at 800 °C in Figure 42 shows the

Al2O3 oxide layer with thickness less than 5 µm.

Figure 40 – Specific mass change as a function of time (100 h cycles) for the specified alloys at 800 °C in air with 10% H 2O on
two different scales (top and bottom). This figure is reproduced form [83].

69
Figure 41 – SEM back scatter electron image of the cross section of the SiMo cast iron after 5000 h in air + 10% H2O at 800 °C.
This figure is reproduced from [83].

Figure 42 - SEM back scatter electron image of the cross section of the CAFA 4 cast alumina forming austenitic stainless steel
after 5000 h in air + 10% H2O at 800 °C. This figure is reproduced from [83].

4.4 LIFE LIMITING MECHANISMS AT ELEVATED TEMPERATURES DURING

OPERATION

Two recent studies by the same research group (Norman et al. [69,101]) have investigated the damage

mechanisms in ductile cast irons under TMF and TMF+HCF loading. The temperature cycle was varied from 100 to

500 °C in one study [69] and from 300-750 °C in the second study [101]. In the former study (temperature range of

100 to 500 °C), the damage mechanisms during TMF loading of a gray cast iron, a CGI, and SiMo5.1 were

investigated. The initial microstructures of these materials prior to TMF testing are shown in Figure 23a, b, and c,

70
respectively. The graphite morphology is lamellar in the gray iron, vermicular in the CGI, and predominately

spheroidal in SiMo5.1. Shrinkage pores are also observed in the microstructure of SiMo5.1 in Figure 23c. The

microstructures of the specimens after TMF testing for approximately half of the lifetime of the specimens are

shown in Figure 24. In Figure 24a and b, representative cracks emanating from the tips of the lamellar and

vermicular graphite particles are shown and were present in the bulk and near the surface. In contrast however,

cracks were only observed to emanate from microshrinkage pores in the SiMo5.1 alloy and not the spheroidal

graphite particles. Therefore, the authors concluded that damage accumulation and failure in SiMo5.1 during TMF

in the temperature range from 100-500 °C occurs by microcrack initiation at microshrinkage pores throughout the

bulk material, crack propagation, and finally, coalescence and the formation of a dominant macroscopic crack which

ultimately leads to failure. Other studies by Nadot et al. [111,112] have also demonstrated that casting defects such

as microshrinkage pores act as crack initiation sites in ductile cast iron albeit during RT HCF testing. In addition,

Nadot et al. emphasized that both defect size and location are important to the fatigue life of ductile cast iron, with

near surface defects significantly more damaging than internal defects (even during testing at RT) as the

environment assists propagation of cracks emanating from surface defects and also due to the lower constraints

associated with surface defects leading to higher stress intensification.

In the TMF investigation by Noman et al.[101] where SiMo5.1 and SiMo1000 were cycled from 300-750

°C, environmentally assisted surface cracks are the primary damage mechanism leading to fracture. The cracks

formed from small oxide intrusions at the surface and grow to macroscopically large cracks that are ultimately

responsible for failure [101]. Figure 43a shows an oxide intrusion at the surface of SiMo5.1 after TMF testing for

250 cycles with a total mechanical strain range of 0.20%. In Figure 43b, under the same conditions but after 500

cycles, a large macroscopic crack over 1 mm in depth from the surface is shown. The authors noted that the growth

of the initial oxide intrusions and subsequent cracks are also facilitated by casting defects such as microshrinkage

pores. As such, the authors recommended that the oxidation resistance of the alloy be improved and that size and

density of casting defects be kept to a minimum in order to mitigate oxide intrusion and subsequent crack growth at

elevated temperatures. It should be noted that surface cracks which develop from oxide intrusions are expected to

become even more detrimental to the TMF life as exhaust gas and exhaust system component temperatures continue

to increase in future HDD engines (see

71
Table 2). Along these lines, internal TBCs that can be applied to cast iron are currently being evaluated as a

means to protect the surface from oxidation and also lower the maximum temperature at the inner surface of the

component, in some cases by up to 50 °C. A 50 °C reduction in temperature was shown by Ekström et al. [76] to in

most cases, have a significant and positive effect on the calculated fatigue life.

With the application of HCF loading on the TMF cycle (e.g., TMF-HCF), Noman et al.[101] observed a

significant reduction in the fatigue life when the HCF strain was increased to 0.12% (see Figure 37d) but observed

no significant difference in the oxide intrusion penetration depth compared to TMF testing in the absence of HCF.

The authors did however observe cracks emanating from spheroidal graphite particles which were not present after

TMF testing. Thus, the authors suggested that the additional loss of fatigue life with the application of HCF with

0.12% strain is a result of crack formation at internal graphite particles. Since TMF-HCF is intended to more closely

approximate operating conditions of exhaust manifolds and turbocharger housings, mitigation or suppression of

these internal cracks at graphite particles, potentially through strengthening the matrix, may be another pathway for

improving the high temperature fatigue life of ductile cast iron. Figure 43b and the discussion above indicates that

various life limiting mechanisms such as creep, fatigue, and oxidation can interact severely under the service

conditions relevant for manifolds and turbocharger housings.

Figure 43 – Cross section of SiMo5.1 after TMF testing with a temperature range from 300-750 °C after (a) 250 and (b) 500
cycles. Oxide intrusions like those shown in (a) can facilitate the development of macroscopically large cracks like that shown in
(b). A total mechanical strain range of 0.2% was applied. This figure is reproduced from [101].

72
4.5 PATHWAYS FOR FUTURE MATERIALS DEVELOPMENT

Exhaust manifolds and turbocharger housing for HDD 10-15 L engines have traditionally been

manufactured from ductile high SiMo or SiMo5.1 cast iron. These cast irons have been the preferred materials due

to their low cost (in terms of alloying and processing) and adequate elevated temperature properties such as static

strength, fatigue strength, and oxidation resistance. However, as exhaust gas and component temperatures have

continued to increase in recent years, high SiMo and SiMo5.1 grades are at their limits in terms of the required long

term durability under the present operating conditions. As shown in the previous discussion, higher maximum

temperatures can severely affect the fatigue life of the exhaust system components. In order to meet future HDD

efficiency goals, PCP, exhaust gas temperatures, and component temperatures are expected to increase substantially

in the coming decades. Therefore, in the near term, new materials technology must be implemented in exhaust

manifolds and turbocharger housings. To solve these challenges, research is currently focused on the following: 1)

enhancing the elevated temperature performance of ductile cast irons, 2) evaluating existing materials for exhaust

manifold and turbocharger housing applications, 3) the development of new materials for these applications, and 4)

development of coatings to enable lower cost/lower performance materials such as high SiMo and SiMo5.1 to

operate in higher temperature environments.

Ferritic materials are typically at a disadvantage in terms of elevated temperature mechanical properties

compared to austenitic steels. However, the improvement or development of new grades of ferritic cast iron or

ferritic cast steels are an attractive option for next generation HDD engine exhaust components due to the

significantly lower cost of these materials. These materials typically contain less than 1 wt. Ni% (see Table 4), a

costly alloying element, giving them a significant cost advantage over austenitic cast irons and steels that are also

being considered for these applications. Recent progress has been made in this area with the development of Al

modified ductile cast irons (e.g., SiMo1000 [93,113]) which can operate at higher temperatures and have improved

fatigue life and oxidation resistance at elevated temperatures compared to ductile iron grades high SiMo and

SiMo5.1. Nonetheless, while these gains are significant, they do not represent the substantial improvement in

elevated temperature properties that is realized by the use of austenitic steels. Continued metallurgical development

of ferritic ductile cast irons will likely be necessary for them to continue to be used in these applications, particularly

as exhaust gas temperatures are expected to increase significantly in the next decades. An alternative approach to

enabling ferritic cast irons to operate in higher temperature applications is internal TBC, which depending on

73
coating material and thickness, and may be able to reduce the temperature of cast iron by up to 50 °C at the cast

iron/TBC interface while preventing oxide intrusions into the base metal [76], which are known nucleation sites for

fatigue cracks. Thus, an internal TBC may be able to extend the elevated temperature fatigue life. However,

metallurgical development of the coating materials and the processes to apply these coatings require further

optimization to achieve the necessary coating durability. Slurry dip and plasma spray coatings are currently under

evaluation [76] and applying the coating to the sand cores with subsequent transfer of the coating to the part during

casting is another method that has been suggested [84]. A fundamental issue with TBCs is the metallic bond coating

to promote adhesion of the low thermal conductivity ceramic layer. Most conventional bond coatings contain Al to

form a thermally-grown alumina layer which has excellent adhesion to ceramics such as Yttria Stabilized Zirconia

(YSZ). For an Al-rich bond coating on ductile iron, the failure mechanism becomes the interdiffusion of Al back

into the iron substrate, which is very rapid above 700°C [114]. Furthermore, TBCs would need to be nearly 100%

free of any spallation that could result in damage to the turbocharger or other downstream components.

Austenitic cast iron (Ni-resist D5S) and a series of austenitic steels are being considered for exhaust

manifold and turbocharger housing materials in next generation HDDE. In some cases, new austenitic steels are

under development for these applications, such as CF8C-Plus [90]. The austenitic steels exhibit significantly

improved high temperature static strength, fatigue resistance, and oxidation resistance over ferritic cast irons and

steels, indicating they may provide enhanced durability at higher temperatures in exhaust manifold and turbocharger

housing applications in next generation HDDE. Nonetheless, in many cases, these materials represent a dramatic

increase in cost due to high Ni content compared to ductile cast irons. Furthermore, the austenitic steels with lower

levels of Cr and Ni and hence lower cost, such as the ~20Cr/10Ni HF, A3N, and CF8C-Plus grades discussed in this

review, exhibit inferior oxidation resistance. Future research should focus on the development of lower cost

austenitic steels that exhibit improved oxidation resistance. One potential avenue is the development of cast alumina

forming austenitic steels with relatively low levels of Ni [98].

New advanced manufacturing techniques such as additive manufacturing should be evaluated for exhaust

manifold production. Additive manufacturing could present numerous advantages compared casting, such as energy

savings and the flexibility to produce these materials without traditional foundries. In addition, additive

manufacturing could enable new microstructures due to vastly different solidification kinetics in comparison to

casting. Faster solidification kinetics could be important in the production of low cost ferritic steel manifolds, where

74
current casting practices can result in excessively large grain sizes in these materials. Furthermore, multi material

additive manufacturing could facilitate the production of manifolds with thin layers of heat and/or oxidation resistant

material on the internal surface.

Finally, the lifetime of exhaust manifolds and turbocharger housings are influenced by a multitude of

factors, including low cycle thermal mechanical fatigue, HCF, CTF, creep, and oxidation. In many cases, the

aforementioned phenomena act synergistically. As such, qualifying new materials for these applications requires

extensive and costly testing and even then, the tests may not accurately replicate actual service conditions. The

metallurgical development of these materials would benefit from improved testing and predictive capabilities that

more accurately take into account the complex mechanical and thermal loading, environmental conditions, and the

interaction between mechanical and environmental modes of damage during service conditions for these

components. Such tools could reduce materials development time and more closely align material properties with

those required based on the service conditions.

75
5 PISTONS

5.1 EVOLUTION OF MATERIALS AND MANUFACTURING

An overview of general piston design trends for HDD engines over the past three decades is shown in

Figure 44. In order to withstand increasing PCP and temperatures, the original aluminum piston designs were

modified to include cooling galleries, fiber reinforcement, pin bore bushings, as well as stronger aluminum alloys

(e.g., Al-12Si-1.0Cu-1.0Mg-1.0Ni wt.%) [115,116]. As PCPs and temperatures continued to increase, stronger

materials for the crown of the piston were required. Thus, articulated pistons, with forged steel crowns (often

manufactured from 4140H quench and tempered steel) and aluminum skirts, were introduced in the late 1980s in the

US and Europe. The main advantages of these pistons over predominately aluminum designs were their strength and

durability up to 1,000,000 miles. The disadvantages of the articulated piston design were increased weight, in part

due to a longer piston pin and a crown piston material that was thicker than necessary from a reliability standpoint,

due to the high cost of machining away material from the inner crown surfaces. One piece steel pistons were

eventually deployed and predominately replaced articulated pistons. A one piece steel piston developed by Mahle,

called the Monotherm® [115], was introduced around the year 2000. The one piece steel design was able to reduce

the weight in comparison to articulated piston designs, in part due to a shorter piston pin [115]. More recently, single

piece steel pistons with full skirts and redesigned cooling galleries have been deployed. Examples of these full skirt

pistons include the MonoWeld®, and MonoLite® (shown in Figure 45)[117], produced by Mahle, and the Magnum

MonosteelTM, produced by Federal Mogul. These pistons provide improved durability over previous one piece steel

designs and the re-designed oil cooling gallery reduces piston temperatures [117].

Figure 44 – Overview of piston types [115]. Reprinted with permission Copyright © 2017 SAE International. Further distribution
of this material is not permitted without prior permission from SAE.

76
Figure 45 – Image of MonoLite® piston. This figure is reproduced from [117].

During operation, pistons are subjected to complex stresses due primarily to temperature distributions

across the piston, gas forces, sliding frictional forces, and residual stresses due to previous manufacturing and

assembly [118]. Gas temperatures in the cylinder exhibit wide fluctuations as the engine goes through a single stroke

revolution. Conversely, due to thermal inertia, temperatures at specific locations on the piston are sometimes

assumed to be quasi-static during engine revolutions at constant engine operating conditions [118]. However,

significant temperature gradients are present on the piston. For instance, in modern steel pistons in HDDE,

temperatures may range from ~350-470 °C at the rim bowl, ~200-260 °C at the top ring grove, and ~150-235 °C at

the pin boss [117][118], depending on piston design and engine operating conditions. Significant thermal expansion

in the areas of higher temperatures induces primarily compressive, but also tensile stresses, depending on local

geometry. The loading due to temperature gradients is superimposed on the cyclic loading due to the gas forces. The

gas force is transmitted from the combustion chamber, to the piston crown and contact surfaces, through the skirt,

and to the piston pin. A detailed description of stresses resulting from these loads is contained in Reference [118].

Medium C low alloy ultra-high strength steel 4140H and micro alloyed grade 38MnSiVS5 (or similar

alloys) are the predominate materials for pistons in HD on road diesel engines. The compositions of these steels are

listed in

77
Table 5. Alloy 4140H is used in the quenched and tempered condition and the microstructure consists of

tempered lath martensite, as shown in the optical micrograph in Figure 46. For applications in pistons, the tempering

temperature (near 620 °C [24]) is typically much greater than the normal operating temperature so that further

structural changes of the bulk material are kinetically unfavorable during the operating lifetime. The Cr in 4140H

promotes hardenability and corrosion resistance, Mo promotes elevated temperature strength and hardenability

[119], Si increases oxidation resistance and elevated temperature strength [119–123]. 4140H is not a deep hardening

steel and variations in strength can occur for thicker sections [118]. The strengthening mechanisms of quenched and

tempered martensitic steel and microstructural development during tempering have been well reported in the

literature (e.g., [124,125]).

Figure 46 – Microstructure of quench and tempered 4140H (left) and 38MnSiVS5 (right) after etching with 2% Nital [126]. In
the 38MnSiVS5 steel microstructure, the ferrite (bright regions) outline the pearlite colonies. Reprinted with permission
Copyright © 2017 SAE International. Further distribution of this material is not permitted without prior permission from SAE.

The 38MnSiVS5 or similar grades of microalloyed steels (MAS) were invented several decades ago in

Germany [127]. MAS are gaining popularity as cost effective replacements for quench and tempered steels in some

applications [128–130]. With the application of controlled cooling after forging to confer suitable mechanical

properties, micro alloyed steels typically do not need subsequent tempering, and as a result, may have a lower initial

cost compared to 4140H quenched and tempered steels. MAS typically have pearlitic microstructures [128] and

more recently, micro alloyed steels with high strength bainitic microstructures have been produced [131]. The

microstructure of 38MnSiVS5 consists primarily of ferrite and pearlite, as shown in Figure 46, with ferrite outlining

the pearlite colonies. The strengthening mechanisms of micro alloyed steel include solid solution strengthening and

78
grain size of the ferrite, pearlite lamellar spacing, and precipitation strengthening and grain size control through the

addition of microalloying elements such as V, Ti, and Nb [126]. V is predominately in solution in austenite at

temperatures above 900 °C although some V carbonitirides, V(C,N), may be soluble depending on the composition

of the alloy [129]. During controlled cooling nucleation and growth of ferrite occurs at the austenite grain

boundaries. Subsequently, planar interphase precipitation of V carbides, nitrides, and/or carbonitrides occurs at the

moving austenite/ferrite interface [129]. These precipitates are typically below 10 nm in size [132] and their

distribution in the pro-eutectoid ferrite often takes the form of curved or straight sheets [129,133]. An example of

the distribution of V precipitates in the ferrite of a 0.1 wt.% C 0.1 wt.% V micro alloyed steel is shown in Figure 47.

Alternatively, a three dimensional atom probe elemental map with a 1.2 at.% V iso surface in Figure 48 shows the

distribution of V carbides in the ferrite of a 0.83C-0.61Si-1.56Mn-0.157V-0.039N at.% steel. V carbide precipitation

in pearlitic ferrite has also been observed in medium C steels [128]. These fine precipitates in pro-eutectoid ferrite

and pearlitic ferrite serve as obstacles to dislocation motion [134]. Various modifications to the base 38MnSiVS5

alloy have been researched to improve the microstructure and mechanical properties, such as HCF strength which is

of interest for piston applications. In MAS, the ferrite phase is generally softer than the pearlite and the deformation

during high cycle fatigue is predominately confined to the ferrite, which can result in crack nucleation at the ferrite-

pearlite interface [132]. Hui et al. [132] explored increasing the V content of the base 38MnSiVS5 alloy from 0.12

to 0.45 wt.% as a means to increase the strength and slip band initiation stress of the ferrite phase, thereby delaying

crack initiation and improving fatigue properties. The authors reported substantial increases in the RT HCF strength

with increasing V content, which they primarily attributed to greater V(C,N) precipitation as well as finer

microstructure. V results in a finer microstructure by a reduction in austenite grain growth, which is caused by a

solute drag effect due to V in solution in austenite or potentially V(C,N) precipitation. V(C,N) precipitates may be

soluble in austenite at typical forging temperatures depending on the bulk V, C, and N content [130]. Additions of

Ti are sometimes added to 38MnSiVS5 to supress austenite grain growth during heat treatments and forging by

precipitation of highly stable Ti nitrides and/or carbides (TiN/TiC) [127]. However, these particles possess high

hardness and can be detrimental to machinability [134].

79
Figure 47 –TEM micrograph of interphase precipitation of V(C,N) precipitates in irregular spaced and curved sheets in a 0.83C-
0.61Si-1.56Mn-0.157V-0.039N at.% steel [129].

Figure 48 (a) Three dimensional atom probe elemental map showing ferrite with sheets of V carbide precipitates with a 1.2 at.%
V isoconcentration surface in a 0.83C-0.61Si-1.56Mn-0.157V-0.039N at.% microalloyed steel. (b) top view of one interphase
plane and (c) proximity histograms of V, C, Mn, Si, and N. This figure is reproduced from [133].

A comprehensive comparison of the physical and mechanical properties of 4140H and 38MnSiVS5,

including engine testing, was conducted by Chen and Worden of Cummins, Inc. [126]. The reported yield strengths,

UTS, and fatigue strengths (corresponding to 107 cycles during rotating bending fatigue) from RT to 550 °C

reported in the aforementioned study are shown in Figure 49a, b, and c, respectively. The quenched and tempered

4140H steel exhibited superior yield strength, UTS, and fatigue strength over the investigated temperature range. In

both materials, the yield strength and UTS begin to decrease significantly with increasing temperature above 300 °C

and, above 500 °C, the drop off in yield strength is more severe in the 4140H steel. HCF strength is an important

criteria for piston materials and the HCF strength of the 4140H remains constant from 300 to 550 °C, whereas the

HCF strength of the MAS steel exhibits a significant reduction. It should be noted however that microstructural

80
variables can change the HCF performance in the two alloys. For instance, Hui et al. [135] studied the influence of

thermal processing (soaking temperature, forging parameters, and cooling rates) on microstructural refinement and

RT HCF behavior of MAS 38MnVS steel of similar composition to that reported in

81
Table 5. The authors reported that the finest prior austenite grain size of ~25 m, which was obtained under

specific forging and cooling parameters, resulted in the finest pro-eutectoid ferrite and pearlite constituents and best

HCF performance which was even superior to the HCF performance of the 4140H Q&T steel evaluated in the same

study for comparison. The prior austenite grain size achieved for the 38MnVS steel in the study by Hui et al. [135]

et al. of ~25 m is substantially smaller than prior austenite grain size of the MAS steel studied by Chen and

Worden (Figure 46), suggesting enhancements in the HCF strengths of current MAS pistons steels are possible.

82
Figure 49 – Yield strength, elongation, and fatigue strength (corresponding to 107 cycles) as a function of temperature [126].
Reprinted with permission Copyright © 2017 SAE International. Further distribution of this material is not permitted without
prior permission from SAE.

5.2 LIFE LIMITING MECHANISMS AT ELEVATED TEMPERATURES DURING

OPERATION

Steady state temperatures in the range of 470-520 °C have been measured near the surface of the bowl of

pistons in HDDE during operation [118][136]. At the aforementioned temperatures or greater, microstructural

changes may occur in 38MnSiVS5 and 4140H pistons in the bulk and/or at the surface over time during operation.

Depending on the severity of the microstructural changes, material degradation and component failure may occur. In

addition, oxidation, scaling, decarburization, and sulfidation may degrade the surface of the material, which can

facilitate cracking.

José and José, in a 1993 study, evaluated the metallurgical stability of articulated pistons with 4140H

crown material during operation in a dynamometer test engine [136]. The 4140H was tempered at 620 °C for 1 h

which is roughly equivalent to a tempering treatment of 520 °C for 1,000 h. The rationale was that tempering at 620

°C would make any further tempering at the operating temperature of 520 °C kinetically unfavorable. The test was

performed on a six cylinder, medium bore, high speed diesel engine. The specific power and the brake mean

effective pressure were 0.41 KW cm-2 and 1.71 MPa, respectively. The flowrate on special cooling oil nozzles was

adjusted to 1.0 l kWh -1 for odd numbered cylinders and 4.5 l kWh -1 for even numbered cylinders, in order to vary the

piston temperatures during operation. For the low and high flowrate conditions, piston temperatures (averaged over

the cycle) near the crown surface ranged between 476 and 517 °C and 436 and 465 °C, respectively, depending on

the measurement location on each piston. Piston temperatures were measured by sensors placed in locations just

underneath the piston crown surface. The authors observed the growth of an oxide layer on both the higher and

lower temperature pistons during testing. A sharp increase in the oxide layer thickness was observed in the first 100

h of operation, followed by a linear increase from 100 to 1500 h, as show in Figure 50. In this linear region, the

oxide layer thickness increased by 0.026 m h-1 at the surface of the higher temperature (514 ±17 °C) piston and

0.011 m h-1 in the lower temperature (457 ±16 °C) piston. However, the authors reported no substantial

decarburization (determined by chemical analyses), reduction in hardness, overtempering, or graphitization in the

base metal during engine testing up to 1500 h.

83
Figure 50 – Oxidation layer growth on the piston crown during engine testing [136]. For the lower temperature piston
configuration, the temperature at the thrust sides reached approximately 457 ±16 °C. For the higher temperature piston
configurations, the temperature at the front sides reached approximately 514 ±17 °C. This data is reproduced from [136].

Chen and Worden [126], in a study published in 2000, performed a full evaluation of the properties relevant

to pistons for the 38MnSiVS5 and 4140H materials. The results of specimen mass gain due to surface oxide during

oxidation testing in ambient air reported in that study are shown in Figure 51. In general, the 38MnSiVS5 performed

better during static oxidation testing, exhibiting less mass gain compared to the 4140H. Oxide spallation, also

measured in the study, was significantly less for the 38MnSiVS5 steel, suggesting the oxide is significantly more

adherent to this steel. The authors noted that a critical temperature of 593 °C exists, above which oxide formation is

significantly accelerated during laboratory testing in both alloys. In the same study, engine abuse testing was

performed to evaluate the durability of 38MnSiVS5 and 4140H. Engine tests were performed with a split set of

38MnSiVS5 and 4140H pistons. During the testing, cracks in the piston rim, like the one in Figure 52, developed in

some of the 38MnSiVS5 pistons. As is common in piston failures, these cracks were typically aligned along the

direction of the pin [137]. Post testing metallography by the authors suggested that temperatures at the crown

surface may have been much higher than the 500 °C limit established for the 38MnSiVS5. In comparison, none of

the 4140H pistons exhibited cracking. Examination of the 38MnSiVS5 steel pistons after engine abuse testing

revealed heavy oxidation on the rim, spalling, evidence of oxide intrusion into the substrate, and spheroidization of

the cementite. The microstructural changes of the 38MnSiVS5 near the crown surface caused the material to soften,

significantly reducing the hardness at specific locations on the piston rim, particularly near the spray plume at the

84
opposite side of the cooling nozzle [126]. Chen and Worden noted that the oxidation resistance of the 4140H was

superior to that of the 38MnSiVS5 during engine testing, contradicting the results of their static oxidation tests in air

(Figure 51). However, these results are not uncommon, and HDDE exhaust gases contain significant amounts of

water vapor, which can alter the kinetics and characteristics of oxidation in some materials, as discussed in more

detail in section 4.3.5. It should also be noted that actual determination of the metal temperatures of pistons during

operation, either by calculation or experimentally, is challenging and substantial uncertainty may exist between

actual metal temperatures and those that are measured or calculated.

Figure 51 – Mass gain as a function of temperature and time (100 and 500 h) during static oxidation tests of 4140H and
38MnSiVS5 specimens [126]. Reprinted with permission Copyright © 2017 SAE International. Further distribution of this
material is not permitted without prior permission from SAE.

Figure 52 – Photograph of 38MnSiVS5 steel piston bowl rim after engine abuse testing [126].. The arrow shows the location of a
crack aligned along the pin bore. Reprinted with permission Copyright © 2017 SAE International. Further distribution of this
material is not permitted without prior permission from SAE.

85
5.3 PATHWAYS FOR FUTURE MATERIALS DEVELOPMENT

Quenched and tempered 4140H and micro alloyed steel 38MnSiVS5 are the two primary materials

currently used in HDDE pistons. The elevated temperature HCF strength and oxidation resistance of 4140H and

micro alloyed steel 38MnSiVS5 limits their peak operating temperatures to ~500 and 450 °C, respectively [118] (

Table 1). Both materials have been subjected to relevant laboratory tests and engine abuse testing where the

components are tested in overloaded conditions, which have resulted in temperatures at the piston crown up to ~550

°C and beyond [126][136]. Microalloyed steel 38MnSiVS5 exhibited significant oxidation and microstructural

instability during engine testing [126], leading to failures of the pistons by cracking at the crown surface in one

study [126]. The pistons manufactured from 4140H also exhibited significant oxidation during engine abuse testing.

Metal temperatures near the bowl surface and rim are expected to be significantly greater than 500 °C in future HDD

engines [118]. Therefore, new materials or coatings may be required that exhibit superior properties above 500 °C

compared to 4140H and MAS 38MnSiVS5.

The following factors may be considered from a materials perspective in the design of next generation

pistons. These materials will require, at minimum, enhanced oxidation resistance and fatigue strength at elevated

temperatures in a combustion chamber environment relative to 4140H and 38MnSiVS5, particularly at the piston

crown and bowl rim. Alloy development strategies to increases HCF strength while maintaining good machinability

should be explored (e.g., enhancement of the HCF strength to UTS ratio[135]). In addition, substantial increases in

maximum piston temperatures would likely require alloys with improved microstructural stability. For Q&T steels,

this may require alloying strategies to promote temper resistance in order to prevent substantial microstructural

changes and loss of mechanical properties during service. Future advanced materials and manufacturing solutions

must also fit the tight cost constraints of the ground transportation industry – which can provide a substantial

scientific challenge – although the lower volume, higher cost HD market has slightly greater cost margins than the

automotive market. Both ferritic and austenitic steels are classes of materials that should be evaluated as candidates

for next generation piston materials. Ferritic steels have been used at temperatures up to 650 °C in pressure vessel

applications [119] and up to 550 °C as dies in hot working applications (e.g., medium alloy air hardening steels H11

and H13 [138]). With proper alloy design and processing strategies tailored towards next generation piston

86
applications, optimized ferritic steels may be suitable. The cost of the bulk material, processing steps such as forging

and subsequent tempering (in the case of quench and temper steels), and machinability are significant contributors to

the final cost and whether a material will be economical for use. Automotive materials trends are towards simpler

and more cost effective processing routes. The replacement of quench and tempered steel pistons with MAS pistons

has been limited in some markets, particularly due to durability concerns with MAS [126]. Pathways to enhance and

tailor the elevated temperature mechanical properties and oxidation resistance of MAS to next generation piston

applications, while maintaining processing simplicity and machinability, should be investigated.

Strategies for local reinforcement of areas with high thermal loads, such as the bowl rim, could be

alternative technologies to enable pistons to operate at higher temperatures [117,139,140]. These technologies may

include coatings and/or inserts [117] of materials of greater temperature resistance or surface modifications

techniques (e.g., laser shock peening [141]). In addition, additive manufacturing of pistons may enable more

complex piston geometries, such as optimized cooling galleries [142], that are difficult to produce using

conventional manufacturing and machining processes.

Finally, the design of the piston for thermal management must also be considered. HDDE pistons are

currently actively cooled by spraying oil into the cooling gallery where it enhances heat transfer from the crown

surface to lower portions of the piston. Heat transfer from the piston also occurs by conduction through the rings and

to the cylinder walls [118]. Oil spray cooling can reduce the maximum temperatures of the piston and enable the

implementation of less costly/less temperature resistant alloys. It may also result in reduced heating of the intake air,

due to a lower piston crown surface temperature, which can enhance volumetric efficiency. A negative consequence

of the oil spray cooling is greater removal of useful energy from the combustion gases, ultimately transferring it to

the cooling system where it is largely unusable [139], and also parasitic losses associated with the oil pumping. In

contrast, if the thermal energy is kept in the cylinder, it can be converted to piston work, used by the turbocharger,

and/or to raise the temperature of aftertreatment catalysts. The application of TBCs on the piston crown surface is a

potential avenue to increase thermodynamic efficiency of the engine. Numerous studies have indicated that low

thermal inertia coating materials of a specific thickness can improve engine efficiency [143–146]. These materials,

with low thermal conductivity and low volumetric heat capacity, allow the surface of the material to “swing” with

the combustion gas temperature, thereby reducing the temperature differential between the gas and metal surface

throughout the engine cycle. The benefit of these materials is to reduce heat transfer from the gas to the

87
piston/cooling system but also to reduce heat transfer from the piston surface to the intake air, thereby improving

volumetric efficiency. The coating material YSZ (thermal conductivity ~1.25 W m-1 K-1, volumetric heat capacity

~2500 kJ m-3K-1 at 500K [147]) has been identified as a potential candidate for this application in the past [143–

145]. More recently, Silica Reinforced Porous Anodized Aluminium (SiRPA) was developed and exhibits thermal

conductivity of ~0.67 W m-1 K-1 and volumetric heat capacity of ~1300 kJ m-3K-1 at 500K [146], both of which are

significantly lower than those values for YSZ. The SiRPA coating on an aluminium piston resulted in significant

efficiency improvements when tested in a research diesel engine. Significant efforts are also underway to improve

thermal management by developing unique piston designs. For instance, pistons with sealed cooling galleries which

contain liquids, low melting point metals, or some combination thereof are a potential avenue for more efficient

cooling of the piston (e.g., [148]). Advanced designs and materials that improve thermal management and efficiency

while maintaining low cost and sufficient durability for HDDE should be pursued for future pistons.

88
6 TURBINE WHEEL

6.1 DESIGN AND MANUFACTURING

Figure 27 shows a schematic of the gas flow on the hot and cold sides of a turbocharger. The turbine is

completely immersed in the hot exhaust gases, can reach ~760 °C in current production HDDEs, may rotate at

speeds on the order of ~100,000 RPM, is subjected to exhaust pulses, and may be impacted by foreign object debris

coming from upstream components. Temperatures at the turbine wheel are expected to reach 900 °C with continued

HDDE evolution (see Table 2). These considerations require the turbine wheel material to possess high resistance to

creep, oxidation, thermal fatigue, mechanical fatigue, and fracture. Turbine wheels in HDDE are typically

manufactured from Ni-based Inconel 713C or 713LC [149] by investment casting. Investment casting can produce

the complex shape of the turbine wheel but involves numerous production steps and high cost. The turbine wheel

may be used in the as-cast condition or subsequently Hot Isostatically Pressed (HIP). Additional heat treatments may

also be applied to improve mechanical properties [149].

Alternative materials and manufacturing approaches are also being considered for turbine wheels in HDDE.

Because of their low density (~4 g cm-3) and high specific strength, TiAl intermetallics are currently being

investigated as possible replacements for Inconel 713C. Lighter weight turbine wheels can improve engine

efficiency by reducing turbo lag. The turbocharger wheel inertia contributes to the delay in building manifold

pressure and engine torque in transient operation. In the current efforts toward “down-speeding” engines for higher

efficiency, it is advantageous to upshift the transmission as soon as there is sufficient engine torque to maintain the

vehicle acceleration or grade-climbing. The delay in available torque from turbo lag (on the order of seconds)

requires delaying the upshift to a higher gear and lower (more efficient) engine speed. More generally, less turbo

lag allows the engine to settle to a higher efficiency operating point more quickly. TiAl is discussed as a potential

candidate for replacement of Inconel 713C or Inconel 713LC in Section 6.3.

89
6.2 CURRENT TURBINE WHEEL MATERIAL

Inconel 713C was developed in the 1950s and provides good high temperature mechanical properties,

microstructural stability, and oxidation resistance [150]. A low carbon variant of Inconel 713C, Inconel 713LC, is

also used in turbine wheel applications. The composition of these alloys is listed in

90
Table 6. The microstructure of Inconel 713C is precipitation strengthened by  and carbides. Figure 53a

shows the microstructure of as-cast Inconel 713C containing NbC precipitates in the interdendritic regions, and

/NbC, / and /Ni7Zr2/M3B2 eutectics [150]. Figure 53b shows a higher magnification BSE SEM image of the

gamma prime phase with cubic morphology in the same as-cast sample [150]. The absence of Co reduces the cost of

Inconel 713C in comparison to some other materials used in turbocharger wheel applications (e.g. MAR-M246/247

or Inconel 100 [151,152]).

Figure 53 – Microstructure of as-cast Inconel 713C showing in (a) interdendritic region containing MC precipitates enriched in
Nb and /NbC and /Ni7Zr2/M3B2 eutectics and (b)  matrix containing  with cubic morphology [150].

The yield strength of Inconel 713C ranges from 690 to 760 MPa from RT to ~830 °C, after which further

increases in temperature result in significant decreases of the yield strength. Metal temperatures at the turbo wheel

are currently near 760 °C and are projected to increase to 900 °C in the future (

91
Table 2), which would result in an approximate 25% reduction in strength of Inconel 713C. Since the

weight of the turbine blades strongly influences the centrifugal forces and resulting stresses on the blades during

operation, the specific UTS of materials is used as one metric to compare candidate turbine wheel materials [153].

The specific UTS of Inconel 713C is shown in Figure 54 and compared with that of select TiAl alloys, which are

potential candidates to replace Inconel 713C and discussed in further detail below. The strength of the alloy at

elevated temperatures in the range of 700 to 900 °C is primarily due to solid solution strengthening, the precipitation

of NbC carbides and as well as the effect of the coherent  interface to inhibit dislocation motion. The

equilibrium amount of is generally in the range of 40 to 60 vol.% [154,155]. The specific strength of Inconel 100

(0.18C-10Cr-15Co-3Mo-4.7Ti-5.5Al-0.9V0.06Zr-0.014B wt.%), a material that has also been used in turbines is

shown in Figure 54 for comparison [152]. Inconel 100 contains higher amounts of forming lightweight elements

Al and Ti relative to Inconel 713C, contributing in part to significantly higher specific strength at temperatures from

760 to 900 °C.

Studies by Harada et al. [155,156] were performed to quantify the contribution to strengthening (tensile and

creep) of the  and  phases of a series of alloys which were developed to lay on the -tie-line of the pseudo

Inconel 713C binary phase diagram. This allowed the composition of the  and  phases to be kept approximately

constant in each alloy of different bulk composition while their amounts were varied from 0 to 100%, enabling the

deconvolution of the contribution of  (103 MPa) and  (~261 MPa) phases as well as the  interface (86 MPa) to

the yield strength (450 MPa) at a temperature of 900 °C. For the bulk composition corresponding to Inconel 713C,

the contributions of , , and the  interface was ~103, ~261, and ~86 MPa, respectively to the yield strength of

~450 MPa at a temperature of 900 °C. The contribution of the  interface to the yield strength was inferred from

the deviation of the measure yield strengths to values predicted by a mixture law. Creep rupture life and tensile

strengths generally increased substantially and peaked as the composition of Inconel 713C was modified to increase

the amounts of from ~53 to ~65 and ~75 vol. %, respectively. The study illustrates potential alloy design

pathways that could yield significantly improved elevated temperature mechanical properties while simultaneously

reducing the density of the alloy.

92
Figure 54 – Specific tensile strength as a function of temperature for Inconel 713C and two TiAl alloys [149,153].

Creep data for Inconel 713C in the as-cast, and heat treated conditions is compiled and reported in [149].

The alloy exhibits power law creep behavior during creep testing at 850 °C [157]. By solution treating then aging

Inconel 713C, it is possible to reduce the creep rate during testing at temperatures around 730 to 850 °C relative to

the as-cast material [149,158]. Azadi et al. [158] attributed the reduced creep rate in part to greater precipitation of

NbC in the interiors of grains in the aged sample relative to the as-cast sample. During creep testing at 850 °C [157],

the decomposition of NbC precipitates has been observed along with the precipitation of Cr enriched M23C6 and

M3B2 borides rich in Mo [150].

The fatigue properties of Inconel 713C and 713LC have been documented [149] and the high cycle fatigue

behavior of Inconel 713 LC has been studied in several recent publications [159–163]. The mechanisms relating to

fatigue damage and failure are discussed below in Section 6.4. However, it is clear that casting defects play an

important role in the fatigue life. Processing by HIP to reduce defect size can significantly improve the average

fatigue life [161]. Al diffusion coatings also impact the high cycle fatigue life of Inconel 713LC [162]. In the latter

study it was reported that Al diffusion coatings can cover casting defects at the surface, causing crack initiation to

occur subsurface where it becomes more difficult, resulting in improved fatigue lives in coated samples for a given

strain amplitude. Conversely, it was also noted that Al diffusion coatings cause structural changes to the base

93
material and shift the cyclic stress-strain curve to lower stress amplitudes, which results in shorter fatigue lives in the

coated samples for specific stress amplitudes.

The mass change due to oxidation in air and in air+10% H2O at 850 and 950 °C as a function of cumulative

time for 1 h cycles is shown in Figure 55a [164]. Inconel 713C exhibits low and positive mass gain at 850 and 950

°C, indicating the formation of a slow-growing, protective, and adherent oxide layer. The presence of water vapor

resulted in only a slight reduction in the mass gain, which is likely due to the increased stability of Al2O3 in the

presence of water vapor relative to oxide layers comprised of chromia or silica [164]. The oxidation resistance of

Inconel 713C and 713LC is expected to be sufficient up to projected future peak metal temperatures of 900 °C at the

turbine wheel.

Figure 55 – (a) Oxidation mass change for Inconel 713C and Ti-48.6Al-1Cr-2Nb at 850 °C in air and air+10% H2O and 950 °C
as a function of cumulative time for 1 h cycles [164] and (b) Oxidation mass change for Inconel 713C and a series of TiAl alloys
at 760 and 870 °C in air as a function of cumulative time for 2 h cycles [165].

6.3 CANDIDATE MATERIALS FOR FUTURE TURBINE WHEELS

Titanium-aluminide (TiAl) alloys are garnering significant interest as a potential replacement for Inconel

713C or 713LC as a turbine wheel material in HDDE. The composition of TiAl alloys are typically within the

following ranges Ti-(44-50)Al-(0-7)Nb-(0-3)Cr-(0-3)Ta-(0-1)Si-(0.016C-0.05)C at.% [166–169]. The TiAl alloys

94
under consideration for HDDE and automotive turbine wheels typically have lamellar microstructures consisting of

thin parallel plates of γ-TiAl (tetragonal L10 structure) and α2-Ti3Al (hexagonal D019 structure) [170]. This

microstructure is formed by reasonably quick cooling rates, such as those resulting from thin-walled castings, from

the α-phase (hexagonal) field [171]. During cooling, γ precipitates on (0001)α planes with (0001)α||{111}γ and at

lower temperatures the α-phase undergoes an ordering reaction to the α2 phase [172]. In addition to γ/α2 interfaces,

there is a random occurrence of six γ variants, which also results in different type of γ/γ interfaces [170]. A BF TEM

image of the lamellar microstructure is shown in Figure 56, showing lamellar thickness ranging from ~100 to 500

nm. The lamellar microstructure generally exhibits superior elevated temperature tensile strength and fatigue

properties near 800 C compared to “duplex” microstructures that can also be formed in TiAl alloys, typically by hot

working in the two phase α+γ regime [173]. In the lamellar microstructure, grain size, interlamellar spacing, and

orientation of the lamellae with respect to the tensile axis are important metallurgical parameters influencing the

mechanical properties, as illustrated by several studies. Hu et al. [174] reported pre-yield cracking in fully lamellar

Ti-44Al-8Nb, Ti-44Al-8Nb-1B, and Ti-46.5Al-2Cr-3Nb-0.2W-0.15B-0.4C at.% alloys by detecting acoustic events

originating in the material during straining below the 0.2% proof stress. The authors noted that these events are

associated with local deformation of the soft orientation γ lamellae and the stress at which these events occur

increases with decreasing colony size/lamellae length. The influence of lamellar spacing and orientation with respect

to loading on mechanical properties was studied by Umakoshi et al. [172] in a series of Ti-48.1-51.6Al at.% alloys

containing a single set of unidirectional lamellae grown using a floating zone method. The finest lamellar spacing

resulted in the best combination of ductility and strength and a Hall-Peth relationship between the RT yield stress

and the lamellar spacing was reported for lamellar perpendicular to the loading direction. More recently, alloy

design concepts such as solidification through the β-phase (rather than peritectically through the L+ β α), B

additions to increase the rate of heterogenous nucleation of α during the β to α phase transformation, and increasing

the stability of α grains against grain growth by Nb, Mo, B, and C additions have been explored [175].

95
Figure 56 – BF TEM image of the lamellar microstructure of a Ti-48Al-2Cr at.% alloy recorded along the <0-11> zone axis
[170].

In terms of mechanical properties for use in automotive turbine applications, TiAl alloys require both

elevated temperature strength, creep resistance, fatigue strength and sufficient room temperature ductility and

fracture resistance, among other properties. Alloy development research over several decades has improved these

properties in TiAl alloys through advancements in understanding the effects of alloying and microstructural

parameters on performance. In the late 1980s, General Electric patented TiAl alloys with nominal composition of

Ti-48Al-2Cr and Ti-48Al-2Cr-2Nb at.% alloy with RT tensile elongations of ~2-3%, which was considered

acceptable and a significant improvement relative to a binary Ti-48Al alloy [166]. The exact mechanism for the

increase in ductility by Cr and Nb additions is unclear. However, the addition of Nb appears to change the plasticity

mechanisms, resulting in widely dissociated super dislocations, large stacking faults, and mechanical twins in

deformed specimens, which is suspected to contribute to the enhanced ductility [170]. In 1996, General Electric

reported additional improvement in the elevated temperature creep resistance by addition of Ta, in a TiAl alloy of

nominal composition Ti-47.2Al-1.5Cr-0.8Nb-1.9Ta [167]. More recently, Mitsubishi Heavy Industries developed a

high Nb Ti-45Al-7.1Nb-1.0Cr alloy [153,169] with specific UTS, shown in Figure 54, superior to that of Inconel

713C up to at least 900 ºC. In the same report, the burst tip speed in turbine wheels manufactured from the high Nb

alloy were ~20% greater at 1000 ºC compared to Inconel 713C and ~10% greater than MAR-M247 [153].

96
Oxidation resistance in HDDE exhaust gases will be an important consideration for the use of TiAl alloys

in the turbine wheel. TiAl alloys generally exhibit poorer resistance to oxidation at high temperatures compared to

Inconel 713C. The threshold for the beginning of severe oxidation in TiAl alloys with increasing temperature

typically lies within the range of 750 to 850 ºC [176], with the specific temperature being dependent on composition,

microstructure, surface finish, and gas composition, among other factors. This temperature range includes the

current (~760 ºC) and predicted future peak metal temperatures (900 ºC) of the turbine wheel in HDDE (see Table

2). Similar thermodynamic stabilities of TiO2 and Al2O3 typically promote a non-protective mixed oxide scale of

TiO2 and Al2O3 in TiAl alloys [176]. In general, to improve the oxidation resistance of TiAl alloys by alloy design,

it is desirable to supress the growth of the fast-growing TiO2, encourage the growth of slow-growing Al2O3, while

improving the resistance of the scale to spallation. A review on the effects of alloy composition on oxide scale

growth rates was performed by Fergus [177]. Additions of Nb were identified as the most effective alloying addition

to improve the oxidation resistance of TiAl alloys. The mass change as a function of cumulative time for 2 h cycles

at 760 and 870 ºC for a series of TiAl alloys reported by Yoshihara and Kim [165], and reproduced here in Figure

55b, illustrate the beneficial effect of Nb additions. Common explanations for the improvement in oxidation

resistance by Nb additions are an increase in the Al diffusivity, a decrease in the oxygen solubility [177], and

improved resistance of the oxide scale to spallation as determined from isothermal and cyclic oxidation tests of

TiAl-X (X=Cr, V, Si, Mo, or Nb) alloys at 900 ºC [178]. Si, W, Mo, and large additions of Cr (greater than 4 at.%)

are also reported to beneficially reduce the rate of oxide growth in TiAl alloys at 900 ºC [177,178]. It should be

noted that under endurance testing in a gasoline engine, the high Nb alloy mentioned above, was reported to reach

temperatures of 900 ºC and greater, but the oxide layer thickness was ~ 1/5th that observed during the static

oxidation test at 850 ºC for 500 h in the same study [153]. Although the exact explanation for this behavior was not

known, the authors did report an adhesive layer enriched in Ca, Zn, P, Fe, and O, that was covering the oxide layer

and suggested that Ca, Zn, P (which might be present from lubricant additives), and or Fe might serve to stabilize

the Al2O3 layer. This observation again highlights the need for real world testing in conjunction with laboratory

testing for the qualification of HDDE materials. While the oxidation resistance of TiAl is inferior to that of Inconel

713C, the oxidation resistance of certain TiAl alloys may be suitable for next generation turbine wheels in HDDE.

In addition, some manufacturers are currently evaluating coatings for TiAl alloys to increase oxidation resistance at

higher temperatures.

97
6.4 LIFE LIMITING MECHANISMS AT ELEVATED TEMPERATURES DURING

OPERATION

6.4.1 Inconel 713C

Microstructural changes and damage due to heat exposure, continuous (creep) loading, fatigue loading, and

erosion or damage due to particle impingement are most likely to limit the useful life of Inconel 713C in turbine

wheel applications. Microstructural changes due to these phenomena will occur in Inconel 713C at temperatures and

loading conditions expected in next generation HDDE turbine wheels. Coarsening of the phase, which can have

negative consequences on strength, creep, and fatigue resistance, has been reported to be significant at temperatures

of 800, 875, and 950 °C by Ges at al.[154]. The coarsening rates of particles in Inconel 713C after the standard

heat treatment (1176 °C/2 h-AC + 926 °C/16 h-AC, where AC means air cooled) at the temperatures of 800 and 875

°C, which are relevant to next generation HDDE turbine wheels, are shown in Figure 57. The coarsening shown in

in Figure 57 over approximately 1000 h of testing would be significantly larger during the actual HDDE lifetime (>

15,000 h [83]). The coarsening rates follow LSW theory and t1/3 kinetics, where t is the exposure time. While the

mechanical properties of the material after long term aging were not measured in the study, it is likely that the

strength of the material would be degraded after such coarsening of the main high temperature strengthening phase.

It is also notable that the authors developed and performed an optimized 4 step heat treatment (1176 °C/2 h-AC +

1080 °C/2 h-AC + 925 °C/16 h-AC + 760 °C/16 h-AC) designed to promote high amounts of fine . The coarsening

rate of the sample initially exposed to the optimum heat treatment was significantly less at 850 °C compared to the

standard initial heat treatment, suggesting it may be possible to manage the coarsening of  expected in this alloy by

optimal initial heat treatments.

98
Figure 57 – Plot of size (a/2, where a is the cube edge length) in Inconel 713C vs time to the 1/3 power for two different initial
heat treatments. The standard heat treatment consisted of a solution treatment at 1176 °C/2 h-AC + 926 °C/16 h-AC. The
optimized heat treatment consisted of 1176 °C/2 h-AC + 1080 °C/2 h-AC + 925 °C/16 h-AC + 760 °C/16 h-AC. AC indicates air
cooling. Date replotted from [154].

Recent research to improve the understanding of the elevated temperature fatigue behavior of Inconel

713LC and Inconel 713C has been conducted by Kunz et al. [159–161], Obrtlík et al. [162], and Petrenec et al.

[163]. Kunz et al.[159] observed a high degree of scatter in the high cycle fatigue lifetimes of specimens tested at

800 °C. The authors observed two types of fatigue crack initiation and propagation, the first occurred and

propagated on slip planes and the second propagated in a non-crystallographic manner and typically perpendicular to

the loading direction. In some cases, cracks of either type clearly initiated at casting defects and in other cases,

casting defects were not observed near the crack initiation. The authors observed slip bands of relatively high

dislocation density by TEM in the tested specimens, which they suggested may occur in part due to the low stacking

fault energy of the alloy, which makes cross slip difficult. Such localized deformation during fatigue testing at 700

and 800 °C was also observed by Petrenec et al. [163]. Kunz et al. [159] suggested these slip bands may also be

crack initiation sites based on the TEM observations. The authors concluded that the high degree of scatter in the

fatigue lifetimes was likely a result of variability of microstructural conditions for slip band formation and crack

initiation as well as casting defect size distribution. A second report [160] from the same research group showed

qualitative agreement between the predicted maximum casting defect in three different lots of Inconel 713LC, based

on extreme value statistical methods, and the shortest fatigue life. In more recent work, Kunz et al. [161] showed

99
that high cycle fatigue life at 800 °C could be improved by HIP processing at 1160 °C and 100 MPa for 3 h.

However, despite the improvement, significant scatter in the high cycle fatigue lifetimes was still present in the

samples processed by HIP, supporting both the role of casting defects as well as other crystallographic phenomena

in contributing to the relatively large scatter in fatigue lifetimes of Inconel 713LC. It should also be noted that the

aforementioned fatigue tests are typically completed in a relatively rapid manner (e.g., ~24 h) and therefore do not

account for the effects of coarsening discussed in the previous paragraph, or oxidation, which would likely act

synergistically with fatigue damage to reduce the overall life of the material.

6.4.2 Titanium Aluminide

Erosion of the turbine wheel material was the most significant degradation mechanism encountered during

endurance testing of TiAl alloys by Tetsui and Ono [153,169]. The authors evaluated the endurance of as-cast turbo

wheels of two different TiAl alloys, Ti-47Al-0.4Nb-0.4Cr-0.6Si and Ti-45Al-7.1Nb-1Cr, in a 2.5L diesel engine

under various different test conditions. The maximum turbo inlet temperature during the test sequence was reported

to be 850 ºC and above. The material lost from the turbine blades was quantified by image analysis of cross sections

of the tested turbine blades. Approximately 7% of the cross-sectional area of the low-Nb turbine blades was lost

during testing while material loss in the high-Nb alloy was negligible. The material loss of the low-Nb alloy was

localized at the leading edge of the turbine blade. The authors concluded the material loss in the low-Nb alloy was

primarily caused by erosion due to contact of the turbine blades with particles such as detached rust and lubricating

oil, rather than spallation of oxides. In addition, the authors reported that the high-Nb alloy contained a fully

lamellar microstructure and finer average lamellar spacing, relative to the low-Nb alloy, which imparted greater high

temperature strength and fracture toughness, and reduced material loss.

A TEM study of the deformation mechanisms during long term creep tests of a duplex Ti-48Al-2Cr at.%

alloy at 700 ºC and 140 MPa was performed by Oehring et al. [179]. The conditions of this creep study may more

effectively replicate conditions at which HDDE turbines are exposed to than typical creep tests, which are often

conducted at higher temperatures and loads. The test lasted ~6000 h, consisting of a relatively rapid and large

amount of creep in the primary creep regime, followed by a long secondary creep regime, lasting ~4000 h, before a

transition to the tertiary creep regime. It was suspected that the high primary amount of creep was in part attributed

to the emission of interfacial dislocations at lamellar boundaries that were present at the start of creep testing. In

addition, the dissolution of α2 lamellae occurred during long term creep at 700 ºC. As temperature decreases below

100
~1100 ºC in TiAl alloys, the thermodynamic stability of the α2 phase decreases relative to the γ phase. This suggests

the α2 volume fraction did not reach the equilibrium amount at 700 ºC during initial cooling prior to creep testing. In

addition, ledges at γ/γ interfaces were observed to form during creep testing, indicating movement of the interfaces,

and in some cases, recrystallized γ grains were observed to form at the ledges, as shown in Figure 58. The

degradation of the lamellar microstructure during long term creep testing at times and temperatures relevant to

HDDE turbine operation suggest the mechanical properties could be significantly degraded and such tests should be

performed during qualification of TiAl materials in future HDDE.

Figure 58 – TEM image of a recrystalized grain that formed at a ledge between two γ lamellae during long term creep tests at
700 °C [179].

101
6.5 PATHWAYS FOR FUTURE MATERIALS DEVELOPMENT

Inconel 713C and Inconel 713LC are two materials currently used in turbine wheels of HDDE. Elevated

temperature strength, fatigue resistance, creep resistance, and microstructural stability are important criteria as the

metal temperature of this component is anticipated to increase from ~760 °C to ~900 °C in future HDDE. Erosion of

turbine blades may also become more severe in future HDDE, due to potentially greater spallation of metal from

upstream components as a result of higher operating temperatures and greater temperature swings in combination

with reduced strength of the turbine wheel material at higher temperatures. The oxidation characteristics, in terms of

mass change, were not observed to change significantly in lab tests from temperatures ranging from 760 to 950 °C

[164,165]. This suggests the oxidation resistance of Inconel 713C or 713LC will be sufficient to meet the anticipated

temperature increase in future HDDE turbine wheels. These alloys do however exhibit significant strength loss

above ~850 °C (see Figure 54) and as future HDDE turbine wheels approach peak metal temperatures of 900 °C,

additional qualification testing, turbine wheel design changes, material processing modifications, and/or material

substitution may be warranted. Strength reductions of this alloy may be mitigated to some extent through optimized

aging treatments [149,158], HIP processing [161], and coatings [162], which have shown to have beneficial effects

on the elevated temperature creep and fatigue properties in the projected temperature range of HDDE turbine

wheels.

Reducing the weight of the turbine wheel is desirable and could improve engine efficiency and

responsiveness. One approach to significant weight reduction of the turbine wheel is the use of TiAl alloys.

Significant improvements to TiAl durability, low temperature ductility and fracture resistance, as well as the

elevated temperature strength and oxidation resistance have been made over the last decades. Additional

investigation into the microstructural stability of TiAl alloys over long durations may be warranted to better

replicate the current and future HDDE operating environment. In general, while TiAl alloys appear to exhibit

properties that could enable them to operate successfully in future HDDE turbine wheel applications, the high cost

of producing TiAl turbine wheels is prohibitive. The cost of a TiAl turbine wheel is reported to be 1.5 to 2 times

greater than an Inconel 713C wheel [180] and such cost increases are generally not tolerable by industry without

extraordinary benefits. The raw material is typically only 10-20% of the cost of the final cast TiAl product [180].

Therefore, significant reduction of the manufacturing costs of TiAl turbine wheels should be a major focus of future

research. These efforts may include optimization of current investment casting processes [180], examining

102
alternative crucible materials/designs that can achieve improved yields [181,182] or the potential use of processes

such as additive manufacturing [183,184].

7 INTAKE AND EXHAUST VALVES

7.1 VALVE DESIGN AND MANUFACTURING

Internal combustion engine intake and exhaust poppet valves control the flow of the intake and exhaust

gases into and out of the combustion chamber. The valves seal the combustion chamber for much of the engine

cycle and are rapidly opened by various mechanical methods. A schematic of a typical ICE valve and associated

components in the cylinder head are shown in Figure 59 [185]. Figure 60 shows various poppet valve designs,

including one-piece, two-piece, tip-welded, internally-cooled, and welded seat face construction [186]. Exhaust

valve temperatures are typically much lower towards the tip of the valve [187], permitting the use of two-piece or

tip-welded valve designs. These designs typically employ martensitic steel in the tip or tip plus portion of the stem.

Tip welded designs are not very commonly used, most are a two-piece design with the stem weld near the hot end of

the guide contact area, unless a one-piece martensitic, in order to reduce stem wear, enhance fatigue strength in the

keeper grove, and reduce material costs [186]. Internally cooled valves include a hollow cavity in the stem typically

filled with sodium or sodium-potassium to enhance heat transfer, ultimately reducing the valve head temperature

and weight [188]. Internally cooled valves are typically not employed in HDDE as a result of their low RPM, which

reduces the effectiveness of the cooling mechanism (shaker effect).

Exhaust valve heads are primarily manufactured from low Ni content austenitic steels (less than 12 wt.%

Ni), nickel based superalloys containing more than 50 wt.% Ni [186], or alloys with intermediate Ni content

(generally 30 to 50 wt.%). Exhaust valve stems are manufactured from martensitic steels and friction welded to the

valve head while the entire intake valve is typically manufactured from martensitic steel (one-piece design) or

intermediate Ni content alloys (for the valve head, two-piece design). Valve materials undergo various processing

steps during production, among these are hot extrusion forging or upset forging to form the basic valve shape.

Exhaust valves may be solution annealed after forging (depending on the microstructural characteristics desired, like

grain size and hardness), followed by precipitation aging to enhance the strength of the material. These processes are

103
performed at high temperatures. For instance, a typical processing route for the Pyromet 31V alloy might include a

solutionizing treatment at 1120 °C for 1 h followed by double aging, first at 860 °C for 4 h then at 732 °C for 4 h.

The temperatures and times associated with these steps will vary depending on the material and desired properties.

Martensitic intake valves and exhaust valve stems require quenching and tempering after forging. Typically, the

majority of exhaust and intake valve stems are chrome plated or salt bath nitrided. These processes produce high

hardness at the stem surface in order to reduce scuffing between the valve stem and guide, as well as reduced

friction (in the case of Cr plating) [187]. Salt bath nitriding is typically not effective on Ni based superalloys,

however. For valve materials that typically exhibit lower hot hardness, such as austenitic steels, the seat of the valve

is often faced with a harder material to improve wear resistance (e.g., Stellite, Eatonite, etc.). The hard facing is

typically performed via plasma transferred arc (PTA) process [187].

Figure 59 - Typical features of diesel exhaust valve. This figure is reproduced from [185].

Figure 60 – Cross section of typical exhaust valve constructions [186]. Reprinted with permission Copyright © 2017 SAE
International. Further distribution of this material is not permitted without prior permission from SAE.

104
7.2 EXHAUST VALVE MATERIALS EVOLUTION

Exhaust valves in modern HDDE can be exposed to combustion pressures in excess of 20 MPa [3], reach

peak temperatures on the valve of at least 760-815 °C [187], high cyclic and contact stresses due to combustion and

valve seating, as well as corrosive combustion products [187]. Exhaust valve temperatures in modern diesel engines

have been estimated by comparing the hardness of fully hardened valve materials after engine operation for a

specified time to known temper curves for the same material [187,189,190]. These types of measurement methods

are subject to some uncertainty. Temperature check valve analysis lacks the ability to account for transient

temperatures at specific locations during the engine cycle and they are more representative of the average to

maximum temperature of the volume of the material that was involved in the hardness indentation. Valve

thermocouple instrumentation studies can be used to capture seat gradients, as well as maximum temperatures on the

valve head under various engine test conditions. These uncertainties notwithstanding, temperatures ranging from

~760-815 °C have been reported to occur at various locations on the valve head and stem [187,189,190] and even up

to 820 °C in some cases [187]. Maximum temperatures often occur at the center of the combustion face and also

along the section of the stem near the fillet [187]. Valve temperatures have historically increased as engine designs

have advanced, and are expected to continue to increase in future diesel engines (see

105
Table 2).

Exhaust valve material selection is typically driven by four main criteria: 1) high temperature fatigue

strength, 2) wear resistance, 3) corrosion/oxidation resistance, 4) manufacturability/compatibility with existing

facilities, and 5) cost (including raw materials and manufacturing). Numerous materials have found use in on-

highway HDD exhaust valve assemblies over the last few decades. A detailed review of the evolution of HDD valve

materials and design up to 1997 was performed by Schaefer et al. [187] and Larson et al.[188]. The purpose of the

current review is to investigate in greater detail some of the major historical metallurgical advancements that

occurred in exhaust valve alloys, as well as report on more recent developments in the last two decades and some

anticipated future trends.

The chemical compositions of typical valve alloys used in HDD engines are presented in Table 7 and the

historical time frame when specific exhaust valve materials were deployed in HDDE is shown in Figure 61. Many

valve materials in service today were developed over 4 decades ago. Shown in Figure 62 are the yield and UTS as a

function of temperature for common exhaust valve materials. These data will be referred to in the text where

appropriate for comparisons. The tensile data for Pyromet 31V, Inconel 751, and Nimonic 80A are unpublished

results on commercial alloys that were evaluated at ORNL. The Pyromet 31V, Inconel 751, and Nimonic 80A

specimens were solution annealed at temperatures of 1121 °C for times of 1 h then double aged, first at 857 °C for 4

h, then at 732 °C for 4 h. The yield strengths and tensile properties of the remaining materials are reproduced from

their corresponding references: 21-4N [186], modified 21-4N [186], 23-8N [186], VAT 32 and VAT 36 [191], and

NCF 3015A [192]. The heat treatment temperatures and durations as well as the test methods used to obtain the

tensile properties in Figure 62 varied. In many cases, the strain rate during testing, duration of the tensile test and the

amount of time the specimen is at temperature, along with prior thermal processing history, can have a significant

influence on mechanical properties. Therefore, Figure 62 should be interpreted in terms of general trends in the

different alloys as a function of temperature and comparisons of tensile properties between materials should be made

with caution.

106
Silchrome 1
8645
Chromo 193 (X 85 CrMoV 18 2)
Cr-Ni-Si-C Steel (Silchrome 10)
Cr-Ni-C Steel (21-12)
Cr-Mn-Ni-C-N Steel (21-4)
Nimonic 80A
Inconel 751
Pyromet 31V
Cr-Ni-Mn-C-N Steel (23-8N)
Cr-Mn-Ni-C-N Steel (Modified 21-4N)
NCF 3015/4015
Crutonite®
1940 1950 1960 1970 1980 1990 2000 2010
Year

Figure 61 – Materials and their approximate time frame of deployment in HDDE valves since 1940.

107
Figure 62 – Temperature dependence of (a) yield strength and (b) UTS for alloys used in HDDE exhaust valves. Data is
reproduced from [191–193] and internal ORNL data. The arrow in (a) indicates the general trend in yield strength requirement
for future HDDE operating at higher temperatures and pressures.

7.2.1 Precipitation Strengthened Austenitic Steels

Up to the late 1970s and early 1980s, HDD engines for on road highway applications commonly used

either 21-12N or Silchrome 10 austenitic steels for exhaust valve applications. Neither alloy was hardenable by

carbonitride precipitation, which promoted adhesive wear at the valve/seat contacts during combustion. In order to

reduce wear, the hardfacing material Stellite was applied via weld overlay to the valve seat face [187]. Stellite

superalloys are Co based with additions of Cr, W, and smaller amounts of C, Mn, Si, Ni, Fe, and sometimes, Mo

(see Table 7). These alloys exhibit excellent strength, wear/erosion resistance, and corrosion/oxidation resistance at

high temperatures [194]. Chromium (20-30 wt.%) in Stellite alloys provides oxidation/corrosion resistance, solid

solution hardening, and precipitates as carbides (predominately M7C3 in high C Stellites or M6C and M23C types in

low-carbon Stellites [195]). These carbides impart significant hardness to the Stellite microstructure. For instance, a

hardness value of 1000 HV at 650 °C was measured for the M7C3 type carbide itself (not the bulk microstructure)

[196].

In the early 1980s, Fe-based austenitic precipitation hardened alloy 23-8N replaced alloys 21-12 and

Silchrome 10 in most on-road HDD exhaust valve applications. Alloy 23-8N was developed jointly by Eaton

Corporation and Armco steel and provided improved strength, hardness, fatigue resistance and oxidation resistance

at 760 °C compared to 21-12N and Silchrome 10. Alloy 23-8N is a precipitation strengthened alloy based on 21-4N,

which in turn was based on the work of Jennings at Armco Steel Corporation. Jennings noted that N additions over

0.2 wt.%, along with sufficient C additions, led to significant increase in hardness at 760 °C in a 21Cr-9Mn-4Ni-

0.6C-0.1Si steel, in part due to the precipitation of carbonitrides [197,198].

108
The mechanisms by which N additions influence the microstructure and elevated temperature mechanical

properties of stainless steels have been investigated by numerous researchers (e.g.,[188,199,200]) subsequent to the

findings of Jennings. Mathew et al. [200] reported significant increases in the creep strength of solution treated

316LN stainless steels with increasing N content from 0.07 to 0.22 wt.%, which the authors attributed in part to the

precipitation of fine carbonitrides in the steels with 0.14 and 0.22 wt.% N. N additions have also been reported to

improve the thermodynamic stability of the microstructure, by delaying the precipitation of M 23C6 carbides

[201,202]. While the precipitation of the M23C6 carbide may impart hardness, the carbide is typically highly

enriched in Cr, which may lead to depletion of Cr in the volume surrounding the carbide, leading to areas of reduced

corrosion resistance. Other N-related strengthening mechanisms besides carbonitride precipitation have been

reported. Byrnes et al. [199] investigated the influence of N additions from 0.04 to 0.36 wt.% to a solution treated

austenitic stainless steel with nominal composition of 0.016C-26Cr-1.7Mn-32Ni-3.3Mo-0.45Si. The authors

reported a significant enhancement of the athermal component of the yield stress, as shown in Figure 63, which is

beneficial in the operating temperature range of exhaust valves. The authors suggested that short range ordering, as a

result of the strong affinity between Cr and N, and possibly interactions of N with Mn and/or Mo, is likely the main

factor contributing to the increase in the athermal component of the yield stress with increasing N content. Additions

of N are also thought to be beneficial to the long term thermal stability and fatigue resistance. Oda et al. [203]

investigated Cr-N substitutional-interstitial complexes (short range order) in 15Cr-15Ni austenitic steels with

additions of N using x-ray absorption near edge structure (XANES) and extended X-ray absorption fine structure

(EXAFS) techniques. In that study, the steels were solution treated and quenched in water to avoid precipitation and

then subjected to strain-controlled LCF testing. The authors reported that fatigue testing eventually resulted in a

breakdown of the Cr-N complexes and that this was accompanied by significant strain softening, suggesting that the

Cr-N complexes may contribute to fatigue resistance.

109
Figure 63 – Influence of temperature and N additions on the 0.2% offset yield strength of an austenitic stainless steel with
nominal base composition of 0.016C-26Cr-1.7Mn-32Ni-3.3Mo-0.45Si wt.%. Data reproduced from [199,204].

In the mid-1980s, a continued need for an Fe based alloy with superior performance drove the development

of modified 21-4N. Modified 21-4N possesses improved high temperature fatigue, strength, and wear properties

over 21-4N, owing to additions of 2.2 wt.% Nb+Ta and 1.2 wt.% W [187]. Modified 21-4N exhibits an approximate

10% improvement in UTS over alloy 21-4N (see Figure 62). In addition, Nb and Ta additions improve the creep

strength as well as intergranular corrosion (i.e. aqueous corrosion) in austenitic steels [205]. Nb improves creep

strength primarily due to the precipitation of fine intragranular M(C,N) carbonitrides [202] [206]. Additional creep

strength in modified 21-4N is obtained by alloying with W, which is known to slow the coarsening rate of M23C6,

although the mechanism is not fully understood [207]. Nb is a stronger carbide former than Cr and may suppress the

precipitation of M23C6 in some austenitic steels. The suppression of M23C6 can enhance the resistance to

intergranular corrosion. M23C6 often precipitates on grain boundaries and contains significant amounts of Cr

(reported to be ~ 65 at.% in austenitic grade 304 [205]) which can deplete Cr from areas surrounding the carbide,

leaving these areas more susceptible to corrosion [202]. In general, the size, morphology, and distribution of

carbides in the microstructure also significantly influence the wear behavior of 21-4N. Arockia Jaswin et al.

[208,209] reported that deep cryogenic treatments of 21-4N resulted in finer carbide particles and improved the wear

resistance at RT and tensile properties at elevated temperatures.

110
7.2.2 Nickel Based Superalloys

Ni based superalloy Pyromet 31V was developed in the late 1970s to address a need for greater sulfidation

resistance in a Ni based superalloy compared to Inconel 751, while maintaining similar high temperature strength

[187]. Due to the relatively high sulfur content in diesel fuel at the time, diesel engine valves were one of several

applications that drove the development of Pyromet 31V [210]. Pyromet 31V contains higher levels of Cr and lower

levels of Ni than Inconel 751, which contributes to better sulfidation resistance [210]. The hard  intermetallic

phase, which develops after a solution treatment and during aging (with aging typically performed between 650 and

815 °C [210]), is predominately responsible for the alloy’s excellent high temperature strength. Additions of Mo

(~2.0 wt.%) also provide solid solution strengthening of the austenitic matrix. The strength of the alloy begins to

decrease significantly as temperatures are increased beyond 650 °C however, as shown in Figure 62. The decrease in

strength is in large part due to a decrease in the  phase fraction with increasing temperature. The phase equilibria

calculated for this alloy, shown in Figure 64 [211], indicates a significant decrease in the volume fraction of the 

phase from 600 to 800 °C, followed by a dramatic decrease to zero with further increase in temperature up to ~975

°C [211].

Figure 64 – Phase equilibria for Pyromet 31, reproduced from [211]. Copyright 2005 by The Minerals, Metals & Materials
Society. Used with permission.

111
Pyromet 31V contains low amounts of C (0.04-0.065%) to enable desirable precipitation of limited

amounts of carbides on the grain boundaries while avoiding tying up significant amounts of Ti (in TiC carbide),

which is needed to form adequate amounts of the strengthening  phase. Evans et al.[211] investigated the

microstructure of solutionized and age-hardened Pyromet 31, as well as the same material in valves after being in

service in natural gas reciprocating engines for times of 2400 h. In the solutionized and aged condition, spherical 

precipitates with average diameters of ~48 nm were dispersed throughout the grains as shown in Figure 65. In

addition, M23C6 carbides, consistent with the predicted phases in Figure 64, precipitated on the grain boundaries as

shown in Figure 66. The higher Cr content can also result in the precipitation of the -Cr BCC lath phase, which

was also observed by Evans et al.[211]. However, it is unclear what, if any, deleterious effect the -Cr phase has on

the alloy. The coarsening behavior of this alloy will be discussed in section 7.4.4.

Figure 65 – TEM micrographs of Pyromet 31V in the solutionized and aged condition. (a) selected area diffraction pattern from
the gamma matrix showing  superlattice reflections and (b) centered dark-field image of  particles using the (100) 
reflection. The figure is reproduced from [211]. Copyright 2005 by The Minerals, Metals & Materials Society. Used with
permission.

Figure 66 – TEM images from solutionized and age hardened Pyromet 31V showing in (a) M 23C6 precipitates decorating a grain
boundary (identified by arrow) and (b) selected area diffraction pattern from the precipitate identified in (a) showing matrix and
M23C6 reflections. The figure is reproduced from [211]. Copyright 2005 by The Minerals, Metals & Materials Society. Used with
permission.

112
7.2.3 Intermediate Nickel Content Alloys

Since the 1980s, there has been significant effort to develop exhaust valve alloys with improved high

temperature properties, but at lower cost than existing Ni-based super-alloys. One strategy to achieve this goal has

been the development of alloys with Ni levels intermediate to that of 21-12N and Pyromet 31V [191,192,212–214].

Nickel is needed for high temperature strength and oxidation resistance. Ni, when combined with Ti, Al, and Nb

also forms stable  intermetallic precipitates, which greatly enhance the high temperature strength of alloys such as

Inconel 751, Pyromet 31V, and Nimonic 80A. Other strengthening mechanisms such as Nb and Mo carbide

precipitation and solid solution strengthening are employed. In these alloys, Fe is also present in large quantities for

cost reduction but too much can destabilize the FCC matrix.

In the mid 1990s, alloys NCF3015A (30 wt.% Ni) and NCF4015A (40 wt.% Ni) were developed in Japan

and deployed in exhaust valve applications. Alloys NCF3015A and 4015A included additions of Nb, 0.8 and 1.3

wt.%, respectively, as well as Mo, 0.7 wt.% each. Niobium promotes (Ti,Nb)C carbide rather than M23C6 and

stabilizes and strengthens the  phase, while Mo is added for solid solution strengthening of the  matrix. Alloy

NCF4015A displayed equivalent high temperature mechanical strength to Inconel 751 despite a significant reduction

in Ni content. Alloy NCF3015A exhibited a significantly better fatigue strength compared to 21-4N, but slightly

lower than that of Inconel 751.

Eaton Corporation and Crucible Materials Corporation jointly developed intermediate Ni content alloy

Crutonite® [212]. This alloy successfully bridged the gap between austenitic alloys (23-8N and mod 21-4N) and Ni

based super-alloys (Inconel 751, Pyromet 31V, and Nimonic 80A). In this alloy, strengthening is enhanced by the

precipitation of finely dispersed intermetallic  phase and (Nb,Ti)C primary carbides (1 to 4 vol.%) improve

adhesive and abrasion wear resistance. The wear resistance of Crutonite® is high enough that hard facing the valve

seat is not necessary in most applications. Crutonite® is also nitridable, enabling further improvements in the wear

resistance. However, in alloys with intermediate Ni content, oxidation resistance and fatigue strength can become a

concern where metal temperatures exceed ~815 °C.

In 2013, researchers from Vilares Metals S/A reported on two intermediate Ni content alloys for exhaust

valve applications, designated VAT 32 and VAT 36 [191,193,215,216]. The nominal compositions of these alloys

are shown in Table 7. VAT 32 contains significantly higher C content relative to Nimonic 80A, Inconel 751,

113
NCF3015A, NCF4015A and higher Nb content than all other valve materials listed in Table 7. VAT 32 is similar in

composition to Crutonite® but contains significantly higher Nb content. The alloying strategy behind the relatively

large additions of Nb and C was to promote significant precipitation of NbC, stabilize the  phase to higher

temperatures, and precipitate the Ni3Nb  phase for improved wear resistance. The measured amount of Nb-

containing carbides was reported to be ~2.5 vol.% in VAT 32 compared to less than 0.25 vol.% in VAT 36, Nimonic

80A, and Inconel 751. The higher amount of carbides in VAT 32 imparts substantially better hot hardness to this

alloy. For instance, the hardness of VAT 32 was ~300 HV at 875 °C compared to ~210 HV for Nimonic 80A, 250

HV for Inconel 751, and 170 HV for VAT 36, as show in Figure 67. The hot hardness of VAT 32 is even superior to

some hard facing materials at temperatures from 700 to 875 °C [191,217], which precludes the need to hard face the

valve seat in some applications, reducing manufacturing complexity and cost. Conversely, hot rotating bending

fatigue tests indicated the VAT 32 exhibited significantly lower fatigue strength than Nimonic 80A and VAT 36

from RT up to 860 °C. The authors suggested that the higher volume fraction of carbides in VAT 32 were

detrimental to fatigue strength, as they provide more crack nucleation sites. The increased wear resistance, coming at

the expense of reductions in fatigue strength, is an example of the challenges associated with balancing materials

properties during alloy design.

Figure 67 – Hot hardness of VAT 32, VAT 36, Nimonic 80A, and Inconel 751 [191]. VAT 32, VAT 36, and Nimonic 80a were
solution annealed and aged while Inconel 751 was solution annealed and double aged according the parameters in [191].
Reprinted with permission Copyright © 2017 SAE International. Further distribution of this material is not permitted without
prior permission from SAE.

Recent research has made further and significant advancements in the elevated temperature mechanical

properties of intermediate Ni (40-50 wt.%) alloys for next generation exhaust valve applications [193]. In the most

114
extreme case, a 78% increase in yield strength at 870 °C, as directly compared to Inconel 751, was achieved by an

alloy developed at ORNL designated 161 [193], as shown in Figure 62. In contrast to the previously discussed

alloys, alloy 161 employs low levels Co (1 wt.%) for solid solution strengthening. A major factor contributing to the

elevated temperature mechanical properties of these alloys is strategic alloying to significantly enhance the thermal

stability of the  strengthening phase. The phase equilibria for alloy 161, shown in Figure 68, indicates the  phase

is stable to significantly higher temperatures, over 1000 °C, compared to Pyromet 31V (see Figure 64) or alloy

Inconel 751. The elevated temperature mechanical properties of these alloys are also enhanced by alloying to 1)

provide solid solution strengthening of the  matrix, 2) precipitate (Ti,Nb,Ta)C carbides, 3) manipulate the 

antiphase boundary energy, the - lattice misfit, the coarsening rate, and the interdiffusion coefficient. This

significant enhancement in elevated temperature mechanical properties suggests these lower-cost alloys could

operate at higher temperatures and/or for longer lifetimes than most current valve materials.

Figure 68 - Phase equilibria for intermediate Ni content alloy designated 161 developed at ORNL. The figure is reproduced from
[193].

7.3 INTAKE VALVE MATERIALS EVOLUTION

Because intake valves are cooled by the incoming intake charge, their typical operating temperatures range

from 430 to 625 °C, which is significantly lower than exhaust valves. Intake valves for the 15 L engines have

historically been manufactured from low alloy martensitic steels such as Silchrome 1, higher alloyed martensitic

steels such as Chromo 193 (see Table 7) or austenitic steels with seat welds when increased wear and oxidation

resistance are required [186]. The martensitic steels typically exhibit excellent hot hardness up to 590 °C. However,

with the introduction of EGR in the 2000 time frame, sulfur based acidic corrosion [25] attack has been observed on

115
some martensitic intake valves in the United States, even in diesel fuel with less than 15 ppm sulfur levels. Acidic

corrosion resistance, with corrosion being a precursor to radial cracking and gutters on intake valves, and fatigue

strength requirements have been driving a switch towards austenitic and nickel base intake valves. For instance,

Crutonite® has been deployed in some 15 L engine intake valve application due to the need for greater acid

corrosion [25], oxidation, and seat wear resistance, as well as improved fatigue strength.

7.4 LIFE LIMITING MECHANISMS AT ELEVATED TEMPERATURES DURING

OPERATION

The mechanisms which contribute to valve degradation in service are numerous and complex. During long

term operation, valve seat wear, HCF, valve stem scuffing, oxidation, sulfidation corrosion, erosion corrosion

(guttering), material aging, and de-alloying may all act synergistically to degrade the valve and associated valve

materials. Valve seat wear is typically the most serious concern during long term operation and can result in chordal

and/or radial cracking. The valve seat wear mechanisms have been investigated by many researchers over the last

several decades, among them: Narasimhan et al., 1981 [217]; Narasimhan and Larson, 1985 [218]; Wang et al., 1997

[219]; Zhao et al., 1997[220]; Lewis and Dwyer-Joyce, 2001/2001/2002 [221–223]; Blau, 2009 [185]; Jaswin et al.,

2011 [208]; Forsberg et al., 2011/2012 [224,225]; Sawant and Jain, 2017 [195]. In a 2009 study, Blau et al. [226]

reported on the interaction of abrasion damage and oxidation in Nimonic 80A and Stellite 6B. The fundamental

mechanisms of sulfidation have been reviewed (Mrowec, 1980/1995[227]; Bornstein, 1996 [228]; Bornstein and

Allen, 1997[229] ) in addition to the sulfidation resistance of various valve materials (Ari-Gur et al., 1991 [190]).

The general factors contributing to valve “guttering”, a form of erosion corrosion, has been investigated by Arnold

et al, [230] (1988) and Scott et al.[231]. Evans et al. [211] (1995) studied the microstructural aging of Pyromet 31V

in service. In addition, numerous analyses to determine the root cause of valve failures have been conducted

(e.g.,[232,233]). The aforementioned studies relating to life limiting mechanisms and failure during valve operation

will be discussed in greater detail in the following sub-sections. In addition, numerous early studies on the

fundamentals of oxidation [234], frictional behavior [235], and role of oxides on the friction and wear of FeCr,

FeCrAl, NiCrAl, and CoCrAl alloys at elevated temperatures [236] were conducted by Stott and co-workers.

116
7.4.1 Valve Seat Wear

Wang et al. [219] (1997) investigated the mechanisms controlling HDDE intake valve seat and insert wear

using an experimental simulator. The valve and seat were installed in the simulator, which used a hydraulic actuator

to simulate high cycle valve opening and closing against the insert during combustion. Both load and temperature

were adjustable to simulate various engine operating conditions. Wear mechanisms at the intake valve seat/insert

interface are different than that occurring at the exhaust valve. The exhaust valve seat, while exposed to significantly

higher temperatures, is thought to contain a thin layer of lubricant due to combustion products which can reduce

wear damage in some cases. As such, wear at the intake valve seat/insert interface is sometimes more severe than the

exhaust valve. The authors studied the wear behavior of the valve seat after testing of various combinations of valve

(Silchrome 1 and Silchrome XB), insert, and seat facing materials by profilometry. The tests were conducted at

510 °C. The authors identified three primary wear mechanisms, including shear deformation (as noted by radial flow

of material towards the outer diameter of the valve seat), adhesion or microwelding (as characterized by material

transfer from the insert to the seat face or vice versa), and abrasion. The authors noted that abrasion in the form of

grooving on the seat face was primarily caused by hard carbide particles from the insert material. The use of higher

yield strength/hardness material and pairing materials with low coefficients of friction was suggested as a means to

reduce material wear associated with high shear strains at asperity contacts between seat face and insert. The

formation of an oxide film reduced the coefficient of friction between the seat face and insert as well as metal to

metal contact, which was beneficial to reducing wear associated with shear strain and adhesion. Reducing the

hardness differential of the valve seat and insert was also reported to reduce the wear of the softer material.

A second study by these same authors [237] aimed to determine the influence of operating parameter such

as load, temperature, and cycle time on valve seat and insert wear, with the valve seat manufactured from Silchrome

1 and the valve insert manufactured from Silchrome XB. The wear scar width and depth on the valve seat were as a

function of load, temperature, and the amount of cycles are show in Figure 69a, b, and c, respectively. In Figure 69a,

the wear scar width and depth increase relatively modestly with load up to 17,640 N but increase at a significantly

greater rate with increasing load from 17,640 N to 24,255 N, suggesting different wear mechanisms are occurring at

higher and lower loads. The authors noted that abrasive, adhesive, and shear strain controlled wear are occurring in

all tests but that adhesive wear dominated at lower loads while shear strain controlled wear dominated at higher

loads. It should be noted that lower valve loads were used in the study than would be equivalent to a typical peak

117
combustion pressure of 18 MPa for that time period but other factors including a narrower seat width, larger seat

angle of 45°, and offset seating (0.76 mm) resulted in more severe wear in the valve simulator compared to typical

operating conditions. These results indicate that as PCP rises in future HDDE, valve seat and insert wear will

become more severe, particularly due to shear strain controlled wear, and materials with higher yield strength and

hardness may be required. The authors reported a decrease in valve seat wear with increasing temperature from 180

to 650 °C for a valve loading of 17,640 N and 864,000 cycles and these results are show in Figure 69b. However,

the wear was significantly more mild in the temperature range of 500 to 650 °C compared to 180 to 400 °C. The

authors suggested this was due to an oxide film that forms at temperatures of ~500 °C and greater, which reduced

metal to metal contact and adhesive wear. The beneficial effect of oxide scales in reducing wear has also been

demonstrated by Stott and co-workers [235,236,238]. The temperature dependence of wear was determined for

valve loading conditions (17,640 N) where adhesive wear dominates. The temperature dependence of wear might

change under higher loading conditions where shear strain controlled where dominates. Under these conditions,

material strength plays a more critical role and is significantly reduced by increasing temperature (see Figure 62). In

Figure 69c, the wear results are shown as a function of the number of cycles at 510 °C and a load of 17,640 N. The

results indicate a significant drop in the wear rate as the number of cycles increases, which the authors attributed to

an increase in contact area and a commensurate decrease in contact stresses.

a 3 0.1 b3 0.2 c 3 0.14

0.09 0.18
0.12
2.5 0.08 2.5 0.16 2.5

Wear scar depth (mm)


0.1
Wear scar depth (mm)

Wear scar width (mm)


Wear scar width (mm)
Wear scar depth (mm)
Wear scar width (mm)

0.07 0.14

2 0.06 2 0.12 2
0.08
0.05 0.1
0.06
1.5 0.04 1.5 0.08 1.5

0.03 0.06 0.04

1 0.02 1 0.04 1
0.02
0.01 0.02

0.5 0 0.5 0 0.5 0


5000 10000 15000 20000 25000 100 200 300 400 500 600 700 0 1 2 3 4
Load (N) Temperature (°C) Cycles (106)

Figure 69 – Wear scar width and depth on a Silchrome 1 valve seat face as a function of (a) load after 864,000 cycles at 510 °C,
(b) temperature at a load of 17640 N after 864,000 cycles, and (c) the number of cycles at a load of 17,640 N and temperature 0f
510 °C. Silchrome XB was used as the insert material. This data is reproduced from [237].

118
Some of the same authors (Zhao et al [220]) performed a similar investigation using the same experimental

simulator to identify the wear mechanisms at the valve seat/insert interface in exhaust valves at 538 and 649 °C. The

valve materials studied were Pyromet 31, 23-8N, and 21-4N mod while the insert material was Silchrome XB for the

538 °C test and Eatonite 6 for the 649 °C test. At both temperatures, the materials ranking, from lowest wear rate to

greatest, was reported as follows: Pyromet 31, 23-8N, 21-4N mod. The Pyromet 31V also exhibited a narrower wear

band, indicating reduced shear deformation, which is likely due to its significantly higher yield and tensile strengths

at 650 °C compared to 23-8N and mod. 21-4N (see Figure 62).

In the early 2000s, Lewis and Dwyer-Joyce[221,223] developed a semi-empirical model to predict material

loss from the valve and insert at the valve/insert interface. The authors identified two main phenomena causing wear

at the valve and insert, impact of the valve on the seat during closure and frictional sliding due to combustion

pressure. The empirical model for the volume (V) of material removed by sliding and impact wear given by the

authors is:

(4)

where the first term in equation ((4) accounts for sliding wear and k, Pavg, x, and h are the wear coefficient (a

function of the sliding materials, lubrication conditions, and type of wear (adhesive, abrasive, corrosive, or fretting)),

average contact force (a function of cylinder pressure and contact geometry), the total sliding distance (the product

of the number of cycles and the sliding distance per cycle), and hardness of the softer of the two materials (valve or

insert), respectively. The second term in equation ((4) is related to wear due to the valve impact, where K and n are

empirically determined wear constants, e is the impact energy per cycle and N is the number of cycles. The impact

energy is equivalent to mv2/2 . Where m is the mass of the valve, follower, and half the mass of the spring, and v is

the velocity of the valve at impact. The valve velocity is itself a function of the engine speed, cam shaft profile,

clearance between valve tip and follower, and any misalignment of the valve and seat, among other factors. The

authors noted that by studying the individual material and engine parameters of the model, reductions in valve wear

can be made. For instance, from a materials perspective, reducing valve mass can reduce impact wear while

increasing material hardness can have a significant reduction in sliding wear, according to their model. A second

wear model was proposed in 2009 by Blau [185]. The model of Blau is similar to the model of Lewis and Dwyer-

Joyce in that it incorporates two wear contributions, a sliding component and a deformation component, which can

119
result from impact or tangential motion. However, the Blau model allows the sliding distance to evolve with wear,

incorporates the influence of temperature on friction coefficients, and adjusts the contributions of sliding wear based

on seat angle, among other factors.

If the valve material does not possess sufficient wear resistance, it often must be hardfaced with a material

of superior wear resistance. In 2011, Forsberg et al.[224] reported on the wear mechanisms between austenitic steel

valve that was hardfaced with Stellite F and a seat insert manufactured from Novofer AR20. Stellite F is a Co and Cr

based alloy with nominal composition shown in Table 7. Novofer AR20 is a high speed steel incorporating

intermetallic phases and high temperature solid lubricants (nominal: composition of 1.05C18.5Co-4.5Cr-15Cu-

0.9Mn-11.5Mo-0.45S-1.25Si-1.8V-3.5W). Cu fills the majority of the pores in Novofer AR20 to enhance cooling.

The valve with hardfacing and the insert material was tested under three different conditions, in a consumer truck

engine which was mildly driven for 249,000 mi., in a consumer truck engine that was relatively hard driven for

130,000 mi, and in an engine test cell for the equivalent of 249,000 hard mi. The microstructures of the Novofer

AR20 and Stellite F before testing are shown in Figure 70. An SEM image of the cross section of the Stellite F

material and corresponding energy dispersive x-ray (EDX) maps are in Figure 71 for the valve installed in the

engine which was driven “relatively hard” for 130,000 mi. A tribofilm of ~5-10 m in thickness, containing Ca, O,

P, S, and Zn, had built up on the Stellite F during operation. The authors noted that the tribofilm consisted of

elements found in common engine oil additives (e.g., zinc dithiophosphate, calcium sulfonates, etc.) and acted as a

protective layer for the Stellite F. Conversely, in the lab engine test cell, which was cleaner running, a cross section

of the material revealed no significant tribofilm had built up. Furthermore, significant plastic deformation of the

Stellite F occurred approximately 20 m below the surface of the valve from the lab engine test cell (see Figure 72)

and the surface was notably rougher, as measured by white light interference microscopy, and material transfer

(adhesion) had occurred. This study [224] and another by the same authors [225] illustrate the influence of

tribofilms and intentionally applied coatings on the wear characteristics, respectively.

120
Figure 70 – Typical microstructure of the seat insert material Novofer AR20 (left) and the valve hardfacing material Stellite F
(right) before engine testing. This figure is reproduced from [224].

Figure 71 – Cross section of the valve surface from the sample driven under hard conditions and corresponding EDX elemental
maps. A tribofilm ~5 m has built on the Stellite F. This figure reproduced from [224].

Figure 72 – Cross section of a valve Stellite F hardfacing after testing in an engine test cell. Notably, no tribofilm formed and a
deformnation ~20 m thick is present after operation in the engine test cell. In comparrison, a tribofilm formed after operation in
a real engine and served to protect the base metal, as shown in Figure 71. This figure reproduced from [224].

121
7.4.2 High Cycle Fatigue

Cyclical stresses due to combustion pressure, spring loads, and repeated valve seating are also imposed on

the valve. In addition, as peak cylinder pressures and temperatures continue to increase, valves become more

susceptible to chordal fracture. In addition, distortion of the head and seat insert, which can be more severe in

modern engines operating at higher PCP and temperatures, can result in poor valve seating and radial fatigue

cracking. These cyclical loads can act on all locations of the valve, including the combustion face, seat, stem, and

tip, etc. These cyclical loadings can result in fatigue and wear damage. Fatigue resistance is typically dramatically

reduced with increasing temperatures, which makes this parameter of significant importance for valve materials,

particularly as engines trend toward increasing temperatures. Hot rotating beam fatigue tests are often employed to

assess the fatigue resistance of valve materials at temperature [191,192,198,213]. In one example, hot rotating

bending fatigue tests were performed on VAT 32, VAT 36, Inconel 751 and Nimonic 80A [191]. Results are

provided in Figure 73. Each data point in Figure 73 represents an experimental series of multiple tests to establish

the stress which results in failure after 100 million cycles. The materials exhibit a significant reduction in fatigue

strength with increasing temperature. VAT 32 exhibited significantly lower fatigue strength than Nimonic 80A and

VAT 36 from RT up to 860 °C. The authors suggested that the higher volume fraction of carbides in VAT 32 were

detrimental to fatigue strength, as they provide more crack nucleation sites. Although austenitic steels are not shown

in Figure 73, hot rotating beam fatigue tests on these materials indicate their fatigue resistance at temperature is

significantly lower than the Ni based alloys [213,239]. It should also be noted that traditional fatigue testing is not an

ideal representation of the combustion chamber environment, where temperature transients, thermal stresses,

oxidation, sulfidation, and wear may act synergistically to reduce fatigue life.

122
Figure 73 – Hot rotating beam fatigue tests for alloys VAT 32, VAT 36, Inconel 751 and Nimonic 80A [191]. Each data point in
Figure 73 represents an experimental series of multiple tests to establish the stress at which failure occurs after 100 million
cycles. Reprinted with permission Copyright © 2017 SAE International. Further distribution of this material is not permitted
without prior permission from SAE.

7.4.3 Oxidation, Sulfidation Corrosion, and Guttering

Oxidation and sulfidation are two microstructural degradation mechanisms which occur in the diesel engine

combustion chamber environment. High temperature alloys are designed to form a protective oxide layer on the

surface which creates a diffusion barrier to oxygen and other corrosive species. Valve alloys form a protective oxide

scale (1 – 10 m thick, depending on temperature and alloy) that is typically chromium oxide-based, but almost

always consists of mixed oxide phases. The ideal protective scale is slow growing, dense, phase stable in the

temperature range of interest, and adherent to the metal surface. Oxide scale characteristics and rates of

formation/degradation are determined by temperature, environment, and the composition, structure and surface

condition of the alloy. There is often alloy design conflict between alloy chemical compositions that maximize high

temperature strength versus compositions that maximize oxidation resistance. In some cases, minor element

additions, and even trace elements, can have substantial positive or negative impact on alloy oxidation behavior. As

operating temperatures increase, oxidation becomes an increasingly important degradation mechanism, as

oxidation/scaling rates increase exponentially with temperature [119]. Oxidation resistance on wear surfaces, due to

abrasive loss of the protective oxide, can be particularly challenging for higher temperature valves. Ni based

superalloys tend to have better oxidation resistance than austenitic stainless steels [198] but are still susceptible to

oxidation degradation at higher temperatures. This is particularly important for valve materials which have exceeded

123
their maximum operating temperature due to engine overheating, which can result in oxide penetration deep into the

base material [240]. Such penetrations can lead to cracking and accelerated fatigue failure and have been observed

on both intake and exhaust valves. Figure 74 shows the mass change as a function of time during oxidation testing of

Inconel 751 and Pyromet 31V at different temperatures and in atmospheres (laboratory air or laboratory air plus

10% water vapor), as summarized from various studies [241–243]. In static laboratory air at 800 °C, the mass gain

behavior of Inconel 751 and Pyromet 31V exhibits parabolic kinetics and modest mass gain [241]. Similar mass gain

behavior of Inconel 751 is observed at 870 °C compared to at 800 °C over the first 500 h of oxidation testing [242].

However, as the test temperature is increased from 870 °C to 930 °C, the rate of mass gain in static air as well as

attack depth [242] increases dramatically for Inconel 751. The oxidation behavior of Inconel 751 and Pyromet 31V

is different in an atmosphere containing 10 vol.% water vapor compared to static air. When tested at 800 °C in air

plus 10 vol.% water vapor, the alloys exhibit para-linear behavior, with an initial mass gain similar to that observed

in static air, followed by slow mass loss attributed to CrO2(OH)2 evaporation. Pyromet 31V, when oxidation tested

in air plus 10 vol.% water vapor but at 850 °C, exhibits a much faster rate of mass gain initially before mass loss due

to CrO2(OH)2 evaporation begins to occur after ~1500 h of testing. With exhaust valve temperatures projected to

increase up to 900 °C in the coming decades (

124
Table 2), oxidation is likely to become an increasingly important degradation mechanism in exhaust valves.

751 lab air, 930°C


3.0 31V lab air +10% water vapor, 850°C
751 lab air, 870°C
31V lab air, 800°C
2.5 751 lab air, 800°C
751 lab air +10% water vapor, 800°C

Specimen mass change (mg cm-2)


31V lab air +10% water vapor, 800°C
2.0

1.5

1.0

0.5

0.0

-0.5
0 1000 2000 3000 4000 5000
Time (h)

Figure 74 – Specimen mass change during oxidation testing for Pyromet 31V and Inconel 751 in laboratory air or in laboratory
air plus 10% water vapor for temperatures ranging from 800 to 930 °C [241–243].

As mentioned previously, Stott and co-workers [235,236,238] have demonstrated the beneficial effects of

oxide layers on reducing wear at high temperatures by reducing metal to metal contact. In turn, surface damage and

wear due to mechanical contact, such as at the valve seat/inset interface during operation, can also influence the

characteristics of oxidation on valves. The interaction of wear and oxidation was studied by Blau et al.[226] on a

series of different alloys, including a martensitic alloy designated 465 with 11 wt.% Cr and 11 wt.% Ni, Pyromet

80A (which is similar to Nimonic 80A), and Stellite 6B. Samples were oxidized for 2 h at 850 °C, cooled to RT

where surface damage was introduced by scratch tests, then oxidation tested for an additional 4 h at 850 °C. A

second series of specimens were oxidized for 6 h then scratched and the grooves were compared, providing insight

into the differences in oxidation behavior of damaged and un-damaged surfaces in each material. The oxygen and

Cr concentration in the scratch groves was significantly greater in the specimens which received additional

annealing after scratching for all materials, suggesting that the defects resulting from the scratching accelerated local

oxidation in the scratch grooves during subsequent annealing. In the case of Pyromet 80a, Ti levels in the scratch

grooves of specimens scratched and annealed for an additional 4 h were higher than the bulk levels. Interestingly,

the hardness of the scratch grooves in Pyromet 80A increased with additional annealing of 4 h at 850 °C, which is

likely due in part to the redistribution of alloying elements in the scratch groove.

125
Hot corrosion in the form of sulfidation is also concern in the diesel engine combustion environment due to

the presence of sulfur along with barium, calcium, magnesium, zinc, and sodium in diesel fuel and engine oils which

can combine to form sulfide salt deposits on valves [190,231]. The mechanism of sulfidation in the diesel engine

atmosphere which is both sulfidizing and oxidizing is complex and dependent on numerous properties relating to the

compositions and structure of the base alloy, any oxide layer present, temperature of the base metal, composition of

the atmosphere, as well as other factors [227,244]. Inward diffusion of sulfur through chromia and alumina scales

has been shown to occur and can ultimately destroy/degrade the protective oxide scale [190,244]. Sulfide scales are

generally significantly inferior, and less protective compared to oxide scales, in part due to a significantly more

pronounced defect structure and lower thermodynamic stability, as well as similar thermodynamic stabilities of

multiple types of sulfides, which promotes greater heterogeneity [227]. In general, the more pronounced defect

structure of sulfides results in enhanced diffusion of elements through the sulfide scale relative to oxide scales at

similar temperatures, as reported by Mrowec [227,244]. As such, the rate constants for sulfidation of pure metals

and alloys is often orders of magnitude greater than for oxidation of the same alloy [227]. Consequently, sulfidation

may proceed at temperatures where oxidation is kinetically limited. A significant amount of work on the

fundamentals of sulfidation of Fe, Ni, and Cr alloys with additions of Mo, Nb, and Al was conducted by Douglass

and co-workers [245–252]. These studies typically consisted of binary or ternary model alloys where the influence

of large additions of Mo, Nb, and Al were evaluated, with the exception of a study on the influence of Al additions

on the sulfidation behavior of MP35N[247]. The studies generally showed that additions of Mo, Nb, and Al reduced

the rate of sulfidation, although typical valve alloys are more complex and have lower amounts of Mo, Nb, and Al,

and the role of these elements on sulfidation behavior at lower quantities may be significantly different.

Testing to determine the sulfidation resistance of typical valve alloys has also been conducted. Ari-Gur et

al. [190] investigated the sulfidation resistance of several diesel exhaust valve and hard facing alloys. In their work,

samples were placed in MgO crucibles containing a salt mixture of calcium, barium, and sodium sulfate with carbon

(ratios provided in [190]), which best simulated the diesel engine combustion environment at the time (1991)

according to the authors. The samples were immersed in the salt bath, which was held at 870 °C for 80 h. The

cylindrical specimens were then sectioned radially to investigate the extent and depth of corrosive attack. The depth

of corrosion penetration from the surface, as measured from SEM images, was determined and is reproduced here in

Figure 75. In the 23-8N sample, a sulfidized region which was previously pure base alloy developed underneath a

126
thin chromia layer. The sulfidized region contained nodules comprised of manganese and chromium sulfides,

suggesting sulfur diffusion through the chromia layer and into the base metal occurred. In the case of Pyromet 31V,

the chromia protective oxide layer was completely destroyed during testing and nodules of chromium and titanium

sulfide were present in the sulfidized region, consistent with the greater thermodynamic stability of chromium,

titanium, and manganese sulfides relative to iron and cobalt sulfides. The most severe corrosive attack occurred in

the Inconel 751 and the least severe in the 21-12N. Similarly, hard facing alloys with lower Ni content, such as

Stellite 6, Triballoy 400, and Eatonite 6 performed significantly better than the Ni based Eatonite 3. In general, the

penetration depth into the Fe and Co based alloys was significantly less deep relative to the Ni based alloys and

Pyromet 31V outperformed Inconel 751. The authors suggested that this may be in part due to the faster kinetics of

sulfur diffusion in Ni based alloys compared to Fe and Co based alloys. The Co based Triballoy 400 had the best

sulfidation resistance among the hardfacing alloy. In Ni-Cr alloys, substantial increases in the rate of sulfidation

were measured with increasing temperature from 800 to 900 °C, which is in the range for valve operating

temperatures in next generation HDDE [227,244,253].

Figure 75 – Depth of corrosion penetration into various exhaust valve materials after immersion in a salt bath of calcium, barium,
and sodium sulfate with carbon at 870 °C for 80 h [190].

127
Valve guttering is a complex mechanical and corrosion phenomenon that is a failure mode in HDDE

exhaust valves which results in a path or “gutter” radially across the seat face for exhaust gas to escape when the

valve is closed [230,231]. Causes of valve guttering are thought to include valve or insert distortion, which is known

to occur in modern HDDE, face peening, and combustion deposits or spalling of deposits, all of which may interfere

with proper valve seating. This improper valve seating can cause higher valve head temperatures as heat conduction

through the contact at the valve seat maybe reduced. In addition, hot exhaust gas can escape, cause additional

deposits to spall, and in many cases, accelerated attack, often intergranular, and erosion of the valve material. In the

study by Scott et al.[231], the structure and chemical composition of deposits on exhaust valves of HDDE that

formed during operation were dependent on the type of oil, but generally contained sulfur. Consequently, sulfidation

corrosion may act during and facilitate the development of valve gutters during operation.

7.4.4 Long Term Material Aging

Numerous microstructural degradation mechanisms act on the valve materials during operation. Continued

exposure to the high temperatures and the long service lifetimes of valves for HDD engines leads to microstructural

changes of the base metal, which can be detrimental to the durability of the material. Evans et al.[211] investigated

the microstructure of solutionized and age-hardened Pyromet 31V in valves after being in service in natural gas

reciprocating engines for times of 2400 h. The valve materials exposed to two different temperatures during the

2400 h, were evaluated. The  particles, shown in Figure 76a and b exhibit average diameters of 158 and 171 nm

after 2400 h of service, respectively, suggesting the temperature of the material corresponding to Figure 76b was

exposed to higher temperatures. These particles are considerably larger than the  particles observed in the

solutionized and aged condition (average diameters of ~48 nm) prior to service, which are shown in Figure 65,

indicating significant coarsening occurred. While not shown here, the M23C6 carbides also underwent coarsening

during valve operation as well. However, the influence of temperature on microstructural changes/coarsening during

valves in service is difficult to quantify due to uncertainty of the actual valve operating temperatures and substantial

temperature gradients that are present in valves during engine operation.

A follow on study by some of the same authors investigated the coarsening kinetics and changes in

hardness of Pyromet 31V under controlled laboratory conditions (aging in furnaces) at temperatures of 700, 750, and

128
800 °C [254]. The specimens were solution treated at 1093 °C for 4 h followed by annealing at 732 °C for 4 h, then

aged for 100, 1000, and 3000 h at the aforementioned temperatures. Figure 77a and b show the coarsening behavior

of the  particles (presented as the average particle radius cubed, r3, which is proportional to the spherical particle

volume) and the Vickers hardness of the Pyromet 31V as a function of aging time, respectively. The value of r3

increases approximately linearly at the three temperatures.  particles exhibited minimal coarsening at 700 °C over

the course of 3000 h and significantly greater coarsening at 750 °C. Remarkably, the particle volume increased by

approximately an order of magnitude after aging for 3000 h at 800 °C. By applying the Lifshitz, Slyozov, and

Wagner (LSW) theory of particle coarsening, the authors were able to calculate an activation energy of 338 kJ mol -1

for volume diffusion controlling particle coarsening in Pyromet 31V, which can then be utilized to predict

coarsening rates at different temperatures or estimate valve temperatures during in service operation from  particle

size measurements. The aging of the microstructure can also lead to changes in the hardness of the material. As

shown in Figure 77b, aging at 700, 750 °C, 800 °C for 100 h all led to increases in the hardness of Pyromet 31V

compared to the solution treated and annealed condition. Conversely, aging from 100 to 1000, to 3000 h resulted in

an apparent monotonic decrease in hardness of the Pyromet 31V at all test temperatures, with the hardness being

reduced below that of the solution treated and annealed condition after prolonged aging at 800 °C. While particle

coarsening undoubtedly affects the hardness of the material, changes in the particle volume, grain size, and other

factors that also occur during long term aging will also influence the hardness of the material. As maximum

temperatures on exhaust valves increase from ~800 to 900 °C in future HDDE, particle coarsening and

microstructural changes during long term operation will become a greater concern in regards to the durability of the

material. A key criteria in the development of next generation valve materials is alloying strategies to reduce inter

diffusion kinetics and minimize the  coarsening rate [193].

129
Figure 76 – Centered dark-field image of  particles in the gamma matrix after service aging for 2400 h at (a) 650 °C and (b) 700
°C. The average diameters after aging at 650 and 700 °C are 158 and 171, respectively, compared to an average diameter of ~48
nm in the solutionized and aged condition, indicating significant coarsening occurs during operation in-service. These images are
reproduced from [211]. Copyright 2005 by The Minerals, Metals & Materials Society. Used with permission.

100 420
No aging
700°C 700°C
90 750°C 750°C
800°C 400 800°C
80

70
380
Hardness (Hv)
r3 (nm3) x 10-4

60

50 360

40
340
30

20
320
10

0 300
0 1000 2000 3000 0 1000 2000 3000
Aging Time (h) Aging Time (h)

Figure 77 –  particle radius cubed and (b) RT Vickers hardness of Pyromet 31V as a function of aging time at temperatures of
700, 750, and 800 °C. The Pyromet 31V was solution treated at 1093 °C for 4 h followed by annealing at 732 °C prior to aging.

7.4.5 Failure Analysis Case Studies

Yu and Xu [233] performed a failure analysis of a diesel engine valve manufactured from 21-4N steel that

failed during service. Figure 78 shows an SEM image of the fractured region steel that failed in service [233]. Both

circumferential and radial cracks are present and indicated by arrows. Radial cracks are often due to distortion,

which can result in poor seating between the valve and insert. The fractures exhibit typical fatigue features, such as

beach marks and fatigue striations. X-ray diffraction and metallography of the failed valve and an identical new

valve of the same material processed in the same manner indicated important microstructural changes in the 21-4N

occurred during service. Among these changes were precipitation of M23C6 carbides and the σ phase, which can

130
embrittle the material [119,211], as well as the formation of a lamellar structure of austenite and M23C6 in only

certain grains, which was undesirable. The microstructure of the material from the new valve contained an abnormal

grain size distribution. The M23C6 carbides were predominately present on the grain boundaries of the relatively

small grains. The authors suggested that the valve failure might have been the result of an improper solutionizing or

aging treatment. For instance, performing the aging at higher than optimal temperatures could result in overaging of

the material, coarsening of carbides, and the precipitation of unwanted phases, which could have set the stage for

additional undesirable microstructural changes during service.

Figure 78 – SEM composite image of a fracture surface from a valve manufactured from 21-4N steel that failed during operation
in a diesel engine. The figure is reproduced from [233].

In a different study [232], a stem fracture that occurred in a valve manufactured from 21-4N steel operating

in an automobile diesel engine was analyzed. Figure 79 shows SEM and corresponding optical micrographs of the

stem fracture. The author of the failure analysis study noted that two cracks were initiated in the valve stem and that

no other material defects or corrosion products were present near the crack initiation sites. Furthermore, finite

element analysis on the stem in the area of fracture indicated that the stress near the fracture (due to non-uniform

131
temperature field and a spring loading) was too low to result in the fatigue fracture. However, significant C deposits

were observed on the seats, and FEA analysis revealed that these deposits could impart a significant bending

moment on the valve stem during seating. As such, the author concluded that the stem fracture was caused by

fatigue which was assisted by a large bending stress that was introduced during valve seating, likely caused by

improper valve seating. It should also be noted that relatively high temperatures occur in the valve stem at the area

of the fatigue crack initiation under normal operation, which likely lowered the fatigue resistance of the material.

While these examples of valve failure clearly indicate the need for materials with high fatigue strength, they also

emphasize how other factors such as initial processing of the material, alignment, and valve seating can exacerbate

failures. Poor valve seating is becoming an increasing concern due to greater elastic and plastic distortion of the

cylinder head and seat inserts as a result of combustion processes, thermal cycling, and long term exposure to

elevated temperatures.

Figure 79 – (left) SEM image and (right) optical image of the fracture surface at the stem of a valve manufactured from 21-4N
steel. Crack zones are labelled as follows: C: first crack origin, D: fatigue beach of first crack, E: initiation point of second crack
when the first crack was in zone D, F: zone of propagation of second crack, G: rupture zone. This figure is reproduced from
[232].

7.5 PATHWAYS FOR FUTURE MATERIALS DEVELOPMENT

Exhaust valve materials trends have historically been driven by the need for greater high temperature

mechanical properties at the lowest cost possible. These improved high temperature properties, such as fatigue

strength, hardness, wear resistance, and oxidation/corrosion resistance, are generally required due to increasing

cylinder pressures and temperatures as diesel engine trends move towards higher power density and thermal

efficiency (see

132
Table 2). Improvements in valve materials performance and without additional cost are needed going

forward. Improvements in the design and processing of both Fe and Ni based alloys, as well as hardfacing and

coating materials, should be considered.

Historically, alloy design and improvements of austenitic steels for exhaust valve applications took place

up to the mid 1980s. In the mid 1970s to mid 1980s, steel alloys 23-8N and modified 21-4N were deployed and are

still used today in some HD on road diesel applications, in part due to their significantly lower cost than alloys with

greater Ni content. To the the authors’ knowledge, no new austenitic steel valve alloys have been deployed in

significant quantities for exhaust valve applications in HD on road diesel engines in the United States since the mid

1980s. The mechanical properties of austenitic valve steels are clearly inferior to alloys with intermediate Ni content

and Ni based superalloys discussed herein (see Figure 62), due to the different strengthening mechanisms employed

and lower heat and corrosion resistance of the austenitic steels materials. The mechanical properties of current

austenitic valve steels generally are not considered suitable for operating temperatures greater than 800 °C [212].

Maximum exhaust valve temperatures are expected to exceed 800 °C in future diesel engines which would require

the optimization and/or development of new austenitic steel valve alloys with enhanced heat resistance or coatings

sufficient to reduce the temperature of the base material. However, austenitic steels for valve applications were

developed three to four decades ago. As such, significant potential may still exist to optimize and/or develop new

cost effective austenitic steel materials with improved mechanical properties, potentially using alternative

strengthening mechanisms, assisted by the use of advanced computation tools and the greater understanding of these

alloys that is available today.

Wear at the valve seat/insert interface is a major cause of valve degradation and failure. Improved coating

and/or hardfacing materials and processes that can reduce the temperature and damage of the substrate, while also

improving the wear characteristics should be evaluated. Physical vapor deposition of CrN, AlCrN, and TiAlCrN

coatings have shown initial promise to reduce wear rate but the durability of the coatings requires improvement

[225]. In addition, -PTA has been evaluated for applying hard facing materials and has enabled smaller heat

affected zones, faster cooling rates, and improved microstructural characteristics (e.g., finer dendrite arm spacing)

compared to conventional PTA [195]. Similarly, thermal spray coatings may enhance the resistance of valve

materials to oxidation and sulfidation in next generation HDDE where temperatures are expected to increase

133
substantially [255]. Laser surface processing is also being explored as a means to increase the hardness of gray cast

iron, making it suitable for valve seating, and eliminating the need for separate inserts in the cylinder head [256].

The strength and temperature limitations of current austenitic steels and excessive cost of Ni based

superalloys have led to the recent development of intermediate Ni based alloys [191,212–214]. In general, these

alloys contain lower amounts of costly Ni, but utilize novel alloy design strategies to provide similar or greater

mechanical properties as some Ni based superalloys, as show in Figure 62. Research efforts are ongoing in this area

to further develop and optimize more affordable versions of higher temperature materials [214].

Manufacturing poppet valves can require multiple high temperature heat treatments, machining steps,

joining steps (in the case of two-piece valves), and in the case of some alloys, chrome plating, nitriding, and/or valve

hardfacing. These steps add significant energy input, labor, and/or cost. New advanced manufacturing techniques

that may be able to simplify valve production and reduce the energy intensity of valve manufacturing should be

evaluated. Furthermore, advanced coatings for oxidation and thermal resistance may enable existing materials to

otherwise operate at higher temperature and should be considered.

134
8 SUMMARY AND CONCLUSIONS

Historical increases in peak cylinder pressures and temperatures in HD on-road diesel engines have resulted

in greater efficiencies. However, the increased peak cylinder pressures and temperatures impart greater thermal and

mechanical loads on engine components, to the point that many high temperature components in current production

HDDE are at or near their limits of required durability. This paper reviewed the historical evolution and major

metallurgical advancements of high temperature materials in HD on road diesel engines (10-15 L displacement) up

to the current state of the art, focusing on materials in the engine block, cylinder heads, pistons, valves, and exhaust

components. These components cover a wide range of materials classes, including cast iron, martensitic steel,

austenitic steel, nickel based super-alloys, and high temperature coatings. The life limiting mechanisms of the

materials associated with the complex mechanical and thermal loading during service (e.g., thermal fatigue,

mechanical fatigue, wear, creep, oxidation, sulfidation etc.) were identified and discussed. Based on these

microstructural limitations of current materials, key areas for future materials research are suggested that overcome

technical barriers while considering cost and manufacturing realities of the industry.

9 ACKNOWLEDGEMENTS

Compilation of this review paper was sponsored by the Propulsion Materials Program, DOE Vehicle

Technologies Office. The authors gratefully acknowledge Victoria Bouwens (Eaton Co.), Sandy Schaefer and

Sundaram Narasimhan (both formerly of Eaton Co.), Dieter Gabriel (MAHLE Engine Components USA, Inc.), and

Michael Brady, Yukinori Yamamoto, Charles Finney, Ercan Cakmak, and Bruce Pint, all of ORNL, for insightful

discussions regarding this manuscript.

10 DISCLAIMER

This report was prepared as an account of work sponsored by an agency of the United States Government.

Neither the United States Government nor any agency thereof, nor any of their employees, makes any warranty,

express or implied, or assumes any legal liability or responsibility for the accuracy, completeness, or usefulness of

any information, apparatus, product, or process disclosed, or represents that its use would not infringe privately

owned rights. Reference herein to any specific commercial product, process, or service by trade name, trademark,

manufacturer, or otherwise does not necessarily constitute or imply its endorsement, recommendation, or favoring

135
by the United States Government or any agency thereof. The views and opinions of authors expressed herein do not

necessarily state or reflect those of the United States Government or any agency thereof.

11 REFERENCES

[1] Davis S, Williams S, Boundy R. Transportation Energy Data Book. 35th ed. 2016.
[2] Annual Energy Outlook 2017. U.S. Energy Information Administration; 2017.
[3] Megel M, Westmoreland B, Jones G, Phillips F, Eberle D, Tussing M, et al. Development of a Structurally
Optimized Heavy Duty Diesel Cylinder Head Design Capable of 250 Bar Peak Cylinder Pressure Operation.
SAE; 2011.
[4] Davis SC, Williams SE, Boundy RG, Moore S. Vehicle Technologies Market Report. 2015.
[5] Jääskeläinen H. Early History of the Diesel Engine. 2013.
[6] Most efficient 4-stroke diesel engine. Guinness World Rec n.d.
http://www.guinnessworldrecords.com/world-records/378372-most-efficient-4-stroke-diesel-engine
(accessed April 7, 2017).
[7] 21st Century Truck Partnership, Roadmap and Technical White Papers. 21st Century Truck Partnership;
2013.
[8] Koeberlein D. Cummins SuperTruck Program, Technology and System Level Demonstration of Highly
Efficient and Clean, Diesel Powered Class 8 Trucks 2014.
[9] Tamm J, Spears M. Looking Ahead To The Next Phase Of Heavy-Duty Greenhouse Gas And Fuel
Efficiency Standards 2014.
[10] Exxon Mobil. 2018 Outlook for Energy: A View to 2040. 2017.
[11] Sripad S, Viswanathan V. Performance Metrics Required of Next-Generation Batteries to Make a Practical
Electric Semi Truck. ACS Energy Lett 2017;2:1669–73.
[12] EPA and NHTSA Adopt Standards to Reduce Greenhouse Gas Emissions and Improve Fuel Efficiency of
Medium- and Heavy-Duty Vehicles for Model Year 2018 and Beyond. Environmental Protection Agency,
Office of Transportation and Air Quality; 2016.
[13] Primus RJ. Visual Thermodynamics: Processes in Log(p)-Log(T) Space. SAE; 1999.
[14] Heywood J. Internal Combustion Engine Fundamentals. 1 edition. New York: McGraw-Hill Education;
1988.
[15] National Research Council. Cost, Effectiveness, and Deployment of Fuel Economy Technologies for Light-
Duty Vehicles. 2015.
[16] Stanton DW. Systematic Development of Highly Efficient and Clean Engines to Meet Future Commercial
Vehicle Greenhouse Gas Regulations. SAE; 2013.
[17] WORKSHOP REPORT: Trucks and Heavy-Duty Vehicles Technical Requirements and Gaps for
Lightweight and Propulsion Materials. U.S. Department of Energy; 2013.
[18] Kocher L. Cummins 55% Brake Thermal Efficiency Project 2017.
[19] Milam DM, Donaldson GE, Easley WL, Bond MS, Roozenboom SD, Brucker JR, et al. Demonstration of a
50% Thermal Efficient Diesel Engine - Including HTCD Program Overview 2006.
[20] Eilts P, Stoeber-Schmidt C-P, Wolf R. Investigation of Extreme Mean Effective and Maximum Cylinder
Pressures in a Passenger Car Diesel Engine, SAE; 2013.
[21] Luff DC, Law T, Shayler PJ, Pegg I. The Effect of Piston Cooling Jets on Diesel Engine Piston
Temperatures, Emissions and Fuel Consumption. SAE Int J Engines 2012;5:1300–11.
[22] Primus RJ. The Evolution of Diesel Engine Performance Prediction 2014:V002T08A001.
[23] Finney C, Simunovic S, Wereszczak AA, Muralidharan G, Haynes A. Applied Computational Methods for
New Propulsion Materials, (PM057), Task 4: Future engine requirements 2016.
[24] Gibble J. Volvo SuperTruck Powertrain Technologies for Efficiency Improvement 2012.
[25] Kass MD, Thomas JF, Wilson D, Lewis SA, Sarles A. Assessment of Corrosivity Associated With Exhaust
Gas Recirculation in a Heavy-Duty Diesel Engine, 2005.
[26] Sharp C, Webb CC, Yoon S, Carter M, Henry C. Achieving Ultra Low NOX Emissions Levels with a 2017
Heavy-Duty On-Highway TC Diesel Engine - Comparison of Advanced Technology Approaches. SAE Int J
Engines 2017;10.

136
[27] Sharp C, Webb CC, Neely G, Carter M, Yoon S, Henry C. Achieving Ultra Low NO X Emissions Levels
with a 2017 Heavy-Duty On-Highway TC Diesel Engine and an Advanced Technology Emissions System -
Thermal Management Strategies. SAE Int J Engines 2017;10.
[28] Sharp C, Webb CC, Neely G, Sarlashkar JV, Rengarajan SB, Yoon S, et al. Achieving Ultra Low NOx
Emissions Levels with a 2017 Heavy-Duty On-Highway TC Diesel Engine and an Advanced Technology
Emissions System - NOx Management Strategies. SAE Int J Engines 2017;10.
[29] Knothe G. Biodiesel and renewable diesel: A comparison. Prog Energy Combust Sci 2010;36:364–73.
[30] Sheehan J, Camobreco V, Duffield J, Graboski M, Shapouri H. An Overview of Biodiesel and Petroleum
Diesel Life Cycles. National Renewable Energy Laboratory; 1998.
[31] Ogunkoya D, Roberts WL, Fang T, Thapaliya N. Investigation of the effects of renewable diesel fuels on
engine performance, combustion, and emissions. Fuel 2015;140:541–54. doi:10.1016/j.fuel.2014.09.061.
[32] Knothe G, Sharp CA, Ryan TW. Exhaust Emissions of Biodiesel, Petrodiesel, Neat Methyl Esters, and
Alkanes in a New Technology Engine. Energy Fuels 2006;20:403–8. doi:10.1021/ef0502711.
[33] Haseeb ASMA, Fazal MA, Jahirul MI, Masjuki HH. Compatibility of automotive materials in biodiesel: A
review. Fuel 2011;90:922–31. doi:10.1016/j.fuel.2010.10.042.
[34] Zieher F, Langmayr A, Jelatancev A, Wieser K. Thermal Mechanical Fatigue Simulation of Cast Iron
Cylinder Heads. SAE; 2005.
[35] Hazime R, Seifert T, Kessens J, Ju F. Lifetime Assessment of Cylinder Heads for Efficient Heavy Duty
Engines Part II: Component-Level Application of Advanced Models for Thermomechanical Fatigue Life
Prediction of Lamellar Graphite Cast Iron GJL250 and Vermicular Graphite Cast Iron GJV450 Cylinder
Heads. SAE Int J Engines 2017;10:350–8.
[36] Seifert T, von Hartrott P, Boss K, Wynthein P. Lifetime Assessment of Cylinder Heads for Efficient Heavy
Duty Engines Part I: A Discussion on Thermomechanical and High-Cycle Fatigue as Well as
Thermophysical Properties of Lamellar Graphite Cast Iron GJL250 and Vermicular Graphite Cast Iron
GJV450. SAE Int J Engines 2017;10:359–65.
[37] Dawson S. Compacted Graphite Iron - A Material Solution for Modern Engine Design. SAE 2011;2011-01–
1083.
[38] Dawson S. Compacted graphite iron: Cast iron makes a comeback. JOM 1994;46:44–7.
[39] Guesser WL, Duran PV, Krause W. Compacted Graphite Iron for Diesel Engine Cylinder Blocks, La Société
des Ingénieurs de l’Automobile (SIA); 2004.
[40] Ghodrat S, Riemslag TAC, Kestens LAI, Petrov RH, Janssen M, Sietsma J. Effects of Holding Time on
Thermomechanical Fatigue Properties of Compacted Graphite Iron Through Tests with Notched Specimens.
Metall Mater Trans A 2013;44:2121–30.
[41] Ghodrat S, Riemslag AC, Janssen M, Sietsma J, Kestens LAI. Measurement and characterization of
Thermo-Mechanical Fatigue in Compacted Graphite Iron. Int J Fatigue 2013;48:319–29.
[42] Mellouli D, Haddar N, Köster A, Toure AM-L. Thermal fatigue of cast irons for automotive application.
Mater Des 2011;32:1508–14.
[43] Rejowski E, Soares E, Uehara S. High Value Gray Cast Iron Material for Heavy Duty Diesel Cylinder
Liners, 2011.
[44] Sreenivasan V., Dhanasekaran S, Sharma S, Prasad MS. Sliding Wear Behavior of Compacted Graphite Iron
Cylinder Liner Material, 2014.
[45] Jackson P, Addis B, Thakare S. Design and Validation of a New 13L Heavy-Duty Diesel Engine Using
Analysis-Led Design, 2008.
[46] Davis JR, editor. ASM Specialty Handbook Cast Irons - ASM International. ASM International; 1996.
[47] Dawson S, Hollinger I, Robbins M, Daeth J, Reuter U, Schulz H. The Effect of Metallurgical Variables on
the Machinability of Compacted Graphite Iron. SAE; 2001.
[48] Park SH, Kim JM, Kim HJ, Ko SJ, Park HS, Lim JD. Development of a Heat Resistant Cast Iron Alloy for
Engine Exhaust Manifolds, 2005.
[49] Gundlach RB. Effects of Alloying Elements on the Elevated Temperature Properties of Gray Irons. AFS
Trans 1983;91:389–422.
[50] Park YJ, Gundlach RB, Thomas RG, Janowak JF. Thermal Fatigue Resistance of Gray and Compacted
Graphite Irons. AFS Trans 1985:415–22.
[51] Gundlach RB. Elevated Temperature Properties of Alloyed Gray Irons for Diesel Engine Components. AFS
Trans 1978;86:55–64.
[52] Gundlach RB. Thermal Fatigue Resistance of Alloyed Gray Irons for Diesel Engine Components. AFS
Trans 1979;87:551–60.

137
[53] Turnbull GK, Wallace JF. Molybdenum Effect on Gray Iron Elevated-Temperature Properties 1959:35–46.
[54] Janowak JF, Gundlach RB. A Modern Approach To Alloying Gray Iron. AFS Trans 1982;90.
[55] Altstetter JD, Nowicki RM. Compacted Graphite Iron - Its properties and automotive applications. AFS
Trans 1982;82–188:959–70.
[56] Ziegler KR, Wallace JF. The effect of matrix structure and alloying on the properties of compacted graphite
iron. AFS Trans 1984;92:735–48.
[57] Sponseller DL, Scholz WG, Rundle DF. Development of Low-Alloy Ductile Irons for Service at 1200-1500
F. Am Foundry Soc Trans 1968;76:353–68.
[58] Norman V, Skoglund P, Moverare J. Damage evolution in compacted graphite iron during
thermomechanical fatigue testing. Int J Cast Met Res 2016;29:26–33.
[59] Diaconu VL, Sjögren T, Skoglund P, Diószegi A. Influence of molybdenum alloying on thermomechanical
fatigue life of compacted graphite irons. Int J Cast Met Res 2012;25:277–86.
[60] Diaconu VL, Sjögren T, Skoglund P, Diószegi A. Stress relaxation of compacted graphite iron alloyed with
molybdenum. Int J Cast Met Res 2013;26:51–7.
[61] Selin M. Tensile and Thermal Properties in Compacted Graphite Irons at Elevated Temperatures. Metall
Mater Trans A 2010;41:3100–9.
[62] Guesser WL, Masiero I, Melleras E, Cabezas CS. Thermal Conductivity of Gray Iron and Compacted
Graphite Iron Used for Cylinder Heads. Rev Matér 2005;10:265–72.
[63] Guesser WL, Martins LPR. Stiffness and Vibration Damping Capacity of High Strength Cast Irons. SAE;
2016.
[64] Gorny M, Lelito J, Kawalec M, Sikora G. Thermal Conductivity of Thin Walled Compacted Graphite Iron
Castings. ISIJ Int 2015;55:1925–31.
[65] Holmgren D, Källbom R, Svensson IL. Influences of the Graphite Growth Direction on the Thermal
Conductivity of Cast Iron. Metall Mater Trans A 2007;38:268–75.
[66] Holmgren DM, Diószegi A, Svensson IL. Effects of transition from lamellar to compacted graphite on
thermal conductivity of cast iron. Int J Cast Met Res 2006;19:303–13.
[67] Fenech H, Rohsenow WM. Prediction of Thermal Conductance of Metallic Surfaces in Contact. J Heat
Transf 1963;85:15–24.
[68] Kempers H. Giesserei 1966;53:15–8.
[69] Norman V, Skoglund P, Leidermark D, Moverare J. The effect of superimposed high-cycle fatigue on
thermo-mechanical fatigue in cast iron. Int J Fatigue 2016;88:121–31.
[70] Metzger M, Seifert T. Computational assessment of the microstructure-dependent plasticity of lamellar gray
cast iron – Part II: initial yield surfaces and directions. Int J Solids Struct 2015;66:194–206.
[71] Metzger M, Seifert T. Computational assessment of the microstructure-dependent plasticity of lamellar gray
cast iron – Part I: Methods and microstructure-based models. Int J Solids Struct 2015;66:184–93.
[72] Da Silva MB, Naves VTG, De Melo JDB, De Andrade CLF, Guesser WL. Analysis of wear of cemented
carbide cutting tools during milling operation of gray iron and compacted graphite iron. Wear
2011;271:2426–32.
[73] Skvarenina S, Shin YC. Laser-assisted machining of compacted graphite iron. Int J Mach Tools Manuf
2006;46:7–17.
[74] Guo Y, Stalbaum T, Mann J, Yeung H, Chandrasekar S. Modulation-assisted high speed machining of
compacted graphite iron (CGI). J Manuf Process 2013;15:426–31.
[75] Balachandran U. High Performance Electrical and Thermal Conductors 2016.
[76] Ekström M, Thibblin A, Tjernberg A, Blomqvist C, Jonsson S. Evaluation of internal thermal barrier
coatings for exhaust manifolds. Surf Coat Technol 2015;272:198–212.
[77] Yonushonis TM. Overview of thermal barrier coatings in diesel engines. J Therm Spray Technol 1997;6:50–
6.
[78] Schulz U, Leyens C, Fritscher K, Peters M, Saruhan-Brings B, Lavigne O, et al. Some recent trends in
research and technology of advanced thermal barrier coatings. Aerosp Sci Technol 2003;7:73–80.
[79] Mahade S, Curry N, Björklund S, Markocsan N, Nylén P. Thermal conductivity and thermal cyclic fatigue
of multilayered Gd2Zr2O7/YSZ thermal barrier coatings processed by suspension plasma spray. Surf Coat
Technol 2015;283:329–36.
[80] Mahade S, Curry N, Björklund S, Markocsan N, Nylén P, Vaßen R. Functional performance of Gd 2 Zr 2 O
7 /YSZ multi-layered thermal barrier coatings deposited by suspension plasma spray. Surf Coat Technol
2017;318:208–16.
[81] Introduction To Turbochargers. Indianapolis, Indiana: Schwitzer Turbochargers; 1991.

138
[82] Cummins X15 tour in New 2018 Peterbilt 389. You Tube 2017.
https://www.youtube.com/watch?v=TiKfsSk5YPQ (accessed November 16, 2017).
[83] Brady MP, Muralidharan G, Leonard DN, Haynes JA, Weldon RG, England RD. Long-Term Oxidation of
Candidate Cast Iron and Stainless Steel Exhaust System Alloys from 650 to 800 °C in Air with Water
Vapor. Oxid Met 2014;82:359–81.
[84] Ekström M. Oxidation and corrosion fatigue aspects of cast exhaust manifolds. KTH Royal Institute of
Technology, 2015.
[85] Ekström M, Jonsson S. High-temperature mechanical- and fatigue properties of cast alloys intended for use
in exhaust manifolds. Mater Sci Eng A 2014;616:78–87.
[86] Shyam A, Hawkins S, Erdman D, England R, Muralidharan G. Constrained Thermal Fatigue Performance of
Several Cast Ferrous Alloys. Mater Sci Forum 2014;783–786:2388–93.
[87] High Temperature Materials for Exhaust Manifolds (J2515 Ground Vehicle Standard). SAE 1999.
[88] Roehrig K, Fairhurst W. High-silicon nodular irons. Foundry Trade J 1979.
[89] Black B, Burger G, Logan R, Perrin R, Gundlach R. Microstructure and Dimensional Stability in Si-Mo
Ductile Irons for Elevated Temperature Applications. SAE; 2002.
[90] Maziasz PJ, Pint BA. High-Temperature Performance of Cast CF8C-Plus Austenitic Stainless Steel. J Eng
Gas Turbines Power 2011;133:092102-092102–5.
[91] Asteman H, Svensson J-E, Norell M, Johansson L-G. Influence of Water Vapor and Flow Rate on the High-
Temperature Oxidation of 304L; Effect of Chromium Oxide Hydroxide Evaporation. Oxid Met 2000;54:11–
26.
[92] Quadakkers WJ, Żurek J, Hänsel M. Effect of water vapor on high-temperature oxidation of FeCr alloys.
JOM 2009;61:44–50.
[93] Kleiner S, Track K. SiMo1000 Ein aluminiumlegiertes Gusseisen für Hochtemperaturanwendungen.
Giesserei 2010;97:28–34.
[94] Fontaine PI. Ni-Resist Type D5S – An Improved Material for Turbocharger Housings, 1980.
[95] Maziasz PJ, Shingledecker JP, Evans ND, Pollard MJ. Publication: CF8C-Plus heat-resistant cast stainless
steel. Adv Mater Process 2008:27–9.
[96] Evans ND, Maziasz PJ, Shingledecker JP, Pollard MJ. Structure and Composition of Nanometer-Sized
Nitrides in a Creep-Resistant Cast Austenitic Alloy. Metall Mater Trans A 2010;41:3032–41.
[97] Maziasz PJ, McGreevy T, Pollard MJ, Siebenaler CW, Swindeman RW. Heat and corrosion resistant cast
CN-12 type stainless steel with improved high temperature strength and ductility. US7255755B2, 2007.
[98] Brady MP, Yamamoto Y, Santella ML, Maziasz PJ, Pint BA, Liu CT, et al. The development of alumina-
forming austenitic stainless steels for high-temperature structural use. JOM 2008;60:12.
[99] Maziasz PJ, Shyam A, Evans ND, Pattabiraman K (Honeywell TT. Cast Cf8c-Plus Stainless Steel for
Turbocharger Applications. Oak Ridge National Laboratory (ORNL), Oak Ridge, TN (United States); 2010.
[100] Öberg C. Thermal Cycling, Creep- and Tensile Testing of Cast Exhaust Materials at Elevated Temperatures.
KTH Royal Institute of Technology, 2018.
[101] Norman V, Skoglund P, Leidermark D, Moverare J. Damage mechanisms in silicon-molybdenum cast irons
subjected to thermo-mechanical fatigue. Int J Fatigue 2017;99, Part 2:258–65.
[102] Lübbehusen M, Mehrer H. Self-diffusion in α-iron: The influence of dislocations and the effect of the
magnetic phase transition. Acta Metall Mater 1990;38:283–92.
[103] Herrmann J, Inden G, Sauthoff G. Deformation behaviour of iron-rich iron-aluminium alloys at high
temperatures. Acta Mater 2003;51:3233–42.
[104] Ekström M, Szakalos P, Jonsson S. Influence of Cr and Ni on High-Temperature Corrosion Behavior of
Ferritic Ductile Cast Iron in Air and Exhaust Gases. Oxid Met 2013;80:455–66.
[105] Xiang SM, Zhu BH, Jonsson S. High-Temperature Corrosion-Fatigue Behavior of Ductile Cast Irons for
Exhaust Manifolds Applications. Mater Sci Forum 2018;925:369–76.
[106] Asteman H, Svensson J-E, Johansson L-G. Oxidation of 310 steel in H2O/O2 mixtures at 600 °C: the effect
of water-vapour-enhanced chromium evaporation. Corros Sci 2002;44:2635–49.
[107] Jacobson N, Myers D, Opila E, Copland E. Interactions of water vapor with oxides at elevated temperatures.
J Phys Chem Solids 2005;66:471–8.
[108] Takeda M, Onishi T, Nakakubo S, Fujimoto S. Physical Properties of Iron-Oxide Scales on Si-Containing
Steels at High Temperature. Mater Trans 2009;50:2242–6.
[109] Sabau AS, Wright IG. Influence of Oxide Growth and Metal Creep on Strain Development in the Steam-
Side Oxide in Boiler Tubes. Oxid Met 2010;73:467–92.

139
[110] Pint BA, Tortorelli PF, Wright IG. Effect of Cycle Frequency on High-Temperature Oxidation Behavior of
Alumina-Forming Alloys. Oxid Met 2002;58:73–101.
[111] Nadot, Mendez, Ranganathan, Beranger. Fatigue life assessment of nodular cast iron containing casting
defects. Fatigue Fract Eng Mater Struct 1999;22:289–300.
[112] Nadot Y. Influence of casting defects on the fatigue limit of nodular cast iron. Int J Fatigue 2004;26:311–9.
[113] Papis K, Tunzini S, Menk W. Cast Iron Alloys for Exhaust Applications, INTEMA - UNMdP-CONICET;
2014.
[114] Pint BA, Zhang Y. Performance of Al‐rich oxidation resistant coatings for Fe‐base alloys. Mater Corros
2011;62:549–60.
[115] Kemnitz P, Maier O, Klein R. Monotherm, a New Forged Steel Piston Design for Highly Loaded Diesel
Engines. SAE; 2000.
[116] Soave R, Leites JMM, Ferran G, Kuhn CW, Gegel GA. Development of a Fiber Reinforced Aluminum
Piston for Heavy Duty Diesel Engines, 1994.
[117] Maurizi M, Hrdina D. New MAHLE Steel Piston and Pin Coating System for Reduced TCO of CV Engines.
SAE; 2016.
[118] MAHLE GmbH, editor. Pistons and Engine Testing. 2nd ed. Stuttgart: Springer Vieweg; 2016.
[119] Davis JR, editor. Heat-Resistant Materials. ASM International; 1997.
[120] Leslie WC. Iron and its dilute substitutional solid solutions. Metall Mater Trans B 1972;3:5–26.
[121] Ban T, Bohnenkamp K, Engell H-J. The formation of protective films on iron-silicon alloys. Corros Sci
1979;19:283–93.
[122] –102.
[123] Svedung I, Vannerberg N-G. The influence of silicon on the oxidation properties of iron. Corros Sci
1974;14:391–9.
[124] Speich GR, Leslie WC. Tempering of steel. Metall Trans 1972;3:1043–54.
[125] Clarke AJ, Miller MK, Field RD, Coughlin DR, Gibbs PJ, Clarke KD, et al. Atomic and nanoscale chemical
and structural changes in quenched and tempered 4340 steel. Acta Mater 2014;77:17–27.
[126] Chen Y-C, Worden JA. Evaluation of Microalloyed Steel for Articulated Piston Applications in Heavy Duty
Diesel Engines. SAE; 2000.
[127] Korchynsky M, Paules JR. Microalloyed Forging Steels-A State of the Art Review, 1989.
[128] Parsons SA, Edmonds DV. Microstructure and mechanical properties of medium-carbon ferrite–pearlite
steel microalloyed with vanadium. Mater Sci Technol 1987;3:894–904.
[129] Baker TN. Processes, microstructure and properties of vanadium microalloyed steels. Mater Sci Technol
2009;25:1083–107.
[130] Baker TN. Microalloyed steels. Ironmak Steelmak 2016;43:264–307.
[131] Xie H, Du L-X, Hu J, Misra RDK. Microstructure and mechanical properties of a novel 1000 MPa grade
TMCP low carbon microalloyed steel with combination of high strength and excellent toughness. Mater Sci
Eng A 2014;612:123–30.
[132] Hui W, Chen S, Zhang Y, Shao C, Dong H. Effect of vanadium on the high-cycle fatigue fracture properties
of medium-carbon microalloyed steel for fracture splitting connecting rod. Mater Des 1980-2015
2015;66:227–34.
[133] Nöhrer M, Zamberger S, Primig S, Leitner H. Atom probe study of vanadium interphase precipitates and
randomly distributed vanadium precipitates in ferrite. Micron 2013;54–55:57–64.
[134] Matlock DK, Speer JG. Microalloying concepts and application in long products. Mater Sci Technol
2009;25:1118–25.
[135] Hui W, Zhang Y, Shao C, Chen S, Zhao X, Dong H. Microstructural effects on high-cycle fatigue properties
of microalloyed medium carbon steel 38MnVS. Mater Sci Eng A 2015;640:147–53.
[136] José M, José A. Metallurgical Stability of Articulated Piston Under Severe Engine Operating Condition.
1993.
[137] Silva FS. Fatigue on engine pistons – A compendium of case studies. Eng Fail Anal 2006;13:480–92.
[138] Elevated-Temperature Properties of Ferritic Steels. Met. Handb. Prop. Sel. Irons Steels High-Perform.
Alloys, vol. 1. Tenth Edition, ASM International; n.d., p. 617–52.
[139] Kamo R. Adiabatic diesel-engine technology in future transportation. Energy 1987;12:1073–80.
[140] Bryzik W, Kamo R. TACOM/Cummins Adiabatic Engine Program, 1983.
[141] Ye C, Suslov S, Kim BJ, Stach EA, Cheng GJ. Fatigue performance improvement in AISI 4140 steel by
dynamic strain aging and dynamic precipitation during warm laser shock peening. Acta Mater
2011;59:1014–25.

140
[142] Ribeiro C, LU O, Egerer T. Piston made using additive manufacturing techniques. WO2014165734 A1,
2014.
[143] Wong VW, Bauer W, Kamo R, Bryzik W, Reid M. Assessment of Thin Thermal Barrier Coatings for I.C.
Engines. vol. 950980, 1995.
[144] Kamo R, Woods ME, Bryzik W. Thin thermal barrier coating for engines. US4852542A, 1989.
[145] Kosaka H, Wakisaka Y, Nomura Y, Hotta Y, Koike M, Nakakita K, et al. Concept of “Temperature Swing
Heat Insulation” in Combustion Chamber Walls, and Appropriate Thermo-Physical Properties for Heat
Insulation Coat. SAE Int J Engines 2013;6:142–9.
[146] Wakisaka Y, Inayoshi M, Fukui K, Kosaka H, Hotta Y, Kawaguchi A, et al. Reduction of Heat Loss and
Improvement of Thermal Efficiency by Application of “Temperature Swing” Insulation to Direct-Injection
Diesel Engines. SAE Int J Engines 2016;9.
[147] Wang H, Dinwiddie RB, Porter WD. Development of a Thermal Transport Database for Air Plasma Sprayed
ZrO<Subscript>2</Subscript>-Y<Subscript>2</Subscript>O<Subscript>3</Subscript> Thermal Barrier
Coatings. J Therm Spray Technol 2010;19:879–83.
[148] Lineton WB, Azevedo M. Piston with a cooling gallery partially filled with a thermally conductive metal-
containing composition. US9127619B2, 2015.
[149] Engineering Properties of alloy 713C. Nickel Institute; n.d.
[150] Matysiak H, Zagorska M, Balkowiec A, Adamczyk-Cieslak B, Cygan R, Cwajna J, et al. The Microstructure
Degradation of the IN 713C Nickel-Based Superalloy After the Stress Rupture Tests. J Mater Eng Perform
2014;23:3305–13.
[151] Baldan R, Silva AAAP da, Nunes CA, Couto AA, Gabriel SB, Alkmin LB. Solution and Aging of MAR-
M246 Nickel-Based Superalloy. J Mater Eng Perform 2017;26:465–71.
[152] Engineering Properties of In-100 Alloy. International Nickel; n.d.
[153] Tetsui T. Development of a TiAl turbocharger for passenger vehicles. Mater Sci Eng A 2002;329–331:582–
8.
[154] Ges A, Palacio H, Versaci R. IN-713C characteristic properties optimized through different heat treatments.
J Mater Sci 1994;29:3572–6. doi:10.1007/BF00352065.
[155] Ro Y, Koizumi Y, Harada H. High temperature tensile properties of a series of nickel-base superalloys on a
γ/γ′ tie line. Mater Sci Eng A 1997;223:59–63.
[156] Harada H, Yamazaki M, Koizumi Y. A Series of Nickel-base Superalloys on γ-γ′ Tie Line of Alloy Inconel
713C. Tetsu--Hagane 1979;65:1049–58.
[157] Azadi M, Azadi M. Evaluation of high-temperature creep behavior in Inconel-713C nickel-based superalloy
considering effects of stress levels. Mater Sci Eng A 2017;689:298–305.
[158] Azadi M, Marbout A, Safarloo S, Azadi M, Shariat M, Rizi MH. Effects of solutioning and ageing
treatments on properties of Inconel-713C nickel-based superalloy under creep loading. Mater Sci Eng A
2018;711:195–204. doi:10.1016/j.msea.2017.11.038.
[159] Kunz L, Lukáš P, Konečná R. Initiation and propagation of fatigue cracks in cast IN 713LC superalloy. Eng
Fract Mech 2010;77:2008–15.
[160] Kunz L, Lukáš P, Konečná R. High-cycle fatigue of Ni-base superalloy Inconel 713LC. Int J Fatigue
2010;32:908–13.
[161] Kunz L, Lukáš P, Konečná R, Fintová S. Casting defects and high temperature fatigue life of IN 713LC
superalloy. Int J Fatigue 2012;41:47–51.
[162] Obrtlík K, Pospíšilová S, Juliš M, Podrábský T, Polák J. Fatigue behavior of coated and uncoated cast
Inconel 713LC at 800 °C. Int J Fatigue 2012;41:101–6.
[163] Petrenec M, Obrtlík K, Polák J. Inhomogeneous dislocation structure in fatigued INCONEL 713 LC
superalloy at room and elevated temperatures. Mater Sci Eng A 2005;400–401:485–8.
[164] Pint BA, Haynes JA, Armstrong BL. Performance of advanced turbocharger alloys and coatings at 850–950
°C in air with water vapor. Surf Coat Technol 2013;215:90–5.
[165] Yoshihara M, Kim Y-W. Oxidation behavior of gamma alloys designed for high temperature applications.
Intermetallics 2005;13:952–8.
[166] Huang S-C. Titanium aluminum alloys modified by chromium and niobium and method of preparation.
US4879092A, 1989.
[167] Austin CM, Kelly TJ, Huang S-C. Titanium aluminide alloy with improved temperature capability.
US5545265A, 1996.
[168] Weimer MJ, Bewlay BP, Jr MFXG, Kelly TJ. Titanium aluminide intermetallic compositions.
US20130251537A1, 2013.

141
[169] Tetsui T, Ono S. Endurance and composition and microstructure effects on endurance of TiAl used in
turbochargers. Intermetallics 1999;7:689–97.
[170] Appel F, Brossmann U, Christoph U, Eggert S, Janschek P, Lorenz U, et al. Recent Progress in the
Development of Gamma Titanium Aluminide Alloys. Adv Eng Mater 2000;2:699–720.
[171] Wu X. Review of alloy and process development of TiAl alloys. Intermetallics 2006;14:1114–22.
[172] Umakoshi Y, Nakano T, Yamane T. The effect of orientation and lamellar structure on the plastic behavior
of TiAl crystals. Mater Sci Eng A 1992;152:81–8.
[173] Kumpfert J, Kim YW, Dimiduk DM. Effect of microstructure on fatigue and tensile properties of the gamma
TiAl alloy Ti-46.5Al-3.0Nb-2.1Cr-0.2W. Mater Sci Eng A 1995;192–193:465–73.
[174] Hu D, Huang A, Jiang H, Mota-Solis N, Wu X. Pre-yielding and pre-yield cracking in TiAl-based alloys.
Intermetallics 2006;14:82–90.
[175] M. Imayev R, Imayev V, Oehring M, Appel F. Alloy design concepts for refined gamma titanium aluminide
based alloys. Intermetallics 2007;15:451–60.
[176] Pflumm R, Friedle S, Schütze M. Oxidation protection of γ-TiAl-based alloys – A review. Intermetallics
2015;56:1–14.
[177] Fergus JW. Review of the effect of alloy composition on the growth rates of scales formed during oxidation
of gamma titanium aluminide alloys. Mater Sci Eng A 2002;338:108–25.
[178] Kim BG, Kim GM, Kim CJ. Oxidation behavior of TiAl-X (X = Cr, V, Si, Mo or Nb) intermetallics at
elevated temperature. Scr Metall Mater 1995;33:1117–25.
[179] Oehring M, Appel F, Ennis PJ, Wagner R. A TEM study of deformation processes and microstructural
changes during long-term tension creep of a two-phase γ-titanium aluminide alloy. Intermetallics
1999;7:335–45.
[180] Kraft EH. Opportunities for Low Cost Titanium in Reduced Fuel Consumption, Improved Emissions, and
Enhanced Durability Heavy Duty Vehicles. EHKTechnologies; 2002. doi:10.2172/814648.
[181] Tetsui T, Kobayashi T, Ueno T, Harada H. Consideration of the influence of contamination from oxide
crucibles on TiAl cast material, and the possibility of achieving low-purity TiAl precision cast turbine
wheels. Intermetallics 2012;Complete:274–81.
[182] Tetsui T, Kobayashi T, Mori T, Kishimoto T, Harada H. Evaluation of Yttria Applicability as a Crucible for
Induction Melting of TiAl Alloy. Mater Trans 2010;51:1656–62.
[183] Murr L, Gaytan S, Ceylan A, Martinez E, Martinez JL, Hernandez D, et al. Characterization of titanium
aluminide alloy components fabricated by additive manufacturing using electron beam melting. Acta Mater
2010;58:1887–94.
[184] Biamino S, Penna A, Ackelid U, Sabbadini S, Tassa O, Fino P, et al. Electron beam melting of Ti–48Al–
2Cr–2Nb alloy: Microstructure and mechanical properties investigation. Intermetallics 2011;19:776–81.
[185] Blau PJ. A Wear Model for Diesel Engine Exhaust Valves. Oak Ridge National Laboratory; 2009.
[186] Engine Poppet Valve Information Report. SAE International; 2004.
[187] Schaefer SK, Larson JM, Jenkins LF, Wang Y. Evolution of Heavy Duty Engine Valves - Materials and
Design, 1997, p. 129–39.
[188] Larson JM, Jenkins LF, Narasimhan SL, Belmore JE. Engine Valves—Design and Material Evolution. J Eng
Gas Turbines Power 1987;109:355–61.
[189] Tanaka N, Kawata A. Measurement Technique of Exhaust Valve Temperature, 2015.
[190] Ari-Gur P, Noble V, Narasimhan S, Larson JM. Hot Corrosion Studies of Automotive Exhaust Valves. SAE;
1991.
[191] Farina AB, Liberto RCN, Barbosa CA. Development of New Intermediate Nickel Alloys for Application in
Automotive Valves of High Performance Engines. SAE; 2013.
[192] Sato K, Saka T, Ohno T, Kageyama K, Sato K, Noda T, et al. Development of Low-Nickel Superalloys for
Exhaust Valves, 1998.
[193] Muralidharan G. Low-cost Fe—Ni—Cr alloys for high temperature valve applications. US9605565 B2,
2017.
[194] Liu R, Yao JH, Zhang QL, Yao MX, Collier R. Microstructures and Hardness/Wear Performance of High-
Carbon Stellite Alloys Containing Molybdenum. Metall Mater Trans A 2015;46:5504–13.
[195] Sawant MS, Jain NK. Investigations on wear characteristics of Stellite coating by micro-plasma transferred
arc powder deposition process. Wear 2017;378–379:155–64.
[196] Liu R, Wu XJ, Kapoor S, Yao MX, Collier R. Effects of Temperature on the Hardness and Wear Resistance
of High-Tungsten Stellite Alloys. Metall Mater Trans A 2015;46:587–99.
[197] Jennings PA. High-Temperature Steel and Articles. 2657130, 1953.

142
[198] Jenkins LF, Larson JM. The Development of a New Austenitic Stainless Steel Exhaust Valve Material,
1978.
[199] Byrnes MLG, Grujicic M, Owen WS. Nitrogen strengthening of a stable austenitic stainless steel. Acta
Metall 1987;35:1853–62.
[200] Mathew MD, Laha K, Ganesan V. Improving creep strength of 316L stainless steel by alloying with
nitrogen. Mater Sci Eng A 2012;535:76–83.
[201] Lai JKL. A review of precipitation behaviour in AISI type 316 stainless steel. Mater Sci Eng 1983;61:101–9.
[202] McGuire MF. Austenitic Stainless Steels. Stainl. Steels Des. Eng. - ASM Int., 2008, p. 69–90.
[203] Oda K, Kondo N, Shibata K. X-ray Absorption Fine Structure Analysis of Interstitial (C, N)-Substitutional
(Cr) Complexes in Austenitic Stainless Steels. ISIJ Int 1990;30:625–31.
[204] Nilsson J 0, Thorvaldsson T. The influence of nitrogen on microstructure and strength of a high-alloy
austenitic stainless steel. Scand J Met 1985;15:83–9.
[205] Sourmail T. Precipitation in creep resistant austenitic stainless steels. Mater Sci Technol 2001;17:1–14.
[206] Erneman J, Schwind M, Andren H, Nilsson J, Wilson A, Agren J. The evolution of primary and secondary
niobium carbonitrides in AISI 347 stainless steel during manufacturing and long-term ageing. Acta Mater
2005.
[207] Ghosh S. The role of tungsten in the coarsening behaviour of M23C6 carbide in 9Cr–W steels at 600 °C. J
Mater Sci 2010;45:1823–9.
[208] Jaswin MA, Lal DM, Rajadurai A. Effect of Cryogenic Treatment on the Microstructure and Wear
Resistance of X45Cr9Si3 and X53Cr22Mn9Ni4N Valve Steels. Tribol Trans 2011;54:341–50.
[209] Jaswin MA, Lal DM. Effect of cryogenic treatment on the tensile behaviour of En 52 and 21-4N valve steels
at room and elevated temperatures. Mater Des 2011;32:2429–37.
[210] Muzyka DR, Whitney CR. Sulfidation resistant nickel-iron base alloy. US3972713 A, 1976.
[211] Evans ND, Maziasz PJ, Truhan JJ. Phase transformations during service aging of Nickel based superalloy
Pyromet 31V. Solid-Solid Phase Transform Inorg Mater 05 2005:809–817.
[212] Sawford MK, Sinharoy S, Narasimhan S, Kajinic A, Wojcieszynski AL, Wright JK. Wear resistant high
temperature alloy. EP2038444 B1, 2015.
[213] Kato T, Uyehara N, Matsunaga K, Isomura T, Matsuno M, lizuka M. A New Iron-Base Superalloy for
Exhaust Valves. SAE 1981;810032.
[214] Muralidharan G. Low-Cost Fe-Ni-Cr Alloys for High Temperature Valve Applications. US20150368760
A1, 2015.
[215] Barbosa CA, Jarreta DD, Sokolowski A. Nickel-based superalloy for valves of internal combustion engines.
WO2011029164 A1, 2011.
[216] Barbosa CA, Jarreta DD, Sokolowski A. Alloys with high corrosion resistance for engine valve applications.
WO2011029165 A1, 2011.
[217] Narasimhan SL, Larson JM, Whelan EP. Wear characterization of new nickel-base alloys for internal
combustion engine valve seat applications. Wear 1981;74:213–27.
[218] Narasimhan SL, Larson JM. Valve Gear Wear and Materials. SAE 1985.
[219] Wang YS, Narasimhan S, Larson JM, Schaefer SK. Wear and Wear Mechanism Simulation of Heavy-Duty
Engine Intake Valve and Seat Inserts. J Mater Eng Perform 1997;7:53–65.
[220] Zhao R, Barber GC, Wang YS, Larson JE. Wear Mechanism Analysis of Engine Exhaust Valve Seats with a
Laboratory Simulator. Tribol Trans 1997;40:209–18.
[221] Lewis R, Dwyer-Joyce RS. Design Tools for Prediction of Valve Recession and Solving Valve/Seat Failure
Problems, SAE; 2001.
[222] Lewis R, Dwyer-Joyce RS. Wear of diesel engine inlet valves and seat inserts. Proc Inst Mech Eng Part J
Automob Eng 2002;216:205–16.
[223] Lewis R. Wear of Diesel Engine Inlet Valves and Seats. University of Sheffield; 2000.
[224] Forsberg P, Hollman P, Jacobson S. Wear mechanism study of exhaust valve system in modern heavy duty
combustion engines. Wear 2011;271:2477–84.
[225] Forsberg P, Hollman P, Jacobson S. Wear Study of Coated Heavy Duty Exhaust Valve Systems in a
Experimental Test Rig, 2012.
[226] Blau PJ, Brummett TM, Pint BA. Effects of prior surface damage on high-temperature oxidation of Fe-, Ni-,
and Co-based alloys. Wear 2009;267:380–6.
[227] Mrowec S. Mechanism of high-temperature sulphide corrosion of metals and alloys. Mater Corros
1980;31:371–86.
[228] Bornstein NS. Reviewing sulfidation corrosion—Yesterday and today. JOM 1996;48:37–9.

143
[229] Bornstein NS, Allen WP. The Chemistry of Sulfidation Corrosion-Revisited. Mater Sci Forum 1997;251–
254:127–34.
[230] Arnold EB, Bara MA, Zang DM, Tunnecliffe TN, Oltean J. Development and Application of a Cycle for
Evaluating Factors Contributing to Diesel Engine Valve Guttering, 1988.
[231] Scott CG, Riga AT, Hong H. The erosion-corrosion of nickel-base diesel engine exhaust valves. Wear
1995;181:485–94.
[232] Witek L. Failure and thermo-mechanical stress analysis of the exhaust valve of diesel engine. Eng Fail Anal
2016;66:154–65.
[233] Yu ZW, Xu XL. Failure analysis and metallurgical investigation of diesel engine exhaust valves. Eng Fail
Anal 2006;13:673–82.
[234] Stott FH, Wood GC, Hobby MG. A comparison of the oxidation behavior of Fe-Cr-Al, Ni-Cr-Al, and Co-
Cr-Al alloys. Oxid Met 1971;3:103–13.
[235] Barnes DJ, Stott FH, Wood GC. The frictional behaviour of iron and iron-chromium alloys at elevated
temperatures. Wear 1977;45:199–209.
[236] Stott FH, Wood GC. The influence of oxides on the friction and wear of alloys. Tribol Int 1978:211–8.
[237] Wang YS, Narasimhan S, Larson JM, Larson JE, Barber GC. The effect of operating conditions on heavy
duty engine valve seat wear. Wear 1996;201:15–25.
[238] Stott FH. The role of oxidation in the wear of alloys. Tribol Int 1998;31:61–71.
[239] Kato T, Fujikura M, Likubo T, Isomura T, Matsuno M, Hirayama H. Application of a Newly Developed
Iron-Base Superalloy to Exhaust Valves of Diesel Engines 1983;830253.
[240] Kwon OG, Han MS. Failure analysis of the exhaust valve stem from a Waukesha P9390 GSI gas engine.
Eng Fail Anal 2004;11:439–47.
[241] Pint BA, Maziasz PJ, Schauer J, Levin V. Protective Aluminide Coatings for NiCr Alloys in Combustion
Environments, Oak Ridge National Laboratory (ORNL); 2008.
[242] Sinharoy S, Narasimhan SL. Oxidation behavior of two nickel-base superalloys used as elevated temperature
valves in spark ignited engines and diesel exhaust recirculation (EGR) applications, The Minerals, Metals &
Materials Society; 2004.
[243] Pint BA, Dryepondt SN, Unocic KA. Oxidation of Superalloys in Extreme Environments, The Minerals,
Metals & Materials Society; 2010.
[244] Mrowec S. The problem of sulfur in high-temperature corrosion. Oxid Met 1995;44:177–209.
[245] Carter RV, Douglass DL, Gesmundo F. Kinetics and mechanism of the sulfidation of Fe-Mo alloys. Oxid
Met 1989;31:341–67.
[246] Wang G, Carter R, Douglass DL. High-temperature sulfidation of Fe-Nb alloys. Oxid Met 1989;32:273–94.
[247] Shing C, Douglass DL, Gesmundo F. Sulfidation behavior of Al-modified MP35N alloys. Oxid Met
1991;35:295–315.
[248] Chen MF, Douglass DL. The effect of Mo on the high-temperature sulfidation of Ni. Oxid Met
1989;32:185–206.
[249] Chen MF, Douglass DL, Gesmundo F. High-temperature sulfidation behavior of Ni-Nb alloys. Oxid Met
1989;31:237–63.
[250] Gleeson B, Douglass DL, Gesmundo F. Effect of Nb on the high-temperature sulfidation behavior of cobalt.
Oxid Met 1989;31:209–36.
[251] Wang G, Douglass DL, Gesmundo F. Effect of Al on the high-temperature sulfidation of Fe-30Nb. Oxid
Met 1991;35:279–94.
[252] Wang G, Douglass DL, Gesmundo F. High-temperature sulfidation of Fe-30Mo alloys containing ternary
additions of Al. Oxid Met 1991;35:349–73.
[253] Mrowec S, Werber T, Zastawnik M. The mechanism of high temperature sulphur corrosion of nickel-
chromium alloys. Corros Sci 1966;6:47–68.
[254] Evans ND, Yamamoto Y, Maziasz PJ, Shingledecker JP. Age Induced Gamma Prime Coarsening and
Hardness Behavior in Pyromet 31V. Microsc Microanal 2006;12:1044–5.
[255] Baiamonte L, Marra F, Gazzola S, Giovanetto P, Bartuli C, Valente T, et al. Thermal sprayed coatings for
hot corrosion protection of exhaust valves in naval diesel engines. Surf Coat Technol 2016;295:78–87.
[256] Slatter T, Taylor H, Lewis R, King P. The influence of laser hardening on wear in the valve and valve seat
contact. Wear 2009;267:797–806.

144
Table 1 – Summary of typical materials currently used in HDDE components along with their mechanical loading and peak
temperature limits, if available. The temperature limit corresponds to a peak temperature occurring under high load conditions
corresponds to high loading and longevity requirements for a particular component in typical HDDE applications. Major life
limiting mechanisms are also summarized for each component/material combination based on typical operating conditions in
production HDDE. The values and mechanisms in this table will change depending on numerous operational parameters.
Mechanical Loading
Component Typical Material in Current Use Temperature Limitation Major life limiting mechanisms
Limitation
a Thermal fatigue [50,69]
Cylinder heads Alloyed gray cast iron ~25 MPa PCP [3] ~400 °C [3]
HCF [3,69]
HCF [126]
Martensitic 4140H steel - ~500 °Ca [118,126]
Oxidation [118]
Pistons
HCF [126]
Microalloyed steel 38MnVS6, ~450 °Ca [118,126]
Oxidation [118]
HCF
Seat wear [219]
~538 °Cb – 590 °Ca Service aging
Silchrome 1 -
[186,187] Oxidation
Sulfur based acid corrosion [25]
Guttering/radial cracking [230,231]
HCF
Seat wear [219]
Chromo 193 - ~593 °Ca [186,187]
Sulfur based acid corrosion [25]
Intake Valves
Guttering/radial cracking [230,231]
HCF
Representative intermediate Ni Seat wear [219]
- N/Ac
content alloy Crutonite® Sulfur based acid corrosion [25]
Guttering/radial cracking [230,231]
HCF
Ni based super-alloys (e.g., Pyromet Seat wear [219]
- N/Ac
31V, Inconel 751, etc.) Sulfur based acid corrosion [25]
Guttering/radial cracking [230,231]
CTF [86]
Exhaust manifolds and Creep [90]
High SiMo ductile cast iron - ~760 °C
turbocharger housings HCF [101]
Oxidation [83,84]
HCF
Precipitation strengthened austenitic
- ~760-788 °C [212] a Seat wear [220]
steel (21-4N mod., 23-8N)
Oxidation
HCF
Representative intermediate Ni Seat wear [212]
- ~815 °Ca
content alloy Crutonite® Oxidation
Exhaust valves
Guttering [231]
HCF
Oxidation
Ni based super-alloys (e.g., Pyromet
- ~870-900 °C [186] a Seat Wear
31V, Inconel 751, etc.)
Sulfidation [210]
Guttering [231]
a
Engine PCP will influence component temperature limitations
b
Approximate temperature limit for Silchrome 1 intake valves in some HDDE applications as set forth by Eaton Co.
c
Intermediate Ni content and Ni based superalloys are generally not temperature limited in intake valve applications

145
Table 2 – HDDE metrics and targets for engine weight reduction, power density, brake thermal efficiency, and peak cylinder
pressures for a baseline 354 kW (475 HP) engine [17]. The estimated peak metal temperatures at the exhaust valve and turbo inlet
under high loads in these notional engines are provided.
2010
(overall 2025 2030 2040 2050
baseline)
Weight Reduction - 15% lighter 15% lighter 18% lighter 20% lighter
Displacement 15L 11L 10.5L 10L 9L

27% augmented 30% augmented 33% augmented 40% augmented


Power density 24 kW L-1
32 kW L-1 34 kW L-1 36 kW L-1 40 kW L-1
Fossil Fuel (32 HP L-1)
(43 HP L-1) (45 HP L-1) (48 HP L-1) (53 HP L-1)

Brake Thermal
Efficiency 42% 50% 50% 55% 60%
(Percent)
Peak Cylinder
19 MPa 25 MPa 25 MPa 27.5 MPa 30 MPa
Pressures
Peak Metal
Temperatures on
700 °C 800 °C 850 °C 875 °C 900 °C
Exhaust Valve and
Turbo Inlet

146
Table 3 – Chemical compositions of gray and compacted graphite iron from various studies along with the RT value of UTS
reported in the study.

Type/ RT
Ref. C Si Mn Cr Mo Cu Mg Ni S P Sn Ti
Designation UTS

Gray [62] 250a 3.43 2.07 0.55 0.20 1.00 0.10

Gray [51] 267 3.43 1.65 0.57 0.11 0.004 0.11 0.11 0.080 0.030

Gray [62] 300a 3.3 2.05 0.56 0.24 0.30 1.2 0.11

Gray [51] 302 3.45 1.74 0.59 0.49 0.004 0.59 0.60 0.082 0.03

Gray [51] 306 3.44 1.69 0.58 0.21 0.38 0.30 0.10 0.082 0.027 0.077

Gray [53] 313 3.31 1.76 0.73 0.64 0.12 0.07 0.17 0.072 0.046

Gray [51] 324 3.45 1.68 0.63 0.30 0.30 0.87 0.97 0.081 0.028

Gray [51] 331 3.43 1.66 0.58 0.5 0.39 0.12 0.10 0.083 0.030

Gray [53] 384 3.21 2.06 0.68 0.96 0.07 0.17 0.072 0.046

CGI [62] 350a 3.65 2.45 0.37 0.029 0.41 0.031

CGI [55] 365 3.78 2.39 0.37 0.01 0.01 0.023 0.08 0.012 0.09

CGI [69] 416 3.38 1.90 0.374 <0.010 0.97 <0.050 0.01 0.019 0.090 0.011
a
CGI [62] 450 3.62 2.41 0.37 0.029 1.17 0.064

CGI [61] 470 3.5 2.18 0.56 0.05 0.46 0.58 0.011 0.03 0.012 0.016 0.140
3.60- 2.00- 0.15- 0.75- 0.06-
CGI [41] 520 <0.10 <0.015
3.90 2.20 0.40 0.95 0.10
a
Nominal value

147
Table 4 – Chemical compositions of common alloys that are used in exhaust manifolds and turbocharger housings in HDDE or
are under consideration for these applications. The compositions are actual compositions reported by various studies which are
discussed in this review [85,90], [86], [101], [83]. Since these are actual compositions reported from a single study, the actual
compositions of these materials may differ from study to study for the same material (nominally). Materials are ranked based on
their elevated temperature mechanical properties and corrosion resistance .
Material Ranking for Specific Tests (1=best)

Short
Short Term
Designation

Long
Term Ox.
Ref. Structure C Si Mn Cr Ni Nb W Co Mo Mg N Ti Al V B Short Term
Ox. Resist
Time LCF Ox.
CTF TMF Creep Resist in CTE
YS, 700°C Resist.,
[86] [101] [90,93,99] in air water [85,93,99]
800 °C [85] water
[93,101] vapor,
[85,90] vapor
700 °C, 900
[83]
750 °C °C,
[84]

Hi SiMo [90] BCC 3.45 4.0 0.3 0.6 5 6

SiMo 5.1 [85] BCC 3.17 4.15 0.4 0.10 0.04 0.86 0.052 6 3 2 6 3 2

SiMo1000 [101] BCC 3.57 2.72 0.25 0.84 0.81 3.08 1 5 1 2

1.4509 [85] BCC 0.02 0.81 0.75 17.9 0.12 0.36 <0.01 0.02 6 1 1

Ni resist D5S [85] FCC 2.41 5.38 0.28 1.77 33.12 0.056 5 4 4 6 2 2 3

HF [85] FCC 0.33 1.5 1.43 18.76 8.87 0.01 0.01 0.02 4 2 3 6

A3N [85] FCC 0.57 0.58 1.05 19.59 10.15 1.75 2.28 0.29 1 4 6

CF8C [90] FCC 0.07 1.0 <1 19.0 10.0 0.8 - 0.3 3

CF8C-Plus [90] FCC 0.1 0.5 4.0 19.0 12.5 0.8 - 0.3 0.25 3 1 1 5 6

CN-12 Plus [97] FCC 0.41 0.72 4.26 24.64 16.06 1.50 0.02 0.09 0.27 0.40 0.01 0.07 <0.001

HK30 [85] FCC 0.51 1.5 1.45 25.41 18.67 <0.005 0.01 0.03 4 1 1 4

HK-Nb [85] FCC 0.37 1.29 0.96 25.38 22.4 1.39 0.02 0.03 2 2 4 1 5

TMA 4705 [83] FCC 0.7 1.5 0.66 25.9 20.8 0.38 0.28 1.16 0.18 0.04 0.04 3 3

CAFA 4 [83] FCC 0.29 0.48 1.92 11.1 25.3 0.94 1.0 1.98 0.05 3.49 0.05 1

HP [83] FCC 0.42 1.61 0.63 25.5 34.2 1.14 0.24 0.1 0.04 0.06 0.04 2

TMA 6301 [83] FCC 0.44 1.23 0.96 25.8 34.2 0.43 0.42 0.07 0.06 0.04 2 2

148
Table 5 – Chemical composition of 4140H and 38MnSiVS5 steels [118]
C Si Mn Cr Mo P S V Ti N
4140H 0.38-0.45 max. 0.40 0.60-0.90 0.90-1.20 0.15-0.30 max. 0.035 max. 0.035 - -
38MnSiVS5 0.34-0.41 0.15-0.80 1.2-1.6 max. 0.30 max. 0.08 max. 0.025 0.020-0.060 0.08-0.20 0.01 0.01-0.02

149
Table 6 – Nominal chemical composition of materials used in HDDE turbine wheels, Inconel 713C and Inconel 713LC, as well
as candidate TiAl alloys.

Designation Ref. Structure C Si Mn Cr Ni Nb W Co Mo Mg N Ti Al V B

Inconel 713C [149] FCC 0.096 0.010 13.6 Bal. 2.16 4.21 0.84 6.08 0.01

Inconel 713LC [159] FCC 0.007 0.06 11.63 Bal. 2.36 0.30 4.66 0.77 5.91 0.12
Multi-
TiAl [169] 0.6 0.4 0.4 Bal. 47.3
phase
Multi-
TiAl [169] 1 7.1 Bal. 45.0
phase

150
Table 7 – HDDE valve, hardfacing, and insert materials and their chemical composition (adapted from [187,188,190,191])
Valve Material Chemical Composition in wt.%

Alloy
C Mn Si Cr Ni Fe Other

Intake Valve Materials


Silchrome 1 0.45 0.40 3.20 8.5 0.40 Bal.
8645 0.45 0.82 0.25 0.5 0.55 Bal. Mo: 0.2
Chromo 193 0.85 1.5 max 1.0 max 17.5 - Bal. Mo: 2.25V: 0.45
Exhaust Valve Materials
21-4N 0.51 8.75 0.15 21.0 3.80 Bal. N:0.45
mod. 21-4N 0.51 9.0 0.26 21.0 4.25 Bal. Nb+Ta:2.2, W:1.2, N:0.5
23-8N 0.31 2.5 0.75 23 8 Bal. N:0.31
Silchrome 10 Steel (Cr-Ni-Si-C) 0.40 1.10 3.00 19.0 8 Bal.
21-12N 0.20 1.30 0.50 21 11.5 Bal. N:0.2
NCF3015A 0.03 0.10 0.10 14 32 Bal. Mo: 0.7, Al: 1.9, Ti: 2.6, Nb: 0.6, B: 0.004
NCF4015A 0.03 0.10 0.10 15 41 Bal. Mo: 0.7, Al: 1.9, Ti: 2.3, Nb: 1.3, B: 0.004
VAT 32 0.3 0.3 0.2 15.5 32.0 Bal. Nb: 3.90, Al: 1.80, Ti: 2.00
Nb:1.8-2.5, Al:1.63-2.3, Ti:2.0-3.5,
Crutonite® 0.16-0.30 <1 0.15 15-20 30-35 Bal.
Mo+W: 0-0.5%
VAT 36 0.05 0.3 0.2 19 36 Bal. Nb: 2.00, Al: 1.95, Ti: 1.2
Al: 1.75, Ti: 3.75, Co: 1.0, W: 0.01,
161 0.025 0.03 0.020 18 47 Bal.
Mo:1.2, Zr: 0.02
Pyromet 31V 0.04 0.1 0.1 22.5 57.0 Bal. Ti: 2.25. Al: 1.25. Mo:2.0
Nimonic 80A 0.1 1.0 0.25 20.5 68.5 2.0 Ti: 2.3, Al:1.4, Co:2.0
Inconel 751 0.1 1.0 0.25 15.5 73.9 7.0 Ti: 2.3, Al: 1.20, Nb: 0.95
Valve seat hardfacing materials
Stellite 1 2.80 0.5 1.10 30.0 3.0 3.00 Co: bal., W: 13
Stellite 6 1.15 0.3 1.05 28.0 1.50 1.50 Co: bal., W: 4.5, Mo: 1.0
Stellite F 1.75 0.5 1.1 25.5 22.5 <1.5 Co: bal., Mo: 0.6, W: 12
Eatonite 3 2.0 0.5 1.2 29.0 Bal. 8.0 Mo: 5.5
Eatonite 5 2 1.00 1.50 29.0 Bal 4.5 Mo:8.0, N:0.1
Eatonite 6 2 1.00 1.50 28.0 10.0 Bal. Mo:5.0
Triballoy 400 0.03 0.02 2.41 8.6 0.6 0.7 Co: Bal., Mo: 28.2
Insert materials
Silchrome XB 1.5 0.40 2.15 20.0 1.30 Bal.

151

You might also like