Sie sind auf Seite 1von 48

262 Int. J. Microstructure and Materials Properties, Vol. 2, Nos.

3/4, 2007

Overview of factors contributing to steel spring


performance and failure

Lauralice C.F. Canale and


Renata Neves Penha
Departamento de Engenharia de Materiais
Aeronáutica e Automobilística
Escola de Engenharia de São Carlos
Universidade de São Paulo
Av. Trabalhador São-carlense 400
São Carlos, SP 13566–590, Brazil
E-mail: lfcanale@sc.usp.br
E-mail: rnpenha@yahoo.com.br

George E. Totten*
Department of Mechanical and Materials Engineering
Portland State University
P.O. Box 751, Portland, OR 97207–0751, USA
E-mail: GETotten@aol.com
*Corresponding author

Antonio C. Canale
Departamento de Engenharia de Materiais
Aeronáutica e Automobilística
Escola de Engenharia de São Carlos
Universidade de São Paulo
Av. Trabalhador São-carlense 400
São Carlos, SP 13566–590, Brazil
E-mail: canale@sc.usp.br

Maria Regina Gasparini


Rassini NHK Auto-Peças Ltda
São Bernardo do Campo, SP, Brazil
E-mail: mrgasparini@rassini-nhk.com.br

Abstract: Leaf springs and coil springs are commonly used in the motor
vehicle industry. Premature fatigue failure is a common failure mode.
However, the reasons for these failures may be more complex and include:
design deficiencies, heat treatment, steel alloy chemistry, intergranular
cracking, presence of Fe-S inclusions, and grain boundary embrittlement. This
paper will provide an overview of spring steel: chemistry, heat treatment,
residual stress, fatigue failure, and failure analysis of leaf and coil springs.

Copyright © 2007 Inderscience Enterprises Ltd.


Overview of factors contributing to steel spring performance and failure 263

Keywords: spring; spring steel; chemistry; heat treatment; residual stress;


fatigue failure; failure analysis.

Reference to this paper should be made as follows: Canale, L.C.F.,


Penha, R.N., Totten, G.E., Canale, A.C. and Gasparini, M.R. (2007) ‘Overview
of factors contributing to steel spring performance and failure’, Int. J.
Microstructure and Materials Properties, Vol. 2, Nos. 3/4, pp.262–309.

Biographical notes: Lauralice C.F. Canale is a Mechanical Engineer and


received her MS and PhD degrees from Escola de Engenharia de São Carlos,
Universidade de São Paulo (USP). She has coauthored chapters and papers
on heat treating, quenching, tribology and biomaterials. Dr. Canale is an
Associate Professor in the same university (USP), responsible for graduate and
undergraduate courses.

Renata Neves Penha is Production Engineer and recently has received her MS
degree from the from Escola de Engenharia de São Carlos, Universidade de
São Paulo (USP). Actually, she is working in her PhD programme at the
same university.

George E. Totten received his BS and MS degrees from Fairleigh Dickinson


University in New Jersey and his PhD degree from New York University. He
has coauthored approximately 25 books and has approximately 500 patents and
publications on various aspects of heat treating, quenching, hydraulic
lubrication and tribology. He is the past President of the International
Federation for Heat Treatment and Surface Engineering (IFHTSE). Dr. Totten
is a Fellow of IFHTSE, ASTM, SAE and ASM International. Currently, he is a
Research Professor at Portland State University, in Portland, OR, USA.

Antonio C. Canale is a Mechanical Engineer and received his MS and PhD


degrees from Escola de Engenharia de São Carlos, Universidade de São Paulo
(USP). He has coauthored papers on automotive and aeronautic areas, mainly
in simulation. Actually, Dr. Canale is Professor at the Departamento de Eng de
Materiais, Aeronáutica e Automobilística of USP.

Maria Regina Gasparini is a Plant Manager of the Rassini NHK Autopeças


Ltda in Brazil.

1 Introduction

Kern and Suess (1979) have defined a spring as “… An elastic body with a primary
function of deflecting or distorting under applied load and returning to its original shape
when the load is removed”. In other words, a spring is a component that is capable of
storing energy, either temporarily or permanently. Because of cyclic loading that
accompanies the use of springs, there is an ongoing interest in improving fatigue strength
of spring materials. Spring efficiency of a is related to its ability to store energy per unit
weight. Typically, steel strengths of >1379 MPa are desired (Kern and Suess, 1979).
There has been a continuing effort to develop spring steels to meet ever-increasing
demands for improved mechanical properties with lower weight suspension materials to
facilitate the larger effort of developing automotive vehicles with lower weights and
lower cost (Murai and Takenoshita, 1989; Nam et al., 2000). For development of
264 L.C.F. Canale, R.N. Penha, G.E. Totten, A.C. Canale and M.R. Gasparini

high-strength spring steels with improved sag strength and fatigue strength and improved
quench embrittlement properties in addition to other thermophysical and mechanical
properties has been of part of this vitally important effort (Kawabe et al., 2002;
Rakhshtadt, 1975; Reguly et al., 2004). This research has focused on both coil and leaf
spring designs.
The two types of springs that will be discussed here are helical coil springs (Prawoto
et al., 2004) and leaf springs (Mukhopadhyay et al., 1997). Coil springs are the most
common springs used in the automotive industry and they are constructed from a length
of round steel wire that is formed into loops which allow for movement. Coil springs may
be further classified as compression and extension springs and they are illustrated
schematically in Figure 1.

Figure 1 Illustration of (a) a compression coil spring; (b) an extension coil spring; and
(c) a leaf spring.

(a)

(c)
(b)

A compression spring is used where the energy of movement acts to compress the
spring. These springs are used in applications such as shock absorber assemblies to
absorb the movement arising from bumps and vehicle weight transfer. Conversely, an
extension spring, such as would be used where the energy of movement will act to extend
the spring.
Coil springs are typically manufactured by one of two processes (Kern and
Suess, 1979):
1 wire with the desired mechanical properties is wound into a finished spring
2 soft wire of the desired composition and hardenability is wound into a spring
(sometimes by a hot-forming process), then quenched, tempered and finished into
a part.
A leaf spring, also occasionally referred to as a semi-elliptical spring or cart spring and
illustrated in Figure 1, is a long, flat, flexible piece of spring steel (or a composite
material which will not be discussed here) that deflects by bending when forces act on it.
Leaf springs are used primarily in suspensions to support the load of wheeled vehicles.
Leaf springs may be used singly or in several layers, often with each successive layer
being shorter than the one bellow it, which are bracketed together into one assembly.
Currently, leaf springs are predominantly used for heavy vehicles such as trucks because
a leaf spring is stresses along its length and transmits a load over the width of the chassis.
This action should be contrasted to a coil spring which transfer a load to a single point.
Leaf springs may perform a locating and to some extent damping functions as well as
springing functions.
Overview of factors contributing to steel spring performance and failure 265

Performance of a leaf spring is related to the spring rate and static deflection. The
spring rate is defined as the change of load per unit of deflection (N/mm). Static
deflection is the static load divided by the spring rate at the static load and this value
determines the stiffness of the suspension and the ride frequency of the vehicle. (The ride
frequency is defined as the low-frequency body motion, up to 5 Hz, when the vehicle is
driven over imperfect roads. Although ride is defined by verticle motions, pitch and roll
also contribute to the ride of the vehicle.)
Automotive leaf springs are typically composed of an assembly of leaves of two or
three thicknesses because the spring rates of such an assembly are typically better than
when all leaves are of the same thickness.
Absorption and release of energy during the lifetime of the usage of a spring is a
cyclic loading and unloading process. Therefore, spring failures generally occur by a
fatigue mechanism although the root cause may be related to material and processing
factors (Mukhopadhyay et al., 1997; Berns et al., 1985; Tekeli, 2002; Kumar et al., 2000;
Barthold, 1998a). The objective of this paper is to provide an overview of the most
important factors involved in either coil spring or leaf spring failures.

2 Discussion
2.1 Design
Material selection and processing parameters for a coil spring are dependent on: stress
level, minimum number load cycles required from the spring, and environmental
conditions such as operational temperature. When coil springs are compressed, the spring
is subjected primarily to torsional stress. For tension springs, the limiting design factor is
typically the stress concentration in the attachment hooks. Stress reduction may be
accomplished by increasing the wire size although other design changes are also required
to maintain the desired spring rate.
In addition to conventional helical spring design, research has been performed to
identify new technologies to achieve high-tensile spring designs with substantially
improved fatigue failure resistance. One method of achieving this is by design
modifications such as using non-cylindrical spring contours and oveate (elliptical) wire
cross-sections to reduce spring stresses and installation heights (Muhr, 1993; Barthold,
1996). Using such design modifications, Muhr has reported that valve springs with 13%
reduction in spring loads and 35% reduction in installation heights can be achieved.
Ayada et al. (2006) and Kimura and Kurebayshi (2001) have reported that tapered leaf
spring designs coupled with modified ausforming technology where the steel is deformed
above Ar1 temperature and quenched to form martensite would facilitate the production
of springs with improved strength and toughness and which are 50%–70% thinner.
The in-service design requirement for leaf spring materials is dependent on the ratio
of yield stress (σy) to Young’s modulus (E) which must be sufficient to avoid a
permanent set (Mukhopadhyay et al., 1997). The ratio of σy/E is dependent on the
material strength which is dependent not only on the steel chemistry and hardenability,
but also thermal processing parameters such as forming and heat treatment (heating,
quenching and tempering) and any mechanical treatments such as shot peening to assure
surface compressive residual stresses (Nakhimov et al., 1963; Todinov, 2000; Lessells
and Murray, 1941).
266 L.C.F. Canale, R.N. Penha, G.E. Totten, A.C. Canale and M.R. Gasparini

Barthold (1998a) has reviewed various common causes of spring failure due to design
and these are summarised in Table 1. Material effects will be discussed below and a more
detailed discussion on proper design criteria is beyond the scope of this review. However,
a comprehensive reference on the topic is provided (SAE Spring Committee, 1996).

Table 1 Failures caused by poor spring design

Design problem Effect Solution


Insufficient dimensioning Fracture, failure due to setting Improve spring design
Material selection Insufficient performance (failure Better material selection
to meet design requirements)
Influence from surroundings Fractures, loss of function Account for environment in
(thermal, corrosion, etc.) which the spring will be used

2.2 Material effects


2.2.1 Steel alloy for wire manufacturing process
A Steel property requirements and chemistry
Certain physical properties are necessary to produce wire suitable for spring production.
In addition to meeting quality requirements, materials suitable for wire production must
exhibit acceptable high toughness, wear resistance and ductility. Metals that have been
used for springs include: beryllium-copper alloy, phosphor-bronze and titanium.
However, steel alloys are the most commonly used materials for spring production.
Examples of spring steel materials include: high-carbon, oil-tempered low carbon,
chrome-vanadium, chrome-silicon and stainless steel. Fully pearlitic carbon steels are
used for applications where high strength, ductility and toughness are required. In many
applications, hypereutectoid steels (>0.80% C) are specified. Recently, high carbon
content steels (1.0%–2.1% C) have been developed where particularly high strength is
required (Taleff et al., 2002).
Alloying elements that are commonly used for spring steels include chromium,
silicon, niobium and vanadium. Chromium is added to increase corrosion resistance,
hardenability and especially high-temperature strength. Chromium facilitates carbide
formation at higher temperatures. Silicon content affects the thickness of scale formation
during hot-rolling as well as scale porosity (Birosca et al., 2004).
Vanadium additions to eutectoid and hypereutectoid steels have been of particular
interest for wire rod development (Taleff et al., 2002). Increasing vanadium content
promotes finer grain size, increases strength, hardenability, fatigue life and wear (Parusov
et al., 2004). Hata et al. (2004) reported that either, niobium or vanadium, or both, are
added to suppress the recovery, recrystallisation and austenitic grain growth which will
enhance pearlite formation, decrease tensile strength and reduce nodule size and improve,
drawability. This will only occur if the content of niobium and vanadium is less than
0.20% and 0.05%, respectively. Drawability will subsequently decrease if these
concentrations are exceeded due to precipitation strengthening (Hata et al., 2004).
However, excessive vanadium addition will result in excess vanadium carbide content
resulting in reduced toughness (Kawabe et al., 2002). Also, recovery and recrystallisation
after hot deformation in medium carbon steel are inhibited by vanadium addition (Ayada
et al., 2006).
Overview of factors contributing to steel spring performance and failure 267

Kimura and Kurebayshi (2001) and Ayada et al. (2006) have discussed the use of
niobium in spring steels to prevent grain coarsening during heat treatment, reduce
brittleness and increase strength. Niobium addition coupled with ausforming process
resulted in the possible manufacture of leaf springs with a 50%–70% reduction in
thickness while still meeting the specified strength (Kimura and Kurebayshi, 2001).
Kimura has summarised advances in spring alloy development by noting that
typically carbon is added to increase corrosion and fatigue life. Silicon provides
sag resistance. Nickel will retard pitting corrosion. Niobium and boron facilitate
grain refinement and strengthen prior austenite grain boundaries (Kimura and
Kurebayshi, 2001).

B Wire manufacturing process and steel metallurgy


To produce the required diameter and tolerance, high carbon steel wire rods are passed
(drawn) through a series of dies, each of which has a smaller bore diameter than the
proceeding die. Drawability is defined as “the ratio of the number of wire breaks to the
length of wire drawn (breaking, frequency)”.1 This is both a measure of rod quality and
wire drawing practice. Phelippeau et al. (2006) used the drawing stain level (ln So/S)
(where: So is the initial section size and S is the final section size) to quantify the
cold-drawing capability of eutectoid steels.
Ashida et al. (1993) has used the following method to assess the drawability for steel:
wire ≤0.4 mm diameter. Tensile Strength (kgf/mm2) > [270 – (130 × logD)], where D is
the wire diameter (mm). In addition, the reduction in area must not be less than 35%.
Kuroda et al. (2002) reported a more general relationship for the drawability of spring
steel rolled wire where: tensile strength ≤ 1200 MPa and 30% ≤ reduction area ≤ 70%.
Li et al. (2003) determined the drawability of 200 nickel wires of different diameters
and with different heat treatment. The relative drawability and strain rates were assessed
by using Voce’s equation (Equation (1)) and determining the effects of various
parameters such as temperature and grain size.
m

⎡ ⎡ . ⎤
. T ⎤⎢ ε ⎥
σ(ε, ε, T) = [σ1 − (σ1 − σ0 )EXP(− nε)] × EXP ⎢ − k. ⎥ (1)
⎣ TM ⎦ ⎢ ε. ⎥
⎣ 0⎦
where:
n = the material dependent strain hardening factor
m = the strain rate sensitivity which is a measure of flow stress to strain rate
K = the temperature co-efficient of stress
TM = the melting temperature of the wire
ε = a scaling factor which is usually assumed to be 1s–1.
The values σ0, σ1, m, n and K are assumed to be temperature independent. The unknown
parameters of Equation (1) are estimated from experimentally measured data for: drawing
force, temperature distribution and radius profile in the deformation region and solving
for the Voce parameters by multiple linear regression analysis.
Before the cold-drawing operation is performed, the wire is heated to a
red-hot temperature (approximately 910ºC) in a hot-rolling operation. Heating the wire to
such high temperatures, produces relatively large grain size austenite. Excessively
268 L.C.F. Canale, R.N. Penha, G.E. Totten, A.C. Canale and M.R. Gasparini

large or coarse grain can also be considerate a defect. Generally, coarse grain
size increases as the hot-rolling temperature increases and the cooling rate decreases
(Ayada et al., 2006).
After austenitising, the wire is subsequently cooled in a ‘patenting’ process where
the austenitised wire is cooled in a bath of molten lead or in a fused salt mixture and
held at a temperature of below 500ºC to form fine pearlite with preferably no separation
of primary ferrite, which is considered to be the best microstructure to assure good
cold-drawability (Taleff et al., 2002). The wire is held at this temperature for at least
10 s. Patenting is typically performed on wires smaller than 13 mm in diameter.
One of the greatest problems with patenting wire in molten lead or salt, is the
relatively low (1.8–3.0 tons/hr) production capacity. Therefore, an alternative air-blast
cooling process has been developed which is capable of producing wire in a continuous
system at greater than 40 tons/hr (Paulitsch et al., 1979). Such a wire production line is
called ‘Stelmore system’. A Stelmore system typically consists of a water-quenching
process to cool the wire (approximately 5.5 mm) to less than 850ºC at the end of
the first conveyor, at which point the remainder of the cooling is performed by air
blasting using a fan system (Kazeminezhad and Taheri, 2003). This process is illustrated
in Figure 2.

Figure 2 Diagram of the Stelmore line

Another disadvantage of lead patenting is that molten lead can only not be dragged out
with the wire upon exiting from the lead pots but it can also lead to surface defects on the
wire, which are caused by lead contamination.
There have been various attempts to use other patenting media such as hot-water, at
least 85ºC. This water temperature is selected to facilitate the patenting process to be
conducted with astable film boiling heat transfer process occurring at the surface of the
wire (Vanneste, 1988). However, such process often results in non-uniform drawability
and an unacceptable brittle wire. Furthermore, small variations in cooling rate, caused by
Overview of factors contributing to steel spring performance and failure 269

non-uniform scale formation on the wire, may result in non-uniform cooling due to
non-uniform film rupture around the wire and the formation of hard spots of martensite
and bainite.
Cold-drawing leads to an increase in mechanical strength and loss of ductility of the
steel wire. Heavily deformed materials exhibit a plastic strain localisation and the
as-draw wire contains residual stresses, which are sufficient to produce a hardening effect
during a post drawing tensile test. After the material can no longer be hardened, the
residual stresses contained within the wire then control the wire elongation during the
cold-drawing process.
A common defect that occurs during the wire drawing process is breakage or necking
down and subsequent fracture (cupping) as the wire exits the die as shown in Figure 3.
Some potential causes of this type of failure include (Hata et al., 2004; Grimwade, 2006):
• The wire is overworked and requires annealing.
• The amount of wire size reduction during the drawing step is excessive
(overdrawing). Although larger reductions in wire size (up to 25%–45% reduction)
may be performed, depending on the alloy, the permissible wire reduction decreases
with each successive drawing step and as the wire diameter decreases.
• There may be a loss of lubrication causing an increase in friction between the wire
and the die resulting in a decrease in the amount of wire reduction possible.
• The presence of inclusions in the wire, this will result in weak spots in the wire. If
nodules within the wire are significantly small will rotate smoothly during the
drawing operation, thus preventing the formation of microvoids and subsequent
‘cuppy breakage’. One method of controlling nodule size is to control the air blast
cooling process on the Stelmore line after hot-rolling (Grimwade, 2006).

Figure 3 Schematic illustration of ‘cuppy’ wire breakage during cold-drawing

One common problem is lubrication failure during cold-drawing. Dhua et al. (2004) have
reported that there have been many reports of the appearance of white etching layers on
the surfaces of high-carbon steel wire that broke prematurely. This is a friction-induced
process (formation of friction martensite) and is associated with the appearance of
transverse cracks.
270 L.C.F. Canale, R.N. Penha, G.E. Totten, A.C. Canale and M.R. Gasparini

Changes in ductility (plastic damage) may occur during the deformation process in
wire forming (McAllen and Phelan, 2004). If the deformation during wire drawing is not
uniform, the resulting tensile stresses may be sufficient to lead to microvoid growth and
crack propagation.
Other causes of wire breakage during cold-drawing, are: improper metallurgy,
incorrect die profile, excessive wire reduction, segregation, porosity and surface defects
including laps, seams and fins (Dhua et al., 2004).
It should be noted that if any structure other than fine pearlite, such as ferrite or
bainite, is produced in amounts of 5% or more, the resulting wire rod is considered to
possess poor drawability (Hata et al., 2004). Zelin (2002) has studied various
microstructural processes that occur during pearlite formation and deformation during the
wire drawing process. The effects that were studied included: lamellae thinning,
evolution of lamellae interfaces, texture changes, localisation of plastic flow and dynamic
strain aging. Figure 4 shows a microstructure obtained by Atomic Force Microscopy
(AFM) of the pearlitic microstructure of a eutectoid steel, patented and cold-drawn,
where it is possible to observe ridges of cementite that have been plastically deformed
(Elices, 2004).

Figure 4 AFM representation of the pearlitic microstructure of a eutectoid steel after patenting
and cold-drawing. Sample was etched. Valleys were filled with ferrite and ridges
are cementite.

Source: Courtesy of M. Elices

Phelippeau et al. (2006) have studied the evolution of the microstructure during
cold-drawing of an eutectoid steel. At the beginning of process, the steel is pearlitic
with a lamellae thickness approaching 200 nm, as shown in Figure 5(a) and the
grain size of 20 µm. After cold-drawing (strain level of 3.5), the thickness of the
lamellae is less than 20 µm, as illustrated in Figure 5(b) and the grain size is reduced to
less than 1 µm.
Overview of factors contributing to steel spring performance and failure 271

Figure 5 Evolution of the eutectoid steel microstructure: (a) initial pearlitic structure (etching
Nital 3%); (b) nanostructure after cold-drawing (TEM bright filled image)

(a) (b)

Zolotorevsky et al. (2004) reported that although it is known that the strength of drawn
pearlite increases as the pearlitic lamellar structure evolves, there is also a reduction of
the cross-sectional area under tensile testing which is due to localised ductility and the
true strain increases from less than 1.5 to 2.0. The localised ductility was reportedly due
to texture development due to stress concentration arising from colony deformation as the
lamellae become oriented to the wire axis (Zolotorevsky et al., 2004).
If the wire rod is cooled too quickly from the rolling temperature, undesirable bainite
or martensite can be formed throughout the cross-section of the wire or locally on the
surface which may produce hard spots. The resulting wire is unsuitable for subsequent
drawing. Hard spots on the surface can also be due to the presence of friction martensite
or bainite which is caused by friction between the wires and guides (Ayada et al., 2006).

2.2.2 Steel alloys for helical coil springs


The fatigue strength that is attainable with a steel increases with carbon content.
However, steel composition varies with the design and desired end-use properties. Alloys
for coil and leaf springs have different requirements which will be briefly discussed here.

Hard-drawn carbon steel spring wire


This spring material is typically a silicon-kilned steel containing 0.45%–0.75% C and
0.60%–1.2% Mn. The steel wire undergoes a patenting (high-temperature process
conducted at 450ºC–570ºC) process then it is cold-drawn which increases strength
without loss of ductility. Because of the high-temperature isothermal patenting process,
the cross-section size is limited to 12.5 mm. To avoid setting the maximum design stress
should not exceed the torsional elastic limit (Kern and Suess, 1979).

Music spring wire


Musical instrument strings are a demanding application since the strengths are typically
under high tension and subject to repeated blows. Yet they must last for decades. The
wire must be consistent in size and variation >0.0076 mm can produce noticeable
272 L.C.F. Canale, R.N. Penha, G.E. Totten, A.C. Canale and M.R. Gasparini

variation in the sound. This material is a high-tensile strength high-carbon cold-drawn


wire with higher tensile strength and higher torsional strength than any other material
available. The high toughness characteristic of this material is obtained by a wire
patenting process (to be discussed subsequently). Music wire is purchased according to
tensile strength, not hardness. This is a high quality hard-drawn spring wire and is the
least subject to embrittling effects of plating.

Oil-tempered carbon-steel spring wire


This is a spring wire commonly used for many applications where outstanding strength
or uniformity is not critical. Heat treatment of this spring material is performed
by austenitising, oil quenching, and tempering by passing the wire through a molten
lead bath (Jarl et al., 2003) at a specification temperature which will produce the
desired mechanical properties. Hardness is determined by the tempering temperature, not
cold-working. Because of the tempered martensitic microstructure, this steel is more
susceptible to embrittlement after plating than hard-drawn wire (Kern and Suess, 1979).

Oil-tempered valve-spring wire


This Cr-Si or C-Cr-Si steel is used for valve spring production is of oil tempered wire and
the use of this steel is one of the most demanding applications with respect to fatigue
failure. Engine values operate under very high dynamic stresses (Bertrand et al., 2003).
To obtain the desired relaxation properties, this steel is used in the oil-tempered
condition. It is critical that the surface be free of defects such as decarburisation and free
of seams, scratches, die marks, pits and other defects that will lead to reduced fatigue life.
Non-metallic inclusions are also critically important defects that must be minimised such
as with the use of superclean steel to assure optimal fatigue life (Barthold, 1996; Bertrand
et al., 2003; Eriksson, 1995).
Kawabe et al. (2002) and Nam et al. (2000) have reported that there has been a more
recent trend to add vanadium to C-Cr-Si steel with an increased level of carbon to
increase fatigue life. Increased silicon is used to reduce the loss of strength when heat
treated at higher temperatures. Higher tempering temperatures are used to: improve
workability and toughness of the steel and to reduce micro defects. Higher strain
relieving temperatures are used to reduce residual stress, which is introduced during
coiling and micro defects that lead to reduced fatigue strength. The presence of vanadium
reduces toughness, thus it is desirable to minimise its presence to achieve optimal
performance properties.
The automotive industry typically stipulates the following requirements for
valve-spring material (Barthold, 1996):
• limitation of aluminium (30–50 ppm), oxygen (25–30 ppm), and titanium (30 ppm)
• absence of Al2O3 inclusions
• limit inclusion size to conform with ASTM guidelines, specification of permissible
distributions of inclusion size
• removal of boundary layers at least 150 µm thick from the wire rod surface by
mechanical working, such as peeling, particularly for Cr-Si-alloyed steel
Overview of factors contributing to steel spring performance and failure 273

• continuous inspection of the wire for surface defects (maximum depth of 40 µm)
using two mutually independent test methods
• fine-tempered microstructures with depth of decarburisation limited to 24–32 µm,
depending on the wire diameter.
Barthold (1996) has reported that in order to obtain significant improvements in spring
durability, there has been a trend from conventional Cr-V alloyed wire to Cr-Si wire
which is stronger by 200 N/mm2 and exhibits better thermal resistance due to its higher
silicon content. Furthermore, the use of ‘Superclean’ processed steel is increasing. The
Superclean process reduces the formation of hard, deformation-resisting Al2O3 inclusions
and high-melting calcium aluminates which are formed by the Si deoxidation process.
Instead, a minimum number of CaO-Al2O3-SiO2 silicate inclusions near the eutectic point
just below the SiO2 corner shown in Figure 6 (Barthold, 1996).

Figure 6 Inclusions present in valve-spring wire


274 L.C.F. Canale, R.N. Penha, G.E. Totten, A.C. Canale and M.R. Gasparini

Alloy spring wire


Alloy spring wire is used when: resistance to relaxation up to temperatures of 232ºC,
higher tensile strength, and higher torsional elastic limits than achievable with
carbon-steel wire are desired. These are Cr-V or Cr-Si steels similar to AISI 6155 and
9254, respectively. To obtain maximum fatigue life, oil-tempered valve-spring quality is
recommended. The 9254 steel exhibits excellent high-temperature (232ºC) and fatigue
(>108 cycles). This steel has a tendency to crack when drawn or coiled.

Stainless steels
When a spring may be used in a corrosive or at a high-temperature in excess of the
operational range of conventional spring steels, a stainless steel material may be used for
springs up to 0.51 mm. Type 302, 303 and 304 stainless steels exhibit excellent corrosion
resistance and resistance to thermal relaxation up to 260ºC. Type 17–7 PH steel, which is
age-hardened by heating up to 482ºC for one hour after coiling exhibits corrosion
resistance comparable to Type 302 but may be used up to higher temperatures 347ºC.
(Table 2 provides approximate use-temperature limits for different spring materials.) To
achieve maximum corrosion resistance, springs manufactured from stainless steel must
be passivated after heat treatment (Kern and Suess, 1979).

Table 2 Approximate use-temperature conditions for different spring materials2

Approx. max. Approx. max.


Wire Material temperature (ºC) Wire material temperature (ºC)
Music wire 120 302 Stainless 260
Hard-drawn carbon 120 17–7 PH 320
Oil-tempered carbon 150 Inconel 600 370
Chrome-vanadium 220 NiCr A286 510
Chrome-silicon 245 Inconel X750 590

Typically, austenitic stainless steel wires do not simultaneously exhibit high tensile
strength, heat resistance and corrosion resistance. However, recent developments utilising
both improved wire drawing technology and heat treatment along with small addition
of nitrogen have reportedly produced stainless steel wires that solve this problem
(Isumida et al., 2005).

2.2.3 Steel alloys for leaf springs


For leaf springs, steels containing at least 0.55% carbon and which are heat treated
to a HRC hardness of 40–45 are typically used. Most steels for leaf spring production
are manufactured with approximately 0.6% carbon, although AISI 1095 and 4068
provide exceptions to this general rule (Kern and Suess, 1979). The most common
leaf spring steel alloys are 4161H, 5160H, 50B60H, 8660H and 9260H. Of these, AISI
9260H (0.65%–1.1% Mn, 1.8%–2.2% Si) and 5160H (0.45%–0.65% Cr) are more used
for leaf spring production when compared to AISI 9260 steel (0.75%–1.0% Mn,
1.8%–2.2% Si), and 5160 steel (0.7%–0.9% Cr).
Overview of factors contributing to steel spring performance and failure 275

AISI 9260 (steel is one of the oldest spring steel alloys. However, the relatively high
silicon content makes it a difficult steel to manufacture. Problems with internal
cleanliness and poor surface quality are common problems with this steel. Other
problems that have been reported with Si-steels include: susceptibility to decarburising,
grain growth and low hardenability (Rakhshtadt, 1975). Owing to its relatively poor
hardenability, this steel grade is suitable for leaf springs up to thicknesses of 0.500 in
(Kern and Suess, 1979). More recent work has shown that the properties of this steel
grade can be improved by decreasing the amount of silicon and/or increasing the amount
of manganese. Properties can be improved further by the addition of boron or titanium.
The addition of 0.5%–1% of Cr or Cr/V is reported to triple the hardenabilty and reduce
the depth of decarburisation (Rakhshtadt, 1975).
Of the different steel grades used for leaf spring production, AISI 5160H (Cr-Mn) is
manufactured in greatest tonnage. And is suitable for leaf springs up to 16 mm thick.
However, although it is a relatively easy steel to melt and yields generally good
cleanliness and surface quality, it does exhibit a number of substantial disadvantages,
such as low hardenability and chromium banding.
AISI 50B60H exhibits considerably better hardenability, fracture toughness and
tensile strength than 5160H and the boron inhibits the formation of proeutectoid
ferrite. Rakhshtadt (1975) reported that the addition of as little as 0.005% B increases the
elastic limit of steel by 20–30 MPa and the relaxation resistance. It exhibits good internal
cleanliness and surface quality. This alloy is suitable for leaf springs up to 35 mm thick.
The highest quality spring steel and the alloy often recommended for optimal
load-carrying properties is AISI 8660 and is suitable for the manufacture of springs up to
35 mm with no boron and 2 in with boron present. Although the Ni content provides
excellent load carrying properties, it tends to form a scale during rolling which produces
quenching defects and poor surface quality.
Finally, AISI 4160H (Cr-V) is a high hardenabilty steel which is suitable for leaves
up to 51 mm thick. However, this alloy is to form chromium banding leading to poor
transverse properties. Some heats are susceptible to quench cracking (Kern and
Suess, 1979).
Kern and Suess (1979) have reported that a good rule of thumb is for steels for leaf
spring production should exhibit a hardenability sufficient to produce at least 80%
martensite in the centre of the cross-section.

2.2.4 Effect of inclusions and surface defects


In the forgoing discussion, the tendency of various spring steels to form surface defects
and inclusions has been noted. This is important because two of the primary causes of
fatigue failure are inclusions and surface defects. For example, Figure 7 shows that
fatigue strength decreases with increasing inclusion size and Figure 8 shows that the
number of cycles obtained by the Nakamura test for fatigue life decreases with increasing
inclusion size for a precipitation-hardening (PH) spring wire (Haldex Garphyttan, 2006).
The potential deleterious effects of inclusions are related to their melting point and
hardness. Typically low melting point and low hardness of an oxide inclusion will result
in lower localised stresses around the inclusion, which increase fatigue strength from
approximately 10 – 20 × 106 to 100 × 107 cycles. This improvement in fatigue properties
has been a major driver in the continued development of fatigue-resistant superclean
steels for spring applications (Bertrand et al., 2003).
276 L.C.F. Canale, R.N. Penha, G.E. Totten, A.C. Canale and M.R. Gasparini

Figure 7 Influence of the size of inclusions on fatigue strength for a PH steel

Figure 8 Goodman diagram illustrating the dependence fatigue resistance on the size of
inclusions in a precipitation hardenable spring steel

In general, non-metallic inclusions are sulfites, oxides, silicates or mixed compounds


of these compositions (Ayada et al., 2006; Eriksson, 1995). Non-metallic inclusions may
be formed by metallurgical reactions. Most non-metallic inclusions of oxides are
represented by the ternary phase diagram for CaO-Al2O3-SiO2 shown in Figure 4
(Paulitsch et al., 1979; Bertrand et al., 2003; Eriksson, 1995; Kawahara et al., 1992;
Barthold, 1998b). Table 3 provides a summary of common inclusion elements and their
resulted effects (Zhang et al., 2002). Bertrand et al. found when MgO and Al2O3 content
was minimal, most of the remaining inclusions possessed a chemical composition close
to the interface wollostonite-anorthite-gelhenite with a melting point <1400ºC (Bertrand
et al., 2003). When the melting point is reduced to this level, the elongation rate is greater
than that of the steel showing that inclusions size is related to their melting point
(Kawahara et al., 1992).
Overview of factors contributing to steel spring performance and failure 277

Table 3 Typical steel inclusion elements and effects on the mechanical properties

Element Form Mechanical properties effects


Si, O Sulfide and oxide inclusions Ductility, charpy impact value
Anisotropy
Formability
Cold forgeability, drawability
Low temperature toughness
Fatigue Strength
C, N Solid solution Enhanced solid solubility, hardenability
Settled dislocations Enhanced strain aging, ductility and lower toughness
Pearlite and cementite Enhanced dispersion, ductility and lower toughness
Carbide and nitride precipitates Precipitation, enhanced grain refining, enhanced
toughness
Embrittlement by intergranular precipitation
P Solid solution Enhanced solid solubility, hardenability
Temper brittleness
Separation, secondary work embrittlement

Nitrogen is an important inclusion element in clean steel wire. For wire applications,
the maximum nitrogen content specification is ≤60 ppm, to total oxygen ≤30 ppm
and maximum size is 20 µm (Zhang et al., 2002). However, even nitrogen inclusions
of approximately 0.01% (100 ppm) may result in a slight increase in the yield point
and tensile strength but deformability will decrease significantly (Ayada et al., 2006).
The manufacture of clean steel for valves involves controlling not only inclusion
content but their size as well (Zhang et al., 2002). Bertrand et al. (2003) found that
if oxide inclusions were <15 µm, no inclusion failures were observed in their fatigue
tests. Hagiwara et al. (1991) have provide numerous characterisation examples of
typical inclusion shape identifications resulting from steel process variations. Typically,
spring steel is characterised as either conventional (Al-deoxidised) and superclean
(Si deoxidised) (Eriksson, 1995). For Al deoxidised, compression springs, the greatest
stresses are just bellow the surface; typically 0.050–0.500 mm from the surface. The most
common inclusion sizes are 0.015–0.040 mm. The values for Superclean steel shown in
Table 4 were reported by Ericksson (1995).

Table 4 Reported cleanness levels for various steel grades

ASTM Size class (µm) 5–10* 10–15* >15*


A877 Superclean 3.0 0.3 0.0
A877 Conventional 9.9 3.4 0.3
A878 Superclean 4.4 0.6 0.0
A878 Conventional 28.6 2.1 0.2
2
Note: * Values as numbers/1000 mm .
278 L.C.F. Canale, R.N. Penha, G.E. Totten, A.C. Canale and M.R. Gasparini

2.2.5 Metallic oxides (scales)


In the following discussion, impurities will be discussed as metallic and metallic oxides
and non-metallic inclusions. Inclusions may be metallic impurities or metallic oxides
(Zhang and Thomas, 2003). Metallic element inclusions (impurities), although typically
in trace quantities, may be originate the scrap used for the steel-making process. These
elements cause intergranular segregation which may lead to crack formation, detrimental
precipitate formation and are often observed as ‘slivers’ in the final product.
The amount of scale formation is dependent on the hot-rolling conditions
(Krzyzanowski et al., 2003). Parusov et al. (2004) studied the effects of scale formation
and removal from wire rod before drawing. Although it is possible to minimise
oxide scale formation on wire rod during hot-rolling, it is extremely difficult to
completely eliminate it. Formerly, scale was removed from wire rod by pickling,
however, currently chemical methods of scale removal are generally prohibited due to
various environmental regulations.
Scale formation due to air contact is related to steel chemistry and methodologies
used for furnace heating and cooling. For example, it has been shown that scale removal
is facilitated by the presence of copper and sulphur while the presence of silicon, nickel
and cobalt make more difficult for scale removal (Birosca et al., 2004; Parusov et al.,
2004). Chromium was found to exhibit no effect (Parusov et al., 2004). There is a
correlation between the colour of an oxide scale and the content of non ferrous metal
impurities (Frank and Kirkaldy, 1988). When the content of chromium, nickel and copper
is less than 0.05% and when the content of chromium, nickel and copper is greater than
0.05% up to more than 0.30%, the surface colour of wire made in a electric furnace is
lighter than if these compositional requirements are not met. Black films on the wire
surface are composed of magnetite and are difficult to remove due to strong adhesion to
the steel surface (Parusov et al., 2004). The formation and growth of oxides layers is
dependent on the time the steel is held at high temperatures, prior to rolling, and wire
rolling speed, oxidation medium and cooling rate, whether air or water is the cooling
medium. The total amounts of scale formation increases with the rolling and coiling
temperature. Total scale formation also increases with decreased air cooling rate (Frank
and Kirkaldy, 1988). Scale formed at temperatures greater than 400ºC has been shown to
contain three layers consisting of: magnetite (Fe3O4), hematite (Fe2O3) and wüstite (FeO)
(Parusov et al., 2004). The composition of the scale on the steel surface is traditionally
considered to have wüstite (FeO), which contains the least amount of oxygen, closest to
the steel substrate. Magnetite (Fe3O4) composes the intermediate layer. Finally, a thin
oxygen-rich layer of hematite (Fe2O3) on the surface (Birosca et al., 2004). However, this
compositional view of scale is affected by heat treatment conditions, the oxidation
environment and steel composition.
Surface oxidation occurs through an ionic diffusion process in the range of
700ºC–1200ºC (Parusov et al., 2004; Frank and Kirkaldy, 1988).
As the oxidation temperature increases, the diffusion of iron ions and oxygen ions
increases which leads to increased scale thickness ionic. Furthermore, the presence of this
diffusion mechanism is also affected by scale porosity, cracks in the scale and adhesion
of the scale on the steel. This is of particular concern because the presence of such scale
can potentially influence heat transfer and friction during the wire drawing process as
well as the surface finish of the final product (Birosca et al., 2004). Thus, the oxide scale
must be removed before the wire drawing operation (Krzyanowski et al., 2003).
Overview of factors contributing to steel spring performance and failure 279

The degree of scale removal increases as the total amount of surface oxides increase
and as the temperature of the metal before air cooling decreases. Poor scale removal is
typically due to thin films composed of higher oxides (Fe3O4) (Parusov et al., 2004). Thin
film scale formation is due to instability of wüstite at low cooling rates.
There are various factors that affect the formation and ease of removal of secondary
scale. One factor is copper formation on the surface of the steel if the steel were heated
just before rolling. This may produce non-uniform formation of secondary scale on the
surface of the wire and produce variable surface roughness. It has been shown that
increasing ratios of Cu/Si and (Cu/Si)/(P/S) increase the difficulty in removing secondary
scale by mechanical processes (Parusov et al., 2004).
Another factor affecting secondary scale removal is increasing surface roughness
which occurs when primary furnace scale is rolled into the surface of the wire. Generally,
the greater the wear of the rolls, the greater the adhesion of the scale.
Relatively low amounts of iron oxides (FeO, Fe2O3, Fe3O4) with a surface layer
thickness of 5–7 µm, are formed when the coiling temperature is 750ºC–800ºC. Such
scale is a rusty ginger colour, although it typically does not produce poorer properties and
it is nearly impossible to be effectively removed by mechanical (or chemical) processes.
At temperatures less than 650ºC, magnetite dominates the overall scale composition
(Birosca et al., 2004). Increasing the coiling temperature, increases the scale thickness to
8–14 µm which can be completely removed from the wire surface (Parusov et al., 2004).
As the temperature increases, wüstite becomes the more dominant component. However,
upon coiling, it is possible for wüstite to transform to magnetite (Birosca et al., 2004).
Figure 9 shows the compositional dependence of the different oxides that are formed in
the scale as a function of temperature (Birosca et al., 2004).

Figure 9 Molar fraction of wüstite, magnetite and hematite as function of temperature


Mol. Fraction (%)

Temperature (ºC)

Increasing the conveyor speed of the Stelmore line (see Figure 2) or increasing the
cooling rate to the critical temperature range (400ºC–570ºC) facilitates scale removal.
Finally, the location of the wire rod on the conveyor in the Stelmore line will also affect
the nature of the oxides formed and thus the ease of removal (Birosca et al., 2004).
Metallic oxide inclusions vary in morphology and composition and may be classified
as illustrated in Figure 10 (Taleff et al., 2002). For steel manufacture, in general, some
sources of these oxides include:
280 L.C.F. Canale, R.N. Penha, G.E. Totten, A.C. Canale and M.R. Gasparini

• Deoxidation products such as alumina inclusions which are formed by the reaction
of dissolved oxygen and the added deoxidant such as aluminium. Alumina inclusions
are dendritic when formed in the presence of high oxygen concentration as illustrated
in Figures 10(a) and 10(b) (Zhang and Thomas, 2003) or they may be formed
by the collision of smaller particles as illustrated in Figure 10(c) (Zhang and
Thomas, 2003).
• Reoxidation products are generated when the aluminium remaining in the liquid steel
is oxidised by FeO, MnO or SiO2 and other oxides in the slag or refractory materials
or by exposure to the atmosphere.
• Slag entrapment occurs in metallurgical fluxes are entrained in the steel. This occurs
during transfer between steelmaking vessels. These inclusions are typically spherical
as illustrated in Figure 10(d) (Zhang and Thomas, 2003).
• Exogenous inclusions from other sources include dirt, broken refractory brickwork,
and ceramic lining materials. They act as sites for heterogeneous nucleation of
alumina and may include a central particle such as that shown in Figure 10(c) (Zhang
and Thomas, 2003).
• Chemical reactions may produce oxides from inclusion modification when Ca
treatment is imperfectly performed. Inclusions containing CaO may also originate
from entrained slag.

Figure 10 Typical steel inclusion morphology and compositions

Source: Zhang and Thomas (2003)

2.2.6 Surface defects


Among the most common surface defects, the authors would like to point out for some
specific damages, such as: seams (described by Roberts and Leahy, 1998, as a tightly
rolled discontinuity in the wire surface often invisible to naked eye), laps (which are
defects caused by hot metal folding back on its own surface during rolling), pits
(characterised by a depression on the rolled surface invisible to the naked eye), and
scratches (furrow-like depressions running longitudinally on the surface of the wire, as
shown in Figure 11, which may occur during the drawing process or they may be caused
by poor machining, worn or broken guides, according to Ayada et al., 2006). These
defects arise from casting, rolling, forming processes. Defects such as these are readily
observable, even continuously, using instrumentation such as eddy current testers. This is
important since these defects act as stress risers (concentrators) and the surface is
subjected to the highest bending and torsion stresses. Generally, a fatigue crack will start
at a surface defect on the spring leaf. Therefore, care needs to be used when
manufacturing and installing springs to reduce these defects to a minimum. Even very
Overview of factors contributing to steel spring performance and failure 281

slight surface damage is enough to shorten the service life of a spring substantially. For
this reason, methods such as electropolishing and hydroabrasive polishing are performed
to remove these defects.

Figure 11 Illustration of typical surface scratch defect on a spring wire as-drawn

Source: Courtesy of Kieselstein GmbH, Chemnitz, Germany

Some specifications for carbon steel permit defect or seam levels as great as 3% of the
wire diameter (Perlick, 2003). Higher levels of defects will result in unacceptable loss of
strength. Lubricants are used during the cold-drawing process to assist in providing more
uniform metal flow around bending or coiling points and to reduce friction and surface
damage of the wire.
Surface defects, which include decarburising, laps, etc., result in lower fatigue
properties. Furthermore, the greater the effective stresses on the springs, the greater the
importance in manufacturing a spring to the highest quality level possible. Optimum
surface quality is obtained by removal of surface defects by grinding, shaving or wire
peeling before coiling (Kuroda et al., 2002; Paulitsch et al., 1979; Eriksson, 1995;
Kieselstein, 2006; Shinohara and Yoshida, 2006; Kawabe and Murai, 2003; Kawabe
et al., 2003). Barthold (1996) has reported surface boundary layer removal of at least 150
µm is generally recommended.
Currently, wire peeling, either turn-peeling (wire turning) or draw-peeling, is
preferred in most cases over grinding (Kieselstein, 2006). Some of the advantages
obtained include: removal of surface decarburised layers, removal of rolling cracks,
scores and pores and smoothing of the surface with low surface roughness removal of
decarburisation by draw-peeling After the peeling process, the wire may be stress-relief
annealed followed by a short shot peening operation. Spring wire with superior
workability for shaving or peeling typically possesses: tensile strength ≤1200 MPa and
30% ≤ reduction area ≤ 70% (Kuroda et al., 2002). Figure 12(a), 12(b) and 12(c) show
microstructure before and after draw-peeling.
Shinohara and Yoshida (2004) studied the effect of growth and subsequent
disappearance of cracks and scratches on wire during drawing. They found that the
largest group of surface defects originate from wire drawing and that many of their
defects are due to misalignment of the moving wire through the die.
Table 5 provides a summary of common material-related defects that may be
encountered with Spring steels (Barthold, 1998a).
282 L.C.F. Canale, R.N. Penha, G.E. Totten, A.C. Canale and M.R. Gasparini

Figure 12 (a) Wire rod with marginal decarburisation (after cross-sectioned grinding);
(b) draw-peeled valve spring wire with martensitic structure in the boundary area
(after cross-sectioned grind); (c) draw-peeled valve spring wire

(a) (b) (c)


Source: Courtesy of Kieselstein GmbH, Chemnitz, Germany

Table 5 Summary of common material defects

Failure (Defect) Effect Solution


Metallurgical defects
Poor alloying Spring does not conform to
specification
Nonmetallic inclusions Vibration fractures Use superclean materials,
optimise residual stress
Segregations Non-uniform hardness Ingot casting (not
through the cross-section avoidable during
continuous casting)
Shrinkage cavities Wire breakage during coiling
Failure due to wire production
Die marks, button-type defects, Wire breakage during Continuous,
rolled-in impurities, transverse coiling, vibration fractures non-destructive
cracks surface inspection
Incorrect thermal treatment Reduced endurance life due Optimisation of thermal
to coarse structure treatment
Twist, wire not drawn straight Variation in spring geometry Wire straightening – if
possible
Rolling defects
Scores/cracks Vibration fractures Shaving or grinding of
wire rod
Edge decarburisation Vibration fractures Shaving or grinding of
wire rod
Transportation damage
Mechanical damage Wire breakage during coiling Shot peening of springs
Corrosion Vibration fractures Shot peening of springs

2.3 Heat treatment processing


In the production of a coil spring, the wire is subjected to a number of thermal and
mechanical processes as shown schematically in Figure 13 (Kawabe et al., 2002).
Although thermal processing may include rolling, drawing, heat treatment, etc., selected
Overview of factors contributing to steel spring performance and failure 283

processes involved in heat treatment including tempering, stress relieving, nitriding, and
will be discussed here. The effects of rolling, drawing, forging, were covered in the
previous section. In addition, mechanical processes, including shot peening, will also be
discussed in this section.

Figure 13 Illustration of decarburisation which has occurred on cold-drawn steel. After quenching,
the surface ferrite, which is detrimental to the product, will remain untransformed.

Decarburisation
When steel is heated at high temperatures, such as those that are used in hot-rolling,
forging, or heat treated in improperly controlled austenitising atmospheres, there may
be a loss of carbon at the surface, which is known as decarburisation. Decarburisation
occurs when the steel is heated for excessively long times at a high temperature or if the
furnace atmosphere reacts with the surface region of the steel resulting in a loss of
surface carbon (Ayada et al., 2006). Decarburisation can significantly reduce strength and
fatigue life of the steel (Kuźmiński and Nowakowski, 2001; Nomura et al., 2000). For
leaf springs, the following limits for decarburising are: total carbon-free depth = 8 µm
and total affected depth = 0.38 mm. Kern and Suess (1979) report that if the steel is shot
peened, an additional 25.4 µm on carbon-free depth and 0.13 mm on the total affected
depth may be allowed. For valve spring steel, maximum depth of partial decarburising is
limited to 24–32 µm depending on the wire diameter (Barthold, 1998b). Furthermore,
decarburisation beyond 10% of the ladle carbon concentration must be avoided.
Therefore, it is necessary to minimise decarburising depth for both rod and wire.
Sometimes this has been done by using induction heating in combination with
shot peening (Sakakibara et al., 1993).
Figure 14 provides a microstructural illustration of a decarburised steel surface
(Prawoto et al., 2004). Although decarburisation occurs during thermal processing, it is
often classified as a surface defect. Decarburisation reduces the sharpness of a stress
concentrator by a stress relaxation mechanism. However, if a crack, for example, passes
through a decarburised layer, crack propagation is accelerated in the transition layer or
the base metal where higher stresses occur at the tip (Rakhshtadt, 1975).
One method of correcting decarburisation problems is by drawing or peeling.
Alternatively, Prawoto et al. (2004) have reported that carbon may be restored to the
decarburised surface. In this work, the steel was preheated to 400ºC and then coated with
a commercial carbon-rich black powder and then heated at 900ºC or 1050ºC (also known
as ‘pack carburising’). After diffusion, the steel was heat treated and quenched.
284 L.C.F. Canale, R.N. Penha, G.E. Totten, A.C. Canale and M.R. Gasparini

Figure 14 Microstructural illustration of a cold-drawn spring steel where decarburisation


has occurred

Source: Prawoto et al. (2004)

Nitriding
Spring wire may be nitrided for improved wear and corrosion resistance. In addition,
significant improvement in fatigue strength is also possible. For valve springs, the
presence of a nitride case provides excellent high-temperature strength (Funatani, 2004).

Dynamic (strain) aging


Strain-aging refers to the alternation in the stress-strain properties of a wire after previous
cold-drawing (deformation) (Bannister, 1953). Heating at a controlled temperature for a
specific time will improve the elasticity of the wire. As the spring wire is drawn through
a series of sequentially smaller dies (drawing), the strength and hardness of the spring
increases with each drawing step. This is due to the sub-microscopic dislocations
interacting (called ‘pile-ups’) with each other or with the grain boundaries. The hardness
of the wire will continue to increase until it reaches a limit called the ‘spring temper’ at
which point the yield strength and tensile strength are approximately the same. At this
point, the wire can no longer plastically deform when the spring is coiled or formed.
This is not only a cold deformation strain hardening process, but is may also include a
strain-induced phase transformation as well. The cold work hardening rate helps to
determine the annealed wire size at the beginning of the process and also how much total
reduction is required to achieve the spring temper value (Perlick, 2003). Dynamic aging
(dynamic strain aging) is an important thermal process in steel wire drawing (Bannister,
1953; Zelin, 2002; Karimi Taheri et al., 1995; Vorob’eva and Rakhstadt, 1996).
Cementite dissolution may be observed during wire drawing. It is thought that during this
process dissolution occurs by dragging the carbon atoms by ferrite dislocations which
cross the cementite lamellae and which results in an over-saturation of ferrite by carbon.
Adiabatic heating due to plastic deformation may potentially increase the wire
temperature by up to 150ºC–250ºC causing a dynamic aging process. Although this may
result in up to a 5% increase in strength and ductility decrease, there is an increased
susceptibility to delamination with increased aging. This can lead to axial cracking when
the wire is tortionally loaded (Funatani, 2004).
Overview of factors contributing to steel spring performance and failure 285

Grain size
Plastic deformation, such as that occurring during hot–rolling and cold-drawing is
dependent on dislocations in the crystal lattice structure of the steel and which are
commonly present in all metals. These dislocations lead to a localised slipping motion
(slip) of a plane of atoms sliding over another plane of atoms in the crystal structure. This
slip is initiated by the application of a shear stress, which is sufficient for plastic
deformation of the crystal structure. With the imposition of shear stress (σy), the
dislocations will move along a slip-plane until they meet an obstacle to movement. Grain
boundaries, precipitates or dislocations (irregularities) in the crystal structure may act as
an obstacle to slip-plane movement at which point further movement of the slip-plane is
prevented. Subsequent dislocations moving on the same slip-plane will result in a
‘pile-up’. These localised dislocations result in pile-ups leading to a reduction of the
critical value of applied stress which is necessary for plastic deformation relative to that
which would otherwise be necessary if the planes of atoms continued to slide over each
other. Therefore, reducing the slip-plane motion increases the material strength.
The Hall-Petch equation provides a correlation of yield stress (σy), and the grain size
(d), (Hall, 1951; Petch, 1953):
σ y = σ0 + Kd −1 / 2 (2)

where:
σ0 = the friction stress due to the grain boundary
K = grain boundary hardening constant that is independent of the temperature.
If the grain size is small, there will be fewer dislocations in the pile-ups over these
small distances until an obstacle (the grain boundary) is encountered. A greater applied
stress is necessary to achieve plastic deformation resulting in greater yield and
tensile strengths. As the grain size increases, plastic deformation increases more
rapidly and yield strength decreases. Ductility improves with decreasing grain size
(Taleff et al., 2002).
It is known that strengthening (yield strength) is determined by the average lamellae
spacing and the drawability and also die life improves as the lamellae spacing (s)
increases which also follows a Hall-Petch relationship (Taleff et al., 2002; Hata
et al., 2004).
Barthold (1996) reported that although CrSi spring steel, for example, exhibits
excellent fatigue properties it is does not contain microalloying elements to allow
the formation of dispersed precipitates which would inhibit grain growth resulting in
grain coarsening during heat treatment. The coarser grain structure reduces tensile
strength making the steel susceptible to internal cracks that will initiate fatigue fractures
in use.
Fatigue life is independent of grain size under low cycle fatigue conditions (high
loads, elastic and plastic deformation) and fatigue life is increased when grain size is
reduced under high cycle fatigue conditions (low load, elastic deformation). The
dependence of fatigue life on grain size is also dependent on the mode of deformation.
Grain size has its greatest effect on fatigue life in the low-stress, high-cycle regime in
which stage 1 cracking predominates. Kawabe et al. (2002) have shown that fatigue life
increase is proportional to grain diameter as shown in Figure 15.
286 L.C.F. Canale, R.N. Penha, G.E. Totten, A.C. Canale and M.R. Gasparini

Figure 15 Correlation between grain size and fatigue limit of an experimental spring steel
(1 × 107 cycles)

Retained austenite
When steel is hardened, it is first austenitised into a face-centred cubic (fcc) austenitic
microstructure then quenched and tempered. During quenching, austenite will transform
expanding into a hard and brittle body-centred tetragonal (bct) martensitic microstructure.
The temperature where the martensitic transformation starts may be calculated from the
chemical composition of the steel (Andrews, 1965): The equation used by Kumar et al.
(2000) and others (Mukhopadhyay et al., 1997) is:
M S (°C) = 561 − 474(%C) − 33(%Mn) − 17(%Ni) − 17(%Cr) − 21(%Mo) . (3)

The temperature where the martensitic transformation is completed is designated as the


Mf temperature.
After quenching, tempering is then performed to produce a tempered martensite
structure which is finer grained, less brittle but tougher with lower hardness than
as-quenched martensite. However, 100% of the austenitic structure may not transform to
martensite thus leaving some residual (retained) austenite in the martensitic structure. As
the MS approaches ambient temperature (and below) the amount of retained austenite
increases. Retained austenite may transform into untempered martensite during use,
which will lead to dimensional instability of the steel. One method of completing the
conversion of retained austenite into as-quenched martensite is by cryogenic treatment at
–70ºC. This is necessary because the presence of retained austenite lowers resistance to
small plastic deformation and increases the rate of relaxation processes after aging
(Rakhshtadt, 1975).
Overview of factors contributing to steel spring performance and failure 287

Stress relieving
Microstructure development and formation during hot-rolling and cold-drawing control,
mechanical properties, such as tensile strength and toughness. Tensile strength increases
exponentially with drawing strain (Elices, 2004). With increasing drawing strain, the
pearlitic lamellae reorients leading to the development of increased dislocations
(interlamellar spacing refinement) which are an important source of wire strengthening.
Alloy strengthening (solid-solution strengthening) occurs as the volume fraction of
cementite in the pearlitic microstructure increases (see Figure 16). Fine dislocation cells
that form in the ferrite of heavely deformed pearlitic lamella is another source of
strengthening as is strengthening that may result from the colony size of the pearlite
microstructure (Elices, 2004).

Figure 16 Volume fraction of ferrite/cementite interfaces as a function of drawing strain.

During the wire drawing process, large plastic deformation occurs. If the resulting strains
do not recover, residual stresses will be formed. The presence of residual stresses will
affect tensile properties by changing the stress-strain behaviour and therefore yielding
behaviour. This is illustrated in Figure 17(a), where the longitudinal residual stresses as a
function of depth from the surface, are shown for a cold–drawn eutectoid steel under due
conditions: as drawn, after for the drawing to obtain a 1% additional area reduction and
after heating to 400ºC under a load of 0.4 σmax. Figure 17(b) shows the stress-strain
curves of the eutectoid steel wire after these treatments. The results of this work showed
that as the residual stresses increase, yield stress decrease (Elices, 2004).
Stress relieving is typically used to remove residual stresses which have accumulated
from prior manufacturing processes. Stress relief is performed by heating to a
temperature below Ac1 and holding at that temperature for the required time to achieve
the desired reduction in residual stresses and then the steel is cooled at a rate to
sufficiently slow avoid formation of excessive thermal stresses. No microstructural
changes occur during stress relief processing (Grenier et al., 2003).
288 L.C.F. Canale, R.N. Penha, G.E. Totten, A.C. Canale and M.R. Gasparini

Figure 17 (a) Longitudinal residual stresses as a function of depth from the surface, of a
cold-drawn eutectoid steel under different conditions; (b) Stress-strain curves of the
eutectoid steel wire after different treatments

(b)

(a)

Stress relieving temperatures are above the recrystallisation temperature. Little or no


stress relief occurs at temperatures <260ºC and approximately 90% of the stress is
relieved at 540ºC. The maximum temperature for stress relief is limited to 30ºC below
the tempering temperature used after quenching. Stress relieving results in a
significant reduction of yield strength in addition to reducing the residual stresses
to some ‘safe’ value and crack-sensitive materials. Typically, stress-relieving times for
specific alloys are obtained from an appropriate industry standard for the spring steel
of interest.
When austenitic stainless steels have been cold worked to develop high strength, low
temperature stress relieving will increase the proportional limit and yield strength
(particularly compressive yield strength). This is a common practice for austenitic
stainless steel spring wire. A two hour treatment at 345ºC to 400ºC is normally used;
temperatures up to 425ºC may be used if resistance to intergranular corrosion is not
required for the application. Higher temperatures will reduce strength and sensitise the
metal, and generally are not used for stress relieving cold worked products.
One expression that may be used in evaluating the stress relief-time of spring steels is
the Larson-Miller equation (Grenier et al., 2003):

P = T(C + log t) × 10 −3 (4)

where t is the time (hours) at temperature T (in Kelvin) and the value for the composition-
dependent constant C is calculated from:
C = 21.3 − (5.8 × %Carbon). (5)

Table 6 provides a summary of typical stress relieving times for various spring steels.
However, more recent work has shown that rapid heating can be used to significantly
reduce these times and still obtain superior residual stress profiles and more uniform
properties (Grenier et al., 2003).
Overview of factors contributing to steel spring performance and failure 289

Table 6 Stress relief temperatures and times for wire

Temperature Time
Material Specifications (ºC) (minutes)
Music Wire ASTM A 228 232 30
Music Wire–tin-coated ASTM A 228 150 30
Music Wire Cadmium-Zinc Coated ASTM A 228 204 30
Music Wire AMS 5112 282 60
O.T.M.B. ASTM A 229 232 30
H.D.M.B. Class I or II ASTM A 227 232 30
High Tensile Hard Draw ASTM A 679 232 30
Galvanised M.B. Class I or II ASTM A 674 232 30
Chrome-Silicon SAE J157 or ASTM A 401 371 60
Chrome-Silicon (Lifens) SAE J157 385 60
Chrome-Vanadium ASTM A 231 371 60
Stainless Steel 301 – 343 30
Stainless Steel 302 AMS 5688 343 30
Stainless Steel 304 ASTM A 313 343 30
Stainless Steel 316 ASTM A 313 315 60
17–7 PH AMS 5678 482 60
Phosphorus Bronze Grade A ASTM B 159 190 30
Hasteloy C – 260 30
Monel 400 – 329 60
Inconel 600 – 454 90
Inconel × 700 Spring Temper AMS 5699 650 240
Inconel × 750/1 Temper AMS 5698 – –
Brass Wire ASTM B 134 190 30
Berilium-Copper ASTM B 134 or ASTM B 197 315 120
Blue-Temper N/A 232 30

In another study, stress relaxation profiles were determined for wire in the condition
as-drawn and after stress-relieving. The results are shown in Figure 18. There is a
relationship between stress relaxation and the magnitude of residual stresses and that the
stress relaxation loss increases with increasing surface tensile stresses (Elices, 2004).
Finally, Elices examined the effect of residual stresses on fatigue life. The results
showed that although there is an effect on fatigue life when the residual stresses are
high. In addition, residual stresses exhibit a large effect on the fatigue threshold
(Elices, 2004).
290 L.C.F. Canale, R.N. Penha, G.E. Totten, A.C. Canale and M.R. Gasparini

Figure 18 Profiles of longitudinal stresses when the wire is stressed at different loads as a function
of the relative depth (σm = 1940 MPa)

When wire is coiled into a compression spring, the steel on the inside of the coil is upset
and becomes shorter due to plastic deformation and the residual stresses on the inside of
the spring reduce the fatigue strength (Muhr, 1993). These residual stresses are decreased
by a stress relieving process. The magnitude of the residual stresses that are formed is
dependent on the tensile strength of the steel wire. Muhr provided a comparison of the
amount of stress relief after coiling as a function of temperature for a CrV and CrSi wire.
Typically, the CrSi wire exhibited approximately 40% greater residual stresses than the
CrV wire as shown in Figure 19 (Muhr, 1993). Figure 20 compares the residual stress as
function of distance from the surface for CrV and CrSi wire after stress relieving (Muhr,
1993). It was also reported that if the negative effects of residual stresses due to coiling
are eliminated, increased residual tensile stresses resulted in improved fatigue strength
(Muhr, 1993).

Figure 19 Residual stresses after stress relieving of CrSi and CrV alloyed valve springs
Overview of factors contributing to steel spring performance and failure 291

Figure 20 Residual stress curves of valve springs of CrV and CrSi alloyed wire, which have been
subject to different heat treatment

Distance from the wire surface [mm]

Tempering
When steel is hardened, the as-quenched martensite is not only very hard but also brittle.
Tempering, also known as ‘drawing’, is the thermal treatment of hardened and
normalised steels to obtain the desired mechanical properties which include: improved
toughness and ductility, lower hardness and improved dimensional stability. During
tempering, microstructure modifications (carbide decomposition and martensite
alterations) allow hardness to decrease to the desired level. The extent of the tempering
effect is determined by the temperature and time of the process. The Larsen Miller
equation shown previously can also be used to predict tempering time to achieve an
as-tempered property such as hardness (Jarl et al., 2003; Grenier et al., 2003). The
tempering process may be conducted at any temperature up to the lower critical
temperature (Ac1).
The specific tempering conditions that are selected are dependant on the desired
strength and toughness. Improper tempering times may result in serious breakage
problems. Segregation, particularly of chromium and molybdenum may occur which will
result in the core being more temper-resistant than the surface. Kern and Suess (1979)
report that the effect of tempering can be evaluated by heating a steel bar for 0.02 h/mm
at the tempering time of interest and a hardness reduction of greater than two HRC units
anywhere along the bar indicate an inadequate tempering time. Properly manufactured
and heat treated steel for leaf springs should exhibit a V-notch impact strength at –29ºC
292 L.C.F. Canale, R.N. Penha, G.E. Totten, A.C. Canale and M.R. Gasparini

of 6 ft-lb min at a HRC of 42. Increasing the tempering temperature reduces fatigue
strength as shown in Figure 21 (Kawabe et al., 2002). This is probably due to the higher
tensile strength achieved at lower tempering temperatures.

Figure 21 Correlation of tempering temperature and fatigue limit for an experimental spring steel
(1 × 107 cycles)

Shavrin and Red’kin (1974) performed a similar study and obtained the opposite results
using a different heat treatment process with a Russian 50KhFA spring steel (0.53% C,
0.65% Mn, 0.17% Cu, 1.06% Cr, 0.29% Si, 0.18% Ni, 0.17% V, and 0.014% P) which is
used for firing pins and ejector springs for hunting rifles. The springs were heat treated
after machining and cold-bending. Heat treatment was performed in an electric muffle
furnace with a charcoal protective atmosphere. The springs were oil quenched after
hot-bending and the interval between the two processes did not exceed 5–6 s. After
quenching, the springs were tempered in a molten salt bath. For the desired time for
20 min and then cooled with running tap water After heat treatment, the springs
were tested in an eccentric-cam testing bench with low-cycle fatigue was determined
using a cyclic loading at a constant sign. The results obtained from this study are shown
in Figure 22 (Shavrin and Red’kin, 1974).
Potentially destructive residual stresses may remain in steels after heat treatment.
These stresses can be relieved by stress tempering (not to be confused with stress
relieving) which is conducted at a temperature that will not affect the heat treated
hardness. One method for doing this is to raise the temperature to about 30ºC below the
last tempering temperature and then hold for 6 h. The steel is then cooled in still air.
Stress tempering improves the elastic properties of steel by relaxing residual
microstresses that remain after martensitic transformation. The nature and distribution of
these microstresses strongly affect the endurance of steel under cyclic loading conditions.
By reducing these microstresses, the fatigue resistance of parts such as helical coil steel
springs may be improved.
Overview of factors contributing to steel spring performance and failure 293

Figure 22 Effect of tempering temperature on low-cycle fatigue strength with a deformation


temperature of 870ºC

Notes: Tempering at (1) 240ºC, (2) 320ºC, (3) 400ºC, (4) commercial production
conditions of bending at 700ºC–750ºC and cooling in air.
Heat treatment at 850ºC and tempering at 350ºC.

The affect of stress tempering on the mechanical properties of 60S2A (0.62% C, 1.73%
Si) and 50 KhFA (typical chemistry shown above) spring wire and springs has been
studied (Zabil’skii et al., 1978). The results of this work showed:
• Stress tempering of spring wire made of steels 60S2A and 50KhFA by the optimum
regimes improves the proportional limit and the yield point by 15–25 kgf/mm2.
Additional plastic strain of 0.3%–0.5% combined with stress tempering improves
these characteristics by 30–40 kgf/mm2.
• The improved mechanical properties (τpr and τy of steel 50KhFA after stress
tempering at 250ºC under tests at elevated temperatures are maintained at the same
level at least to 200ºC.
• The endurance of wire and springs under cyclic loading as a result of stress
tempering increases several times over; the residual set of the springs during fatigue
tests is reduced to one-third.
Intergranular fracture along prior austenite grain boundaries may lead to either tempered
martensite embrittlement or temper embrittlement. Either process may lead to decreased
fatigue strength (Fyfitch et al., 2001). Tempered martensite embrittlement may
potentially occur during tempering at approximately 350ºC and temper embrittlement
may occur during tempering at approximately 500ºC. Temper embrittlement is related to
grain boundary segregation of impurity elements in the steel. Tempered martensite
embrittlement is thought to result from the combined effects of cementite precipitation on
prior-austenite grain boundaries or interlath boundaries and the segregation of impurities
at prior-austenite grain boundaries. In both cases, hardness and strength decrease
294 L.C.F. Canale, R.N. Penha, G.E. Totten, A.C. Canale and M.R. Gasparini

continuously with increasing tempering times (Reguly et al., 2004). Temper


embrittlement can be reversed by retempering above the critical temperature range, then
cooling rapidly.

Quenching
Quenching involves heating a steel above the upper critical temperature (about 760ºC),
the iron crystal structure will change to Face Centred Cubic (FCC) structure called
austentite (γ-iron) and the carbon atoms will migrate into the central position formerly
occupied by an iron atom. This is called the ‘austenitising temperature’ and the selection
of the proper austenitising temperature depends on the steel chemistry. The steel is held
at the austenitising temperature for a sufficient time until all of the microstructure has
transformed to austenite. The soaking time in air furnaces should be 1.2 min for each mm
of cross-section or 0.6 min in salt or lead baths. Non-uniform heating, overheating and
excessive scaling should be avoided.
After austenitisation, the red-hot steel is then quenched by immersion into the
appropriate medium such as: air, oil or water depending on the chemical composition of
the steel. If the steel is cooled too slowly, the iron atoms move back into the cube forcing
the carbon atoms out, resulting in soft steel called pearlite. If the sample was formerly
hard, this softening process is called annealing. If the steel is quenched quickly, the
carbon atoms are trapped, and the result is a very hard, brittle steel. This steel crystal
structure is now a Body Centered Tetragonal (BCT) form called martensite.
Quenching is necessary to suppress the normal breakdown of austenite into ferrite and
cementite, and to cause a partial decomposition at a lower temperature to form
martensite. To obtain this, steel requires a critical cooling velocity. The critical cooling
velocity for alloy steels is much slower for more hardenable alloy steels than for carbon
steels. Alloy steels are typically oil hardened (or aqueous polymer quenchants may, in
some cases, be used).
The quenching process should be performed sufficiently fast to achieve the maximum
transformation to martensite and it should continue until the temperature of the steel is
less than the martensite-finish (Mf) temperature to minimise the presence of retained
austenite. The as-quenched surface hardness (below and decarburised layer) should
correspond to 99.9% martensite as indicated in Figure 23 (Kern and Suess, 1979).

Figure 23 Hardness of a steel at 99.9% martensite as a function of carbon content


Overview of factors contributing to steel spring performance and failure 295

The difference in the cooling rate between the surface and the core of the steel leads to
the development of residual stresses. If these stresses are greater than the yield and
fracture stresses for the steel, quench cracking may result.
To assess the hardenability of a steel, it is useful to determine the ideal critical
diameter (DI) which is defined as the diameter that will produce an as-quenched
microstructure of 50% martensite. Hardenabilty is determined by: grain size, carbon
content and alloy composition. (Hardenabilty the ability to harden steel, must not be
confused with hardness, which is dependent on the carbon content of the steel as shown
in Figure 18.) The DI can be calculated from:
D I = D I(base) (carbon and grain size) × fMn × fSi × fCr × fMo × fV × fCu . (6)

The values for DI (base) and alloy factors (fx) are available in tabular form in metallurgy and
heat treatment texts.
In addition to DI, another value related to steel hardenability is also helpful which is
referred to as the critical diameter (DC) which is defined as that diameter in which
50% martenite is formed in a particular quenching medium. For a leaf spring the critical
values corresponding to DI is Lt (ideal plate thickness) and DC is LC (critical plate
thickness). Thus for a leaf spring, for example, the thickness of the spring must be
comparable to the critical plate thickness (LC) if excessive stresses are to be avoided
(Mukhopadhyay et al., 1997).
It is possible to develop a qualitative expression to estimate if the internal stress of the
steel is in excess of the fracture stress which is required to cause quench crack. This is
done by first estimating the surface stress (σsurface) from (Mukhopadhyay et al., 1997):
σsurface = σ T + σS (7)

where:
σT = the thermal stress
σS = the structural stress.
The thermal stress is due to the non-uniform temperature temperature between the surface
(TS) and the core (TC) and can be calculated from:
σ T = Eα (TC − TS ) (8)

where:
E = Young’s modulus
α = the thermal expansion coefficient.
The value of σS is dependent on the martensitic transformation occurring at the centre
and surface of the component. The most critical stage is the amount of transformation in
the centre and at the surface after the quench process is complete. At this point, σS can be
qualitatively approximated from:
σS = EεfS [1 − 2(fS − fC )] (9)

where:
ε = (1/3)(∆V/V) which is the strain due to martensitic phase transformation
f = the fraction of transformation and which fS = 1 and fS = 0 otherwise.
296 L.C.F. Canale, R.N. Penha, G.E. Totten, A.C. Canale and M.R. Gasparini

At the centre, fC = 1 if martensite if complete transformation has occurred at the centre or


fC = 0 otherwise.
Taken together, the value for σsurface can be approximated from:
σsurface = Eα(TC − TS ) + EεfS [1 − 2(fS − fC )]. (10)

For thin cross-sections as for many springs, fS = fC = 1. In this case, where


∆T = (TC – TS), the equation may be simplified to (Mukhopadhyay et al., 1997):
σsurface = Eα(∆T) + Eε. (11)

Baggerly (1994) has reported that quench cracking can be exacerbated by decarburisation
of the steel which leads to cracking caused by unfavourable residual stresses caused by a
raid quenching process.
Kayushnikov (1963) has reported that quench distortion control of coil springs
manufactured from U9A Russian steel could be accomplished by heating the springs in a
horizontal position and attaching them at several points to the bottom of the basket then
immersion quenching in oil. Distortion control was achieved by tempering at 320ºC to
remove the distortion due to quenching.
Quench embrittlement, the susceptibility to intergranular fracture in as-quenched
low-temperature tempered high-carbon steel due to cementite formation on austenite
grain boundaries during austenitisation, was examined by Reguly et al. (2004)
for AISI 5160 spring steel. Their results showed that quench embrittlement was
affected by elements such as manganese but especially phosphorous segregation
on the austenite grain boundaries at all temperatures. Phosphorus reduces carbon
solubility in the steel. The relatively high 0.6% carbon content makes this alloy
particularly susceptible to quench embrittlement. If the phosphorus content of 5160 steel
is reduced to very low levels, quench embrittlement can be suppressed by austenitising
at 830ºC–920ºC.

2.4 Corrosion
Corrosion and corrosion-control processes may lead to the diffusion of hydrogen into a
component which can seriously reduce the ductility and load-bearing capacity of the steel
and lead to cracking and brittle fracture failures at stresses below the yield stress of the
steel (Fyfitch et al., 2001). Two common parts of the mechanism of hydrogen
embrittlement are: application of a tensile stress and hydrogen dissolution in the metal.
Hydrogen embrittlement does not affect all steels equally. The most vulnerable are
high-strength steels, titanium alloys and aluminium alloys.
Sources of hydrogen causing embrittlement for springs include: steelmaking and as a
by-product of general corrosion reactions such as rusting, cathodic processes such as
galvanic surface treatments, and electroplating. Hydrogen diffuses along the grain
boundaries and combines with carbon in the steel to form methane gas which
concentrates in small voids along the grain boundaries where it builds up enormous
pressures that initiate cracks. If the steel is under a high tensile stress, brittle failure
can occur.
At normal room temperatures, the hydrogen atoms are absorbed into the metal lattice
and diffuses through the grains, tending to gather at inclusions or other lattice defects
shown schematically in Figure 24 (Barthold, 1998a). Cracking of martensitic and
Overview of factors contributing to steel spring performance and failure 297

precipitation-hardened (PH) steel alloys is believed to be a form of hydrogen stress


corrosion cracking that results from the entry of a portion of the atomic hydrogen that is
produced in the corrosion reaction into the metal. Hydrogen embrittlement is typified by
open grain boundaries, microcavities, and ductile hairlines in the fracture surface
(Barthold, 1998a).

Figure 24 Illustration of the mechanism of damage caused by hydrogen diffusion into the
steel matrix

To reduce hydrogen embrittlement, the amount of residual hydrogen in steel must be


controlled, by controlling the amount of hydrogen pickup in processing, developing
alloys with improved resistance to hydrogen embrittlement, the development developing
low or no embrittlement plating or coating processes, and restricting the amount of in situ
(in position) hydrogen introduced during the service life of a part.
Interestingly, Lantsman et al. (1973) reported that there was no direct correlation
between the amount of hydrogen absorbed in the steel and the tendency for delayed
fracture. They found that the susceptibility to hydrogen embrittlement was largely
dependent on the heat treatment and microstructure, which contribute to the internal
stresses of the steel. The stresses due to hydrogen embrittlement are then added to the
pre-existing internal stresses. The critical factor is total stresses. Clearly, if the spring can
be produced with lower stresses, the susceptibility to fracture due to hydrogen
embrittlement will be reduced. For example, if the steel is austempered rather than
conventionally quenched, it would be expected to be less susceptible to hydrogen
embrittlement which is what Lantsman et al. observed.
Another source of cracking due to corrosion that can lead to fatigue failures is pitting.
Figure 25 illustrates a fatigue crack of a coil spring that was originated from a pit caused
by corrosion.
298 L.C.F. Canale, R.N. Penha, G.E. Totten, A.C. Canale and M.R. Gasparini

Figure 25 Illustration of pitting due to corrosion of a spring used as a vibration isolator for a motor
in a pharmaceutical plant

Source: Courtesy of Neville W. Sachs, P.E., Sachs, Salvaterra & Associates,


Inc., Syracuse, NY

2.5 Shot peening


Shot peening, as it relates to spring manufacture, is a controlled process of blasting a
large number of hardened spherical or nearly spherical particles (shot) against the softer
surface of the steel spring. Each impingement of a shot makes a small indentation in the
surface of the part, thereby inducing compressive residual stresses. This process is
conducted to improve the resistance of the spring to fatigue fracture or stress-corrosion
cracking (arising from hydrogen embrittlement).
In a very early study, Lessells and Murray (1941) showed that shot peening increased
the fatigue strength of steel by increasing the surface compressive stresses of the
component. This has been an active area of spring steel research since that time. For
example, more recently, Tekeli (2002) studied the use of shot peening to increase the
fatigue life of AISI 9245 steel used. It was shown that shot peening increased the fatigue
life of this steel by about 30%.
Nakhimov et al. (1963) examined the use of shot blasting to increase the compressive
stresses of a 55KhGR (0.57% C, 0.36% Si, 1.14% Cr, 0.5% Ni, 1.03% Mn, 0.0037% S)
spring steel after heat treatment, quenching and tempering. As a result of this work, it was
shown that:
• oil quenching creates surface tensile stresses which were nearly completely relieved
by tempering
• cooling with water after tempering has minimal effect on surface stresses
Overview of factors contributing to steel spring performance and failure 299

• The magnitude of thermal stresses depends on the microstructure, and hardness


of the steel. As hardness decreases, thermal stresses rise sharply. (This is consistent
with work by Berns et al. (1985) and Gao et al. (2002) who showed that the
surface residual stresses were dependent on both material properties and shot
peening parameters.
• shot blasting creates considerable surface compressive stresses
• Tempering after shot blasting causes only minimal reduction of surface stresses and
the fatigue properties also change very little. There is also a further improvement in
corrosion resistance.
Naisai et al. (1987) examined the effect of operational parameters on the shot peening of
3 mm 0.65%–0.70% carbon steel wire used for coil spring manufacture. Their work
showed that compressive stress exhibits a greater effect than hardness on fatigue strength
in the range of Rc = 40 – 50. Fatigue failure originates with surface defects for unpeened
wire. For shot peened wire, fatigue failure originates below the surface. When this
occurs, fatigue strength can be increased by increasing the depth of compressive stresses
by adjusting the shot peening parameters. The magnitude of the peak values of
compressive stress are less important.
It has been shown that it is not necessary to measure the compressive residual stresses
to estimate fatigue life improvement since it is possible to correlate fatigue life with a
damping factor (Aggarwal et al., 2005). Determination of a damping factor takes only a
few minutes versus a few hours of x-ray analysis time. The damping factor (ξ) can be
calculated from the logarithmic decrement (δ):
δ = 1 / N ln X1 / X N (12)

δ = 2πξ /(1 − ξ2 )1 / 2 (13)

where:
N = number of cycles
X1 = amplitude of vibration for the first cycle
XN = amplitude of vibration after the N-th cycle.
Decarburisation is a common cause for fatigue life problems of spring steels. (see above
discussion). Sato et al. reported that shot peening can substantially improve the fatigue
strength of SUP6 (0.55%–0.65% C, 1.50%–1.80% Si, 0.7%–1.0% Mn, <0.055% P,
<0.035% S) spring steel plate. As Figure 26 shows, although shot peening provides a
substantial improvement in fatigue life, the limit of improving fatigue life for this steel
and shot peening conditions is when total depth of decarburisation is <0.15 mm for this
steel (Sato et al., 1981).
Todinov (2000) also studied the effect of shot peening on quenched and tempered
spring wire (60Si2) which possessed a decarburised surface layer. Interestingly, he found
that oil quenched steel had tensile surface stresses while water exhibited compressive
surface stresses and that the residual stresses do not disappear after tempering. More
importantly, if the shot peening process is not conducted properly small compressive
stresses or even tensile stresses at the surface will result dramatically reducing the fatigue
life of the spring.
300 L.C.F. Canale, R.N. Penha, G.E. Totten, A.C. Canale and M.R. Gasparini

Figure 26 Correlation between total depth of decarburisation (DM-T) and fatigue life

Notes: Shot peening condition (2) is: 1 pass with 0.7 mm steel shot (ss) at 87 Kg.min,
wheel speed 51.3 m/s and conveyor speed of 3 m/min
Shot peening condition (5) is: 1 pass with 0.7 mm steel shot (ss) at 87 Kg.min,
wheel speed 51.3 m/s and conveyor speed of 3 m/min followed by one pass
0.4 mm cut-wire (cw) at 95 Kg.min, wheel speed 56.4 m/s and conveyor speed
of 3 m/min

2.6 Fatigue failure


There are three common failure mechanisms in leaf and coil springs that lead to fatigue
failure that are readily characterisable microscopically. Two sources of fatigue cracking
are those due to tension such as Figures 27(a) and 27(b); and those due to steel inclusion
such as that illustrated in Figure 28 (Hayes, 1984). A third source of fatigue cracking has
been reported as due to shear cracks (Hayes, 1984). Fatigue strength is typically
improved by shot peening (Xu et al., 1982; Hirose, 1982). Table 7 provides a summary of
the frequency that these types are failures are observed (Hayes, 1984).

Figure 27 Two examples of tension crack formation

(a) (b)
Source: Courtesy of Kieselstein GmbH, Chemnitz, Germany
Overview of factors contributing to steel spring performance and failure 301

Figure 28 Failure due to steel inclusion

Source: Courtesy of Neville W. Sachs, P.E., Sachs, Salvaterra & Associates,


Inc., Syracuse, NY

Table 7 Summary of failure-types for unpeened and shot peened spring wire

Failure type (%)


Wire type Shear Tensile Subsurface
BS5216
As-coiled 75 25 0
Shot peened 25 75 0
BS2056
As-coiled 100 0 0
Shot peened 0 100 0
BS2803C
As-coiled 70 30 0
Shot peened 5 95 0
CrV
As-coiled 100 0 0
Shot peened 5 90 5
SiCr
As-coiled 95 5 0
Shot peened 5 50 45

Fatigue failure may be initiated by the presence of a defect which may act as a stress
raiser or point defect (Beretta and Boniardi, 1999). Fatigue strength is related to the
surface quality (smoothness) of a material. Interestingly, particularly with high cycle
fatigue, cracks actually appear under the surface (subsurface) and the failure process is
called ‘incipient subsurface failure’ (Beretta and Boniardi, 1999; Teretév, 2004). In
order to obtain higher quality materials, an increase in frequency of online testing
(Zoughi et al., 1997) and even for subsurface defects is recommended (Duncan and
Bashkansky, 1998).
The effect of microstructure on the fatigue properties of high-strength spring wire
(0.59% C, 2.49% Si, 0.50% Mn, 0.54% Cr. 1.98% Ni, 0.18% V) was investigated and the
following observations were made (Lee et al., 1998):
302 L.C.F. Canale, R.N. Penha, G.E. Totten, A.C. Canale and M.R. Gasparini

• As the tempering temperature increases from 300ºC–450ºC, the precipitation


evolves from VC to (VC + ε-carbide + (Fe, Cr, V)3C) and the tempered
martensite microstructure changes from a retained austenite-containing mixture
to a lath-type tempered martensite. These two structures contribute to
slip deformation.
• Fatigue endurance increases with increasing temperature with a maximum at 450ºC.
• With further increases in tempering temperature, the martensite matrix is softened
due to excessive precipitation resulting in decreased fatigue endurance.
In the automotive industry, it is recognised that effective life of springs may be affected
by yielding, not fracture. Therefore, the elastic limit is design parameter. The value of
(σf)2/E is used for material selection and manufacturing (Yan, 1998): The value σf
is the stress to fracture and E is Young’s modulus. Typically, the greater this value, the
better. Another design parameter is fatigue strength and maximum endurance limits
are desired.
Various approaches to quantifying the effects of small cracks, defects, non-metallic
inclusions on fatigue life have been reviewed by Murakami and Beretta (1999). Timmins
(1979) has reported that a fracture mechanics approach to the fatigue failure of springs
offers a number of advantages relative to conventional fatigue testing since the important
problem of fatigue crack propagation can be addressed. Fatigue testing provides the
number of cycles at which failure occurs. It does not provide information on crack
nucleation or the growth of a crack prior to final failure. The reader is directed to
Timmins (1979) for further information.
Automotive leaf springs, for example, may progressively sag or deform over the time
that they are in service. Sag is caused by a cyclic microcreep and/or plastic deformation
occurring over a large numbers of loading cycles at stress levels less than the yield
strength of the material. The ability of a material to resist deformation, also known as
‘sag resistance’, results from both static and dynamic loading during use and is
logarithmically dependent on time.
Sag resistance can be predicted using the Bauschinger effect which is the weakening
phenomenon of a material when it is deformed in the reverse direction after a plastic
strain is imposed in the forward direction (Yan, 1998). Sag resistance is estimated using a
Bauschinger torsion test which measures the weakening of a metal when strained in the
reverse direction after a plastic strain in a forward direction is applied. The ‘actual
Bauschinger twist’ – ABT – measures the different in strain in the forward-twist and
reverse-twist curve at maximum torgue as illustrated schematically in Figure 29 (Yan,
1998). The greater the Bauschinger effect, the greater the sag resistance (resilience).
Using known Bauschinger effect data, it is possible to determine the load loss or sag
resistance of spring steels. Yang and Wang (1996) examined the effect of prestain on the
cyclic creep behaviour of a high-strength spring steel and they found that when prestain
reached a critical value, compressive cyclic creep was observed that was attributed to a
strong Bauschinger effect. However, it is important to note the mechanisms of the sag
effect and the Bauschinger effect are not the same since the Bauschinger effect is the
direction-dependent, asymmetrical yield behaviour which measured after the first few
cycles of loading and sag resistance occurs over thousands of normally elastic stresses
(see Figure 30).
Overview of factors contributing to steel spring performance and failure 303

Figure 29 Schematic illustration of the Bauschinger torsion test

Figure 30 The Bauschinger effect is the uniaxial stress-strain behaviour obtain by forward and
reverse flow tests
304 L.C.F. Canale, R.N. Penha, G.E. Totten, A.C. Canale and M.R. Gasparini

The effects of Mo, W, and V additions on the hysteresis loop of the Bauschinger test was
determined to predict sag resistance of Si-Cr spring steels (SAE 9254). This work showed
(Nam et al., 2000):
• The sag-resistance of Si-Cr steels is directly influenced by the distribution
of precipitates.
• Sag-resistance increases with tempering temperature and then decreases at tempering
temperatures greater than 350ºC.
• The addition of Mo and/or W does not affect the behaviour of tempered carbides
at low tempering temperatures. However, the rate of spheroidisation and
cementite coarsening is reduced at tempering temperatures >400ºC and
sag-resistance is improved.
• It was recommended that the ratio of loop area in the Bauschinger test to
hardness be used as a spring steel design parameter to estimate sag-resistance
relative to hardness.
• Vanadium additions improve sag-resistance due to precipitation of vanadium
carbides at all tempering temperatures.

Murai and Takenoshita (1989) also used the Bauchinger effect to study the sag-resistance
of Si-Cr spring steels. They were concerned with problems encountered when high-
strength steel is used for improved sag-resistance such as damaged cutting tools and
decreased shot peening effect. In their work, Si-Cr (SAE 9254) steel was quenched and
tempered and then a strain was imposed prior to cold-forming the coil spring. These
lower strength springs exhibited high sag-resistance equivalent to coil springs
manufactured from high-strength steel without cutting tool damage. Methods of imposing
a strain on steel wire include: tortion, drawing, and cyclic bending.

3 Conclusion

Various factors that are involved in coil and leaf spring failure have been discussed. Of
these, four factors are predominately involved in spring failures. In addition to steel
chemistry, the use of clean steels with minimal inclusion content (Superclean steels) is
particularly important if fatigue life is to be maximised. Surface defects not only
including decarburisation but the presence of laps, seams, and other defects also lead to
premature spring failure. To address these problems, high-quality outer layer of spring
wire by a grinding or a draw-peeling (shaving) process. Finally, there is an increased
emphasis on spring design innovations, which can be used with higher loads at reduced
spring weight and size with substantial improvements in fatigue strength.

References
Aggarwal, M.L., Khan, R.A. and Agrawal, V.P. (2005) ‘Smart system for measurement of
compressive residual stress field in shot peened components’, Proceedings of International
Conference on Smart Materials, Structures and Systems, 28–30 July, Bangalore, India,
Paper No. ISSS-2005/PS-15, pp.PS–91–PS–98.
Andrews, K.W. (1965) ‘Empirical formulae for the calculation of some transformation
temperatures’, The Journal of the Iron and Steel Institute, Vol. 203, pp.721–727.
Overview of factors contributing to steel spring performance and failure 305

Ashida, S., Ibaraki, N. and Mizutani, K. (1993) Wire Rod for High Strength and High Toughness
Fine Steel Wire, Twisted Products Using Fine Steel Wires and Manufacture of the Fine Steel
Wire, U.S. Patent 5,211,772, 18 May.
Ayada, M., Yuga, M., Tsuji, N., Saito, Y. and Yoneguti, A. (2006) ‘Effect of vanadium and
niobium on restauration behavior after hot deformation in medium carbon spring steels’,
Journal of Materials Processing Technology, Vol. 171, pp.232–239.
Baggerly, R.G. (1994) ‘Quench cracks in truck spring “U” bolts and the implications for spring
failure’, Engineering Failure Analysis, Vol. 1, No. 2, pp.135–141.
Bannister, J.L. (1953) ‘Cold drawn prestressing wire’, The Structural Engineer, August,
pp.203–218.
Barthold, G. (1996) Recent Developments in Spring Steel Materials, Springs, Vol. 35, No. 3,
pp.64–76.
Barthold, G. (1998a) Spring Failures and their Causes, Springs, Vol. 37, No. 3, pp.21–37.
Barthold, G. (1998b) ‘Recent developments in spring steel materials’, Wire, Vol. 45, No. 4,
pp.223–228.
Beretta, S. and Boniardi, M. (1999) ‘Fatigue strength and surface quality of eutectoid steel wires’,
Intl. Journal of Fatigue, Vol. 21, pp.329–335.
Berns, H., Siekmann, G., Kősters, R. and Schreiber, D. (1985) ‘Structure and rupture of heat treated
spring steels (fatigue dailure)’, Wire, Vol. 35, No. 1, pp.8–13.
Bertrand, C., Molinero, J., Londa, S., Elvira, R., Wild, M., Barthold, G., Valentin, P. and Shifferrl,
H. (2003) ‘Metallurgy of plastic inclusions to improve fatigue life of engineering steels’, Iron
Making and Steel Making, Vol. 30, No. 2, pp.165–169.
Birosca, S., Dingley, D. and Higginson, R.L. (2004) ‘Microstructural and microtextural
characterization of oxide scale and steel using electron backscatter diffraction’, Journal of
Microscopy, Part 3, Vol. 213, pp.235–240.
Dhua, S.K., Ray, A., Jha, S. and Chakravorty, S. (2004) ‘Metallurgical investigation onto the
causes of premature failure of steel wires and screws during manufacture’, J. of Failure
Analysis and Prevention, Vol. 4, No. 2, pp.17–23.
Duncan, M.D. and Bashkansky, M. (1998) ‘Subsurface defect detection in materials using optical
coherence tomography’, Optics Express, 22 June, Vol. 2, No. 13, pp.540–545.
Elices, M. (2004) ‘Influence of residual stresses in the performance of cold-dracon pearlitic wires’,
Journal of Materials Science, Vol. 39, pp.3889–3899.
Eriksson, L.H. (1995) Superclean Steel Improves Fatigue Properties of Valve Springs, Springs,
Vol. 34, No. 1, pp.41–55.
Frank, A.R. and Kirkaldy, A. (1988) ‘The effect of boron on the properties of electric arc-sourced
plain carbon wire-drawing qualities’, Wire Journal International, May, pp.100–103.
Funatani, K. (2004) ‘Low-temperature sale bath nitriding of steels’, Metal Science and Heat
Treatment, Vol. 46, Nos. 7–8, pp.277–281.
Fyfitch, S., Moore, K.E., Xu, H. and Hyres, J.W. (2001) ‘Failure investigation of type 17-7PH leaf
spring material’, Report available from Framatome ANP, Lynchburg, VA, USA,
http://www.us.framatome-anp.com/pdf/17-7PH-Xu-2556%20.PDF.
Gao, Y-K., Yao, M. and Li, J-K. (2002) ‘An analysis of residual stress fields caused by shot
peening’, Met. And Mat. Trans., Vol. 33A, pp.1775–1778.
Grenier, M., Canale, L.C.F., Crnkovic, O. and Totten, G.E. (2003) ‘Rapid tempering and stress
relief via high-speed convection heating’, Industrial Heating, Vol. 70, No. 5, pp.40–45.
Grimwade, M.F. (2006) Causes and Prevention of Defects in Wrougth Alloys, 15 June, World Gold
Council, International Technology, 45 Pall Mall, London SW1Y5JG, UK, www.ganoksin
.com./borisat/nenam/wrought-alloys.htm.
306 L.C.F. Canale, R.N. Penha, G.E. Totten, A.C. Canale and M.R. Gasparini

Hagiwara, T., Kawami, A., Ueno, A. and Kido, A. (1991) ‘Superclean steel for valve spring quality
wire’, Wire Journal International, April, pp.29–34.
Haldex Garphyttan, A.B. (2006) ‘Stainless precipitation spring wire with high-strength fatigue’,
Eurowire, 24–28 April, http://www.read-eurowire.com/haldex.cfm.
Hall, E.O. (1951) ‘The deformation and ageing of mild steel III: discussion of results’, Proc. Phys.
Soc. Ser. B., Vol. 64, pp.747–753.
Hata, H., Nagao, M. and Minamida, T. (2004) High-Carbon Steel Wire Rod with Superior
Drawability and Method for Production Thereof., U.S. Patent 6,783,609 B2, 31 August.
Hayes, M.P. (1984) ‘Physical appearance of spring fatigue failures’, The Metallurgical and
Materials Technologist, Vol. 16, No. 11, pp.572–573.
Hirose, S. (1982) ‘Effect of shot peening on fatigue strength of small spring wire and coil springs’,
1st International Conference on Shot-Peening, Oxford, UK: Pergamon Press, pp.51–59.
Isumida, H., Kawabe, N., Takamura, S., Morita, H. and Murai, T. (2005) ‘Development of
high-tensile-strength stainless steel wire’, SEI Technical Review, June, No. 60, pp.24–29.
Jarl, M., Berntsson, T., Sapcanin, E. and Azizi, B. (2003) ‘Replacement of lead baths for the
tempering of spring wire’, Scandanavian Journal of Metallurgy, Vol. 32, No. 5, pp.241–246.
Karimi Taheri, A., Maccagno, T.M. and Jonas, J.J. (1995) ‘Dynamic strain aging and wire drawing
of low carbon steel rods’, ISIJ International, Vol. 35, pp.1532–1540.
Kawabe, N. and Murai, T. (2003) High Fatigue –Strength Steel Wire and Spring and Processes for
producing These, U.S. Patent 6,627,005 B1, 30 September.
Kawabe, N., Izumida, H., Murai, T., Yamao, N., Matsumoto, S., Yamaguchi, K. and Fujimoto, S.
(2002) ‘Development of a highly fatigue-resistant oil-tempered wire for valve springs’, SEI
Technical Review, January, No. 53, pp.99–106, https://www.sei.co.jp/tr_e/t_technical
_e_pdf/53-14.pdf.
Kawabe, N., Murai, T., Yamaguchi, K. and Oishi, Y. (2003) Steel Wire and Method of
Manufacturing the Some, U.S. Patent 6,527,883 B1, 4 March.
Kawahara, J., Tanabe, K., Banno, T. and Yoshida, M. (1992) ‘Advance of value spring steel’, Wire
Journal International, November, pp.55–61.
Kayushnikov, P.Y. (1963) ‘Quenching without deformation’, Metal Science and Heat Treatment,
Vol. 5, No. 3, pp.149–152.
Kazeminezhad, M. and Taheri, A.K. (2003) ‘The effect of controlled cooling after hot rolling on
the mechanical properties of a commercial high carbon steel rod’, Materials and Design,
Vol. 24, pp.415–421.
Kern, R.F. and Suess, M.E. (1979) ‘Chapter 12 – Selection of materials for heat treated springs’,
Steel Selection: A Guide for Improving Performance and Profits, New York, NY: John Wiley
& Sons, pp.223–249.
Kieselstein, S. (2006) ‘Surface treatment of valve spring wires’, Produkt Informationen-wire 2006,
Wire Düsseldorf, International Fachmesse Draht and Kabel, 24–28 April, Düsseldorf,
Germany, kiesel, GmbH, Otto-Schmerbach-Strasse 19, D-09117, Chemnitz, Germany,
www.wire.de.
Kimura, T. and Kurebayshi, Y. (2001) ‘Niobium in microalloyed engineering steels, wire rods and
case carburized products’, Niobium Science and Technology. Proceedings of the International
Symposium Niobium 2001, TMS, Warrendale, PA, pp.801–819.
Krzyzanowski, M., Yang, W., Sellars, C.M. and Beynon, J.H. (2003) ‘Analysis of mechanical
descaling: experimental and modeling approach’, Materials Science and Technology, January,
Vol. 19, pp.109–116.
Kumar, B.R., Bhattacharya, D.K., Das, S.K. and Ghosh Chowdhury, S. (2000) ‘Premature fatigue
failure of a spring due to quench cracks’, Engineering Failure Analysis, Vol. 7, pp.377–384.
Kuroda, T., Ibaraki, N. and Yoshihara, N. (2002) Spring Steel Superior in Workability, U.S.
Patent 6, 372,056 B1, 16 April.
Overview of factors contributing to steel spring performance and failure 307

Kuźmiński, Z. and Nowakowski, A. (2001) ‘Plastometric evaluation of the properties of spring


steel during hot forming’, Metal 2001 Conference, 15–17 May, Ostrava, Czech Republic,
http://www.tanger.cz/metal2001/papers/44.pdf.
Lantsman, P.S., Vernovaskaya, G.G. and Kudryavtsev, V.N. (1973) ‘Use of austempering to reduce
hydrogen embrittlement of steel 65S2VA’, Metal Science and Heat Treatment, Vol. 15, No. 5,
pp.434–435.
Lee, C.S., Lee, K.A., Li, D.M., Yoo, S.J. and Nam, W.J. (1998) ‘Microstructural influence on
fatigue properties of a high-strength spring steel’, Mat. Sci and Eng., Vol. A241, pp.30–37.
Lessells, J.M. and Murray, W.M. (1941) ‘Effect of shot blasting on strength of metals – Part III’,
Heat Treating and Forging, November, pp.557–558, 567–568.
Li, Y., Quick, N.R. and Kar, A. (2003) ‘Effect of temperature distribution on plasticity in laser
dieless drawing’, J. of Materials Science, Vol. 38, pp.1953–1960.
McAllen, P.J. and Phelan, P. (2004) ‘Comparison of ductile failure models using a simple
elastic-plastic based degradation model’, Journal of Materials Processing Technology,
Vol. 155–156, pp.1214–1219.
Muhr, T.H. (1993) ‘New technologies for engine valve springs’, SAE Technical Paper Series,
No. 930912.
Mukhopadhyay, N.K., Das, S.K., Ravikumar, B., Ranganath, V.R. and Ghosh Chowdhury, S.
(1997) ‘Premature failure of a leaf spring due to improper materials processing’, Engineering
Failure Analysis, Vol. 4, No. 3, pp.161–170.
Murai, S.T. and Takenoshita, N. (1989) ‘High sag resistance spring wire for automotive
suspension’, SAE Technical Paper, No. 850059.
Murakami, Y. and Beretta, S. (1999) ‘Small defects and inhomogeneities in fatigue strength:
experiments, models and statistical implications’, Extremes, Vol. 2, No. 2, pp.123–147.
Naisai, H., Xuanzhou, L., Jun, Y. and Kui, J. (1987) ‘Improve fatigue limit on coil spring’, Shot
Peening: Science, Technology; Application, Garmisch-Partenkirchen, FRG, pp.541–546.
Nakhimov, D.M., Bron, D.I. and Fertikov, D.G. (1963) ‘Residual stresses in the leaves of
automobile springs’, Metal Science and Heat Treatment, Vol. 5, No. 11, pp.670–672.
Nam, W.J., Lee, C.S. and Ban, D.Y. (2000) ‘Effects of alloy additions and tempering temperature
on the sag resistance of S-Cr steel springs’, Mat. Sci. and Eng., Vol. A289, pp.8–17.
Nomura, M., Morimoto, H. and Toyama, T. (2000) ‘Calculation of ferrite decarburizing depth,
considering chemical composition of steel and heating condition’, ISIJ International, Vol. 40,
No. 6, pp.619–623.
Parusov, V.V., Savýuk, A.N., Sychkov, A.B., Nesterenko, A.M., Oleinik, A.A., Zhigarev, M.A.
and Perchatkin, A.V. (2004) ‘Study of the possibilities for maximizing scale removal from
wire rod before drawing’, Matallurgist, Vol. 48, Nos. 5–6, pp.292–297.
Paulitsch, H., Babilon, A., Kieter, G., Koerfer, K., Vlad, C. and Koch, U. (1979) Method of Making
a Steel Wire Adapted for Cold Drawing, U.S. Patent 4,180,418, 25 December.
Perlick, R.A. (2003) Stainless Steel and Nickel Alloy Wire for Demanding Spring Applications,
Springs, Vol. 42, No. 2, p.36.
Petch, N.J. (1953) ‘The cleavage strength of polycrystals’, J. Iron and Steel Institute, May,
pp.25–28.
Phelippeau, A., Pommier, S., Tsakalakos, T., Clavel, M. and Prioul, C. (2006) ‘Cold drawn steel
wires-processing, residual stresses and ductility – Part I Metallography and finite element
analysis’, Fatigue Fract. Engrg. Mater. Stuct., Vol. 29, pp.243–253.
Prawoto, Y., Sato, N., Otani, I. and Ikeda, M. (2004) ‘Carbon restoration for decarburized layer in
spring steel’, Journal of Materials Engineering and Performance, Vol. 13, No. 5, pp.627–636.
Rakhshtadt, A.G. (1975) ‘The problem of improving the properties of spring steels and alloys’,
Metal Science and Heat Treatment, Vol. 17, No. 8, pp.679–686.
308 L.C.F. Canale, R.N. Penha, G.E. Totten, A.C. Canale and M.R. Gasparini

Reguly, A., Strohaecker, T.R., Krauss, G. and Matlock, D.K. (2004) ‘Quench embrittlement of
hardened 5160 steel as a function of austenitizing temperature’, Met. And Mat. Trans.,
Vol. 35A, pp.153–162.
Roberts, B. and Leahy, T. (1998) ‘Process improvements on rod mills’, Proceedings of the Annual
Convention of Wire Association International, Wire Assoc. Int. Inc., Guilford, CT,
pp.179–189.
SAE Spring Committee (1996) Spring Design Manual Second Edition AE-21, SAE International,
Warrendale, PA, USA.
Sakakibara, T., Kusakari, W., Nakano, O., Yasuda, S., Sugimoto, A. and Watanabe, M. (1993)
‘Development of a new light-weight suspension coil spring’, SAE Technical Paper Series,
No. 930263.
Sato, S., Inoue, K. and Ohno, A. (1981) ‘The effect of shot peening to decarburized spring steel
plate’, Conference Proceedings for ICSP-1, Document Number 1981071, pp.303–312,
http://www.shotpeener.com/library/pdf/1981071.pdf.
Shavrin, O.I. and Red’kin, L.M. (1974) ‘Thermomechanical hardening of laminated springs’, Metal
Science and Heat Treatment, Vol. 16, No. 7, pp.573–575.
Shinohara, T. and Yoshida, K. (2004) ‘Deformation analysis of surface defects in wire drawing’,
AIP Conference Proceedings, 10 June, Vol. 712, No. 1, pp.594–599.
Shinohara, T. and Yoshida, K. (2006) ‘Deformation analysis of surface defects in wire drawing’, in
S. Ghosh, J.C. Castro and J.K. Lee (Eds.) CP 712, Materials Processing and Design:
Modeling Simulation and Applications, NUMIFORM 2004, American Institute of Physics,
New York, NY, pp.594–599.
Taleff, E.M., Lewandowski, J.J. and Pourladian, B. (2002) ‘Microstructure-property relationships
in pearlitic eutectoid and hypereutectoid carbon steels’, Journal of Materials (JOM), July,
pp.25–30.
Tekeli, S. (2002) ‘Enhancement of fatigue strength of SAE 9245 steel by shot peening’, Materials
Letters, Vol. 57, pp.604–608.
Teretév, V.F. (2004) ‘On the problem of the fatigue limit of metallic materials’, Metal Science and
Heat Treatment, Vol. 46, Nos. 5–6, pp.244–249.
Timmins, P. (1979) ‘Spring fracture’, Engineering, April, pp.468–473.
Todinov, M.T. (2000) ‘Residual stresses at the surface of automotive suspension springs’,
J. of Mat. Sci., Vol. 35, pp.3313–3320.
Vanneste, G. (1988) Method for the Treatment of Steel Wires, U.S. Patent 4,767,472, 30 August.
Vorob’eva, I.G. and Rakhstadt, A.G. (1996) ‘Dynamic aging of alloyed spring steels of the pearlitic
class’, Metal Science and Heat Treatment, Vol. 38, Nos. 5–6, pp.207–213.
Xu, J-C., Zhang, D-q. and Shen, B-j. (1982) ‘The fatigue strength and fracture morphology of leaf
spring steel after pre-stressed shot-peening’, 1st International Conference on Shot-Peening,
Oxford, UK: Pergamon Press, pp.367–387.
Yan, J. (1998) ‘Study of the Bauschinger effect in various spring steels’, MS Thesis, University of
Toronto, Toronto, Canada.
Yang, Z. and Wang, Z. (1996) ‘Effect of prestrain on cyclic creep behaviour of a high strength
spring steel’, Mat. Sci. and Eng., Vol. A210, pp.83–93.
Zabil’skii, V.V., Ismagilov, M.M., Sarrak, V.I., Suvarova, S.O. and Ubukhov, N.S. (1978) ‘Effect
of stress tempering on the mechanical properties and the endurance of spring wire and
springs’, Strength of Materials, July, Vol. 107, No. 7, pp.835–839.
Zelin, M. (2002) ‘Microstructure evolution in pearlitic steels during wire drawing’, Acta
Materialia, Vol. 50, pp.4431–4447.
Zhang, L. and Thomas, B.G. (2003) ‘State of the art in evaluation and control of steel cleanliness’,
ISIJ International, Vol. 43, No. 3, pp.271–291.
Overview of factors contributing to steel spring performance and failure 309

Zhang, L., Thomas, B.G., Wong, X. and Cai, K. (2002) ‘Evaluation and control of steel
cleanliness-review’, 85th Steelmaking Conference Proceedings, ISS-AIME, Warrendale, PA,
pp.431–452.
Zolotorevsky, N.Y., Titovets, Y.F. and Vasiliev, D.M. (2004) ‘Microstresses developing and severe
cold drawing in nanoscale pearlitic steel’, Rev. Adv. Mater. Sci., Vol. 4, pp.91–96.
Zoughi, R., Huber, C., Qaddoumi, N., Ranu, E., Otashevich, V., Mishahi, R., Ganchev, S. and
Johnson, T. (1997) ‘Real-time on-line near – field microscope inspection of surface defects on
rolled steel’, 1997 Asia Microwave Conference, Paper No. 5PO1–2, pp.1081–1084.

Notes
1 Definition is published by BIRLA COPPER at http://www.birlacopper.com/faqs.htm,
15 June 2006.
2 Information obtained from Century Spring Corporation, 222 East 16th Street, Los Angeles,
CA 90015; Internet link: www.centuryspring.com.

Das könnte Ihnen auch gefallen