Sie sind auf Seite 1von 24

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/272729745

Material properties of high-strength beryllium-free copper alloys

Article  in  International Journal of Materials and Product Technology · January 2015


DOI: 10.1504/IJMPT.2015.067820

CITATIONS READS

14 2,560

6 authors, including:

Igor Altenberger H.‐A. Kuhn


Wieland-Werke AG Wieland-Werke AG
136 PUBLICATIONS   1,674 CITATIONS    53 PUBLICATIONS   355 CITATIONS   

SEE PROFILE SEE PROFILE

Hilmar R. Müller Mansour Mhaede


Wieland-Werke AG 74 PUBLICATIONS   464 CITATIONS   
51 PUBLICATIONS   127 CITATIONS   
SEE PROFILE
SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Development of innovative treatments for ductile cast Ion View project

Ultrafine-grained high strength Cu-Ni-Si alloys View project

All content following this page was uploaded by Mozhgan Gholami on 24 February 2015.

The user has requested enhancement of the downloaded file.


124 Int. J. Materials and Product Technology, Vol. 50, No. 2, 2015

Material properties of high-strength beryllium-free


copper alloys

Igor Altenberger*, Hans-Achim Kuhn and


Hilmar R. Müller
Wieland-Werke AG,
Graf-Arco-Str. 36,
89079 Ulm, Germany
Email: igor.altenberger@wieland.de
Email: achim.kuhn@wieland.de
Email: hilmar.mueller@wieland.de
*Corresponding author

Mansour Mhaede,
Mozghan Gholami-Kermanshahi and
Lothar Wagner
TU Clausthal,
Institute of Applied Materials Science and Engineering,
Agricolastr. 6, 38678 Clausthal-Zellerfeld, Germany
Email: mansour.mhaede@tu-clausthal.de
Email: mozhgan.gholami.kermanshahi@tu-clausthal.de
Email: lothar.wagner@tu-clausthal.de

Abstract: High strength copper alloys can be produced either by generating


very fine grained low alloyed single phased or precipitation hardened copper
alloys or by highly alloyed precipitation hardened copper alloys. The latter
process requires special processing methods such as spray forming in order to
achieve a sufficiently homogeneous microstructure. Systematic investigations
on the aging behaviour of the highly alloyed nickel-manganese bronze
CuNi20Mn20 demonstrate that fully crystalline copper alloys with hardness
exceeding 500 HV can be produced. In addition to age hardening, swaging or
severe plastic surface deformation can be used for additional grain refinement
and strain hardening before precipitation hardening. In contrast to
CuMn20Ni20, the low-alloyed precipitation hardened copper alloy
CuNi3Si1Mg exhibits excellent thermal and electrical conductivity while
maintaining acceptable strength after swaging and precipitation hardening.
Finally, a systematic comparison between spray-formed or precipitation high
strength hardened copper alloys and classical well-known materials such as
steels or aluminium alloys was carried out by using material property charts
(Ashby-maps) and highlighting the fields of application and unique property
combinations of copper alloys.

Keywords: high-strength copper alloys; materials selection; spray forming;


precipitation hardening; fatigue; mechanical surface treatment; ultra fine
grained materials.

Copyright © 2015 Inderscience Enterprises Ltd.


Material properties of high-strength beryllium-free copper alloys 125

Reference to this paper should be made as follows: Altenberger, I.,


Kuhn, H-A., Müller, H.R., Mhaede, M., Gholami-Kermanshahi, M. and
Wagner, L. (2015) ‘Material properties of high-strength beryllium-free copper
alloys’, Int. J. Materials and Product Technology, Vol. 50, No. 2, pp.124–146.

Biographical notes: Igor Altenberger studied Materials Science at the


University of Erlangen – Nürnberg. He obtained his PhD degree with summa
cum laude in Materials Engineering from University of Kassel (Germany) in
1999. In 2000, he joined Prof. R.O. Ritchie’s Fatigue and Fracture Group at the
Lawrence Berkeley National Laboratory, Berkeley (USA), where he worked as
a Post-Doctoral Research Associate on fatigue-related issues such as
foreign-object damage in titanium alloys. After his return to Germany, he
headed his own DFG-funded Emmy Noether-Group of Excellence and
investigated the high temperature fatigue behaviour of various laser-shock
peened materials. In 2007, he became Professor and Head of Department in
Materials Science at the German University in Cairo (Egypt). Today, he works
in Research and Product Development at Wieland Werke AG (Germany) and
teaches materials science at the University of Applied Sciences in Ulm
(Germany). He has authored or co-authored more than 100 publications.

Hans-Achim Kuhn has studied Materials Science at the University of


Erlangen-Nürnberg where he also obtained his PhD on the Microstructure of a
Nickel-Base Superalloy. After several years as a senior scientist in the group of
Prof. H. Mughrabi, he obtained a position as Developing Engineer and Project
Leader in the central laboratory of Wieland Werke AG in the Metals Division.
From 1992 to 1997, he worked in the foundry of Wieland Werke AG in
Vöhringen, where he devoted his professional activities to the improvement of
continuous casting processes. Since 1998, he is leading the Product Technology
Department of the Metals Division. He holds numerous patents on
semi-finished copper alloys and products. Between 2011 and 2013, he was
elected Chairman of the GDMB Copper Committee. In addition, he is also a
Lecturer for Materials Science at the University of Applied Sciences in Ulm,
Germany.

Hilmar R. Müller studied both Mechanical and Process Engineering at the TU


Clausthal, from 1973 to 1978, following which he worked as Scientific
Assistant and Senior Engineer at the Institute of Heat Engineering and
Industrial Furnace Design. After his graduation to Dr.- Ing., he joined
Wieland-Werke AG, Ulm, in 1982, where, between 1983 and 2001, he headed
the Process Development Department in the Central Lab and Development
Services, essentially involved in continuous casting. By the end of 2001, the
spray forming project had become a production department, managed by him.
Following many years of cooperation in various groups of the expert committee
on continuous casting of Deutsche Gesellschaft für Materialkunde (DGM), the
author was elected its Chairman, at the end of 2002. Since 2014, he is speaker
of the DGM-committee chairmen and member of the DGM-board. In 2009, he
became Chairman of the expert committee on spray forming of AWT.

Mansour Mhaede received a Bachelor’s degree (1996) and a Master’s degree


(2001) in Chemical Engineering from Minia University, Egypt. In 2008, he
obtained a Dr.-Ing. degree from TU Clausthal (Germany). From 2004, he has
been working as a Research Assistant at the Institute of Materials Science and
Engineering of TU Clausthal. His current research interests are surface
treatments of copper as well as of light alloys in correlation with fatigue
performance, corrosion fatigue and corrosion behaviour.
126 I. Altenberger et al.

Mozghan Gholami-Kermanshahi has studied Metallurgical and Material


Engineering at Shahid Bahonar Kerman (Bachelor’s degree in 1996) and
Esfahan University of Technology (Master’s degree in 2000) in Iran. From
2001 to 2010, she was working as a Technical Inspector in Iran Standard and
Quality Inspection Company. She is currently a Research Assistant in the group
of Prof. Lothar Wagner. Her research interests are centred around the topics
heat treatment, mechanical properties and corrosion of metallic alloys.

Lothar Wagner studied Mechanical Engineering with specialisation in


Materials Technology. He holds Dipl.-Ing. (1978) and Dr.-Ing (1981) degrees
from the University of Bochum and a Dr.-Ing. habil. (1989) degree from TU
Hamburg-Harburg. In 1993, he was offered the Chair of Materials Technology
at TU Cottbus and in 2002, the Chair of Applied Materials Science and
Engineering at TU Clausthal where he became head of this institute in 2005.
His main research interests are microstructure/property relationships with a
strong focus on fatigue performance and surface treatment.

This paper is a revised and expanded version of a plenary lecture by


I. Altenberger entitled ‘High strength copper alloys’ presented at the CLT
2011, Clausthal-Zellerfeld, Germany, 21–22 September 2011.

1 Introduction

With a density of 8.9 g/cm³ copper does not classify as a light weight material, however,
high specific values of material properties (e.g., specific tensile strength, specific yield
strength or specific resilience) can nevertheless be achieved by extreme strength values.
Furthermore, many copper alloys are excellently work-hardenable (i.e., wires of low
alloyed single phased copper alloys can be drawn to yield strengths greater than 1,000
MPa) and often solid solution hardenable within a wide range of alloying element
content. Consequently, a multitude of possible combined hardening mechanisms render
copper alloys into perfect high strength materials if mechanical metallurgy principles are
applied.
If specific strength is combined with other physical properties, copper alloys are
especially attractive. Typical examples are low permeability (for off-shore and on-shore
oil drilling tools) or excellent electrical and thermal conductivity (e.g., for heat
exchangers or electromechanical connectors), low galling tendency (for cold and hot
working tools in the steel industry) or sea-water corrosion resistance.
As of 2013, high strength copper alloys are used in numerous industrial applications
such as springs, drilling tools, lead frames, electromechanical connectors or prematerials
for low-temperature super-conductors.
This paper mainly concentrates on three classes of alloys: Precipitation-hardened
Corson-type- or Hyper-Corson-alloys (Kuhn et al., 2012), fine grained or ultrafine-
grained (UFG) alloys (single-phased or precipitation hardened and finally spray-formed
copper alloys. Alternative copper-based alloys (composites as well as bulk metallic
glasses) are outlined briefly.
Material properties of high-strength beryllium-free copper alloys 127

2 Experimental procedures

Backscatter/channelling micrographs of the investigated microstructures were taken by


using an angle selective backscatter detector in a ZEISS ultra scanning electron
microscope (SEM), equipped with a thermal Schottky field emission cathode. For
Electron channelling contrast imaging (ECCI) an angle selective detector was used.
Typically, acceleration voltages of 15–20 kV and an aperture of 120 microns were used.
The working distance while recording the SEM pictures was typically 2–6 mm
[Altenberger et al., (2012a), p.79]. These adjustments were also used for the electron
backscatter diffraction (EBSD) investigations, where an EBSD-unit by Oxford, using a
4 × 4 binning, was applied. For the data registration and indexing of the Kikuchi-patterns
the commercial software CHANNEL 5 was used. Prior to characterisation by
SEM/EBSD the specimens were carefully mechanically ground up to 2,400, then
polished up to 1 micron and finally vibration-polished using a magnesium oxide
suspension in order to minimise preparation-induced near-surface cold-work.
The modulus of resilience is defined by the stored energy of a spring as: resilience =
(σYield)²/2E. The Young’s modulus E is taken from the classical tensile test. For more
precise measurements of the modulus of resilience the Young’s modulus is preferably
measured by the dynamical resonance method (Förster, 1937).
Hardness was measured by instrumented indentation using a hardness tester by
Fischer company. Tensile tests were carried out on a standard tensile testing machine by
Zwick company according to standard DIN EN ISO 6892-1 to obtain engineering stress-
strain data.
Fatigue tests were performed on hour-glass shaped specimens with 6 mm minimum
gage diameter in rotating beam loading (R = –1) in air at a frequency of 50 s–1. A
mechanically polished (MP) condition was used as a reference to which the various
mechanically surface treated conditions were compared.
Laser shock peening (LSP) was performed on the as-machined specimens at Toshiba,
Yokohama, Japan. The process Parameters were: 50 mJ pulse energy, 0.4 mm focal spot
diameter and 119 pulses/mm2 irradiation density were applied.
Shot peening (SP) was done using a direct pressure blast system of OSK at IWW of
TU Clausthal. Spherically conditioned cut wire (SCCW14) having an average diameter of
0.35 mm was used. All peening was performed to full (100%) coverage at an Almen
intensity of 0.20 mmA. Deep rolling (ball-burnishing) was performed at IWW of TU
Clausthal by using a conventional lathe and a hydrostatically driven device from Ecoroll
AG, Celle, Germany. A hard metal ball of ∅ 6 mm (HG 6) was utilised at a working
pressure of 300 bar.

3 Mechanical properties of high strength copper-alloys

In former times, high-strength copper alloys were dominated by precipitation hardened


copper alloys containing toxic beryllium due to outstanding strength at acceptable
conductivity. Within the last two decades, however, health and safety issues, legislation
128 I. Altenberger et al.

urges and a series of health-related scandals in the beryllium manufacturing industry have
clearly demonstrated the need for Be-free high strength copper alloys (Schuler et al.,
2005). Today, a variety of Be-free medium conductive copper alloys exists. A selection
of these is exhibited in Figure 1. In respect to mechanical properties Cu-alloys are
equally attractive to steel. They can be produced into conditions with a wide range of
mechanical properties, from highly ductile to ultra-high strength conditions (Figures 2
and 3). Disadvantageous, however, is the high density (8.94 g/cm³ for pure, oxygen-free
copper) rendering light-weight applications almost impossible. Table 1 illustrates the
specific strength of different metallic materials in the high strength condition, derived
from tensile strength and density.
One can see that even ultra high strength copper alloys are still inferior to titanium
alloys (in both tension or bending/torsion) in terms of specific strength in spite of higher
tensile strength. Unfortunately, high strength copper alloys exhibit either similar or only
slightly smaller densities than steels. In many cases (as for CuMn20Ni20) the density of
copper alloys is even higher than those of steels. Comparably low densities in copper
alloys can be achieved in Al-bronzes such as spray-formed CuAl15Fe5Co2Mn2 (density
of 7.0 g/cm²). Obviously, a reduction of density in copper alloys by varying alloying
elements is very limited and not very likely to contribute significantly to a substantial
increase of specific strength. A more promising approach is therefore the optimisation of
tensile and yield strength of copper alloys by applying different microstructural
hardening mechanisms of the bulk or surface regions of the material. These strengthening
mechanisms will be discussed in the following chapters while regarding typical Cu-based
materials in an exemplary manner.

Figure 1 Tensile strength and elongation of some be-free high-strength copper alloys with
electrical conductivities from 10% to 70% IACS (see online version for colours)

18
CuCrAgFeTiSi
16
CuNi3Si1Mg (UFG)
14

12
Elongation (%)

10 CuNi7Si2Cr
spray formed
CuNi3Si1Mg
8
CuSn6 CuNi7Si2Cr spray formed (UFG)
6

CuFe2P CuCo1Ni1Si
4 CuSn8 supralloy®
CuNi3Si1Mg
2 thin wire
CuNi4Co1Si1Mg

0
400 500 600 700 800 900 1000 1100
Tensile Strength (MPa)
Material properties of high-strength beryllium-free copper alloys 129

Figure 2 Tensile strength and elongation of copper alloys as compared to steels and titanium

Source: Altenberger et al. (2009)


Table 1 Specific strengths (for tension and bending/torsion) of different metallic materials
including high strength spray-formed and drawn wire CuMn20Ni20 and Cu-based
metallic glass

Tensile strength 2/3/


Tensile strength Density Tensile strength/Density Density
Material
(MPa) (g/cm³) (MPa cm³/g) (tension) (MPa cm³/g)
(bending or torsion)
Aluminium alloy 600 2,7 222 26
Titanium alloy 1,450 4,5 322 28
Steel 2,500 7,8 320 23
Copper alloy 1,800 8,5 211 17
Magnesium alloy 350 1,8 194 27
Cu47Ti34Zr11Ni8 2,100 7,5 280 22
Notes: In this table the strength values of the following material classes were inserted:
precipitation strengthened 7xxx – aluminium alloy, high strength Beta-titanium
alloy, high alloyed quenched and tempered steel, spray-formed and precipitation
hardened CuMn20Ni20 wire, magnesium wrought alloy ZK60 and Cu-based
metallic glass Cu47Ti34Zr11Ni8.

4 Spray-formed copper alloys

Spray forming is a technology for producing high alloyed homogeneous semi-finished


metallic products. One can see from Figure 1 that spray-forming is an excellent process
technology for producing high-strength copper alloys. The strength and ductility of spray-
formed alloys are often superior to those processed by conventional technologies such as
continuous or semi-continuous casting. This is due to the very homogeneous, fine-
grained and almost segregation-free microstructure prior to hot working (usually hot
extrusion or hot rolling).
130 I. Altenberger et al.

Figure 3 Yield strength of copper-based alloys at elevated temperature (see online version for
colours)
1000
CuNi7Si2Cr
900
Amz irc

800 X10CrNiNb18 9
CuNi20Mn20
700 CuCrZr
Rp0.2 (MPa)

CuNi30Mn
600
Cu-DHP
500 AA 2218 T61

400

300

200

100

0
0 200 400 600 800 1000

Temperature (°C)

Note: CuNi7Si2Cr and CuNi20Mn20 are typically produced by spray forming.


At Wieland-Werke AG, spray forming was performed by using the Osprey process for
billet production. In this facility, a single free-fall atomiser is used with a typical
deposition rate of 30 kg/min. Rotation-symmetrical billets with a typical diameter
between 160 mm and 500 mm and a length of 2.2 m are produced. For detailed archival
literature about this facility see Müller and Zauter (2003).
Spray forming has been widely recognised as an elegant manufacturing method for
generating microstructures or material states which by other conventional manufacturing
methods, such as continuous casting, are not easily or not at all producible due to
problems in segregations or casting ability, especially in alloy systems with broad
solidifying intervals (Abächerli, 2005).
Today, spray forming is able to produce homogeneous billets of large geometry with
fairly low porosity, especially after surface region removal by machining. Hot and cold
porosity, which are inevitable in spray-formed materials, can be controlled by the
gas/metal flow rate and by reactive elements, e.g., titanium (Müller et al., 2003, 2004).
By careful selection of process parameters porosity can be kept to a minimum. Porosity
levels of 1% or lower can be achieved in the centre of the billet for mass production. In
addition, a subsequent non-destructive inspection by an ultrasonic probe detects
remaining critical porosity and thus maintains quality assurance. Smaller porosities are
mostly eliminated during the subsequent hot-extrusion process.
Since spray-formed materials are usually highly solid solution- or precipitation
hardened, they are also a very attractive choice for copper-based high temperature
materials. Figure 4 shows the high temperature strength of spray-formed copper materials
as compared to aluminium and austenitic steel.
Material properties of high-strength beryllium-free copper alloys 131

Figure 4 Fatigue strength/yield strength relation for unnotched Cu- and Al-alloys (see online
version for colours)
800

10 Fatigue strength (MPa) 700

600 Cu alloys

500

400

300

200
7

Al alloys
100

0
0 200 400 600 800 1000 1200

Yield Strength (MPa)

Source: Raw data taken from Ashby (2005)


Up to a temperature of 650°C, spray-formed copper alloys (e.g., CuNi20Mn20 or
CuNi7Si2Cr) exhibit superior yield strengths as compared to austenitic steels or the high-
temperature copper alloy AMZIRC (CuZr0.15) (Figure 3).

Figure 5 Fatigue strength and tensile strength of various be-free copper alloys
500

450
CuSn8
400
Fatigue strength (MPa)

350
CuNi1Co1Si
300
CuCr0.5Ag0.1
250 CuZn15
200 CuZn20 CuZn30 CuNi3Si1Mg
150

100 CuFe2P CuZn37

50

0
0 200 400 600 800 1000
Tensile strength (MPa)

Source: Kuhn et al. (2012)

5 Fatigue behaviour and effect of surface treatments

One of the most important properties of many copper-based components (e.g., springs or
electromechanical connectors) from the point of applicability, is the ability to withstand
132 I. Altenberger et al.

repeated cyclic loads for many million cycles under bending-, push-pull or torsional
loading conditions. The fatigue strengths of high-performance copper alloys exceed those
of pure copper up to several hundred MPa and are superior to those of non-hardened
plain carbon steel, austenitic steels or at least equal to quenched and tempered steels.
According to difference in slopes in Figure 4, copper alloys reveal higher fatigue strength
increases than aluminium alloys for equal increase of yield strength. Figure 5 illustrates
the fatigue strength behaviour (endurance strength for 107 cycles) of several copper
alloys.

Figure 6 Effect of different mechanical surface treatments [SP, roller burnishing (deep rolling)
and laser peening] on the stress-controlled s/n-behaviour of the spray-formed high
strength copper alloy CuNi20Mn20

Similar to other metallic materials and alloys, there is (for smooth, unnotched specimens)
an empirical relationship between the fatigue strength and the tensile strength (this is
sometimes practical, since strain-life data, as required by the Coffin-Manson law is not
always readily available). Within the same chemical composition the fatigue strength
rises with increasing tensile strength. The fatigue strength/tensile strength ratio strongly
differs from alloy to alloy and is in the range of 0.25–0.55. The fatigue strengths of
electromechanical connectors alloys such as CuSn8 and CuNi1Co1Si under reversed
bending are in the range 300–450 MPa (Kuhn et al., 2012).
After certain fully aged and cold worked conditions, beryllium copper alloys may
reach up to 700 MPa fatigue strength (in alternating bending).
In most cases fatigue damage originates from the surface or near surface regions of
highly loaded components due to stress gradients, roughness (extrinsic or intrinsic by slip
line formation) or corrosion. Many methods for fatigue strength optimisation therefore
concentrate on optimising the microstructure, hardness and stress state of the surface
areas. Mechanical surface treatments such as SP, ball burnishing (deep rolling) or laser
(shock) peening (Altenberger, 2003, 2005; Schulze, 2005) can significantly improve the
fatigue strength and -life of high strength copper alloys through induced near-surface
work hardening as well as near-surface compressive residual stresses, as demonstrated in
Material properties of high-strength beryllium-free copper alloys 133

Figure 6–9 for the precipitation hardened spray-formed high strength alloys
CuNi20Mn20 and CuNi7Si2Cr.
These treatments are especially effective in the high cycle fatigue (HCF) regime,
where cyclically stable near-surface work hardening and compressive residual stresses act
to suppress or retard crack initiation and diminish fatigue crack growth rates (Altenberger
et al., 2012b). Some surface treatments such as roller burnishing (deep rolling) also
effectively reduce the surface roughness originating from prior machining. If
precipitation hardening and mechanical surface treatment are combined with severe
plastic deformation methods to obtain an ultra fine grain structures, fatigue strengths in
Be-free copper alloys of 600 MPa can be reached, as demonstrated for the spray-formed
alloy CuNi20Mn20 (Figure 7).

Figure 7 Effect of swaging (sw), subsequent aging and ball burnishing on the s/n-behaviour of
spray-formed CuNi20Mn20

Figure 8 Effect of SP on the s/n-behaviour of spray-formed precipitation hardened CuNi7Si2Cr


134 I. Altenberger et al.

Figure 9 Surface hardness increase by laser (shock) peening in the spray-formed high strength
copper alloys CuNi20Mn20, CuNi7Si2Cr (peening intensity: 60 mj) and the aluminium
bronze CuAl13Fe5Mn1Co1 (peening intensity: 200 mj) (see online version for colours)

Figure 10 Electrical conductivity and yield strength of various copper alloy groups
(see online version for colours)

6 Classical precipitation hardening

Precipitation hardening is applied frequently to copper alloys if the primary aim is to


combine high strength with high (thermal or electrical) conductivity. By concentrating
the alloying elements in fine precipitates the Cu-matrix remains relatively pure with only
few interstitial or substitutional atoms left in the Cu-matrix. Consequently conductivity is
Material properties of high-strength beryllium-free copper alloys 135

not detrimentally affected by solid solution impurities while maintaining high yield
strength by finely dispersed precipitates acting as very effective dislocation obstacles.
The working horse of copper alloys combining high strength with high conductivity
are essentially Corson-alloys (Corson, 1927) which are mostly based on the ternary
system Cu-Ni-Si. In addition, Cu-Ag, Cu-Be or Cu-Nb alloys are known to provide
excellent combinations of strength and conductivity, however as compared to Cu-Ni-Si
alloys they are commercially/industrially less relevant due to higher metal prices or
toxicity problems. Figure 10 gives an overview of the electrical conductivity and yield
strength of the most relevant copper alloy classes.

Figure 11 Ni-silicide-precipitates in fully aged CuNi3Si1Mg (SEM-micrograph using backscatter


electrons)

Figure 12 Ashby-map of the electrical resistivity versus fracture toughness for steels, copper
alloys and aluminium alloys (see online version for colours)
40
Electrical Resistivity (µOhm cm)

30

Steels
20

10 Aluminium alloys Copper alloys

Fracture Toughness (MPa m1/2)


0
50 150 200

Source: Data from Ashby (2005)


136 I. Altenberger et al.

The Cu-Ni-Si-system has been thoroughly already studied by (Corson, 1927).


Microstructurally, the high yield strength of up to 800–900 MPa (after precipitation
hardening and cold working such as wire drawing) is caused by finely dispersed semi-
coherent Ni-Si-precipitates with a diameter of < 20 nm (Figure 11).
Numerous consumers and manufacturers promote Be-free copper alloys, hence
product developments of silicide hardened copper alloys aim at replacing copper-
beryllium alloys completely. Targets are combinations of yield strength and electrical
conductivity of 800 MPa and 50% IACS or 950 MPa and 30% IACS. This is significantly
superior to the first generation of Corson alloys used for connector materials such as
CuNi2Si, CuNi3Si1Mg and CuNi3Si, where volume fractions of precipitates of 2%–3%
are associated to 700 MPa and 50% IACS. In addition to these basic physical and
mechanical properties new materials have to meet requirements for excellent thermal
stability with 70% remaining stress after 1,000 h at 150°C as well as satisfactory
formability.
Variations of the chemical compositions of classical Corson-alloys are able to
increase the volume fraction of particles and/or improve thermal stability of precipitates.
In this context an understanding of the role of alloying elements is crucial.
Similarly as cobalt, nickel reduces the solubility of silicon in copper. From the quasi
binary phase diagrams, one can deduce an increase of solvus of silicides as a function of
the sum of the silicide forming elements Ni, Co and Si. The temperature of solvus line
increases also with the ratio of Co versus Ni.
For example the solvus temperature of pure Ni2Si is 770°C for 2.5 wt% of Ni plus Si
and 800°C for 3 wt%. For comparison the solution temperature of pure Co2Si is
approximately 1070°C for 2.5 wt% Co plus Si. With a Ni:Co ratio of 1 at composition
sum of Ni, Co and Si of 2.5 wt% the solution annealing temperature is reduced to 960°C
(Mandigo et al., 2004). The superior strengthening effect of Co-silicides as compared to
Ni-silicides was also demonstrated by Fujiwara (2004). The hardness increase depends on
the size, crystallographic structure and the distance between the involved precipitates.
The radius of effective silicides is less than 100 nm.
Successful solution annealing requires subsequent quenching of annealed strips. For
solution annealing strips, typical solution annealing temperatures are above 850°C. A
minimum of electrical conductivity is an indicator for optimised solution annealing. For
further processing of strips, grain sizes of 20 µm and less are desirable.
Eventually, the quenched strips are age hardened at 400 to 500°C. With respect to
isotropy of bending the amount of final cold roll prior to final age hardening is usually
25%–50%. An optimised combination of yield strength and electrical conductivity can be
obtained when the (Ni + Co)/Si ratio is between 3.8 and 5.
Due to the enhanced solution temperature, copper alloys exhibiting mixed silicides
(Ni, Co)2Si reveal improved stress relaxation resistance (evolution of stress with time at a
certain temperature at a given strain) as compared to first generation Corson alloys
containing only Ni2Si.
Non-shearable precipitates lead to further hardening in dependence of their volume
fraction and their radii. Strengthening of Copper-alloys hardened by silicides takes
advantage of the Orowan–mechanism due to the semi-coherency of the orthorhombic
crystal structure oP12 of (Ni, Co)2Si precipitates (Kuhn et al., 2007). The contribution of
particle hardening to yield strength can be improved by an increase of volume fraction f
and a reduction of particle radius r of precipitates. For an increase of 100 MPa in yield
strength a microstructure with finely distributed small precipitates with radii less than
Material properties of high-strength beryllium-free copper alloys 137

10 nm and at a volume fraction of three and more percent is required. Higher volume
fractions of 4 - 5% can be achieved in Hyper-Corson alloys exhibiting 4–5 wt% Ni.
Figure 12 exhibits typical precipitate arrangements in CuNi3Si1Mg strip. The radii of
most particles vary between 2 and 100 nm.
In addition to cobalt, nickel and silicon, other elements such as chromium and
magnesium lead to further strength increase and thermal stability of Corson-type alloys.
In contrast to lower alloyed CuNi1Co1Si chromium forms coarse chromium containing
silicides. One disadvantage of such coarse precipitates is increased wear of stamping
tools in the connector production root. Magnesium atoms improve stress relaxation
resistance by solid solution hardening due to their large difference of atomic radii rMg as
compared to copper rcu. (rMg – rCu / rCu) is of the order of 25%. Moreover, Mg also
promotes the formation of silicides (as known for CuNi3Si1Mg).
In the following, the novel Corson-type alloy CuNi3.9Co0.9Si1.2Mg0.1 (containing
Ni- and Co- mixed silicides) with a yield strength above 900 MPa and a composition of
more than 5 wt % alloying elements is called hyper corson alloy. In contrast to the
microstructure of lower alloyed C70250 and C70350 some silicides of hyper corson
alloys exhibit sizes of 5 µm. However large silicides do not contribute significantly to
strengthening, their main benefit is rather to inhibit grain growth.
The elastic behaviour of electromechanical connectors is crucial since they, besides
conducting electric current, act as springs. For maintaining their grip during service the
connectors have to exert high elastic forces onto their elastic contact partner and time- or
temperature-dependent stress relaxation has to be minimised.
The ability of a spring to store elastic energy is described by the modulus of
resilience. Springs with a high resilience can store and release (upon unloading) a lot of
elastic energy, springs with a low resilience can store and release only little elastic
energy. In a stress-strain-curve the modulus of resilience of any material is defined as the
area under the stress-strain curve (or strain energy density) up to the elastic limit (yield
strength) of the material. The modulus of resilience rises with increasing yield strength
and with decreasing Young’s modulus.

Figure 13 Ashby-map of the fracture toughness versus thermal conductivities for copper alloys
and aluminium alloys (see online version for colours)

Source: Data from Ashby (2005)


138 I. Altenberger et al.

The modulus of resilience is depicted in Table 2 for several copper alloys (including the
aforementioned electromechanical connector alloys CuNi3Si1Mg and CuNi1Co1Si). The
modulus of resilience of suitable copper based alloys is typically in the range of
0.8–3.5 MPa (depending on the strength and temper condition, see also (Kuhn et al.,
2007). If even higher resilience is desired, a perfect copper-based spring would be made
from metallic glass (such as Cu47Ti34Zr11Ni8). However, the electrical and thermal
conductivity of bulk metallic glasses are rather low.
Table 2 Resilience and Young’s modulus of various copper alloys

Modulus of resilience (MPa) Young’s modulus (GPa)


Cu-DHP 0.15–0.7 132
CuZn37 1.2–1.7 110
CuNi3Si1Mg 0.9–2.5 130
CuNi1Co1Si 1.0–2.5 137
CuNi5Si1,5Co1Cr 1.1–3.0 135
CuSn8 1.7–3.5 115

The modulus of resilience is temperature- and orientation (and hence texture-) dependent.
In strips exhibiting a typical rolling texture, a high modulus of resilience occurs
perpendicular and parallel to the rolling direction, whereas in angles in between these
orientations lower values for the modulus of resilience occurs.

Figure 14 EBSD-orientation maps of the microstructure of ECAP-processed (8 passes) oxygen-


free pure copper (see online version for colours)

For safety-relevant applications, it is important to point out that copper or copper alloys
are extremely suitable materials for damage-tolerant design. The most important criteria
for damage-tolerant design is the plain strain fracture toughness, which describes the
resistance against propagation of microstructurally large cracks. Within the metallic
Material properties of high-strength beryllium-free copper alloys 139

material universe, copper alloys exhibit a unique combination of very high fracture
toughness and excellent electrical and thermal conductivity and are obviously highly
superior to aluminium alloys in both properties. This fact is illustrated in Figures 12 and
13. Knowing this, we can deduce that copper alloys are much more suitable materials for
current-bearing safety-relevant parts than all other traditional metallic alloys. This is
already being taken advantage of in high strength copper-based electromechanical
connector or heat exchangers, where electrical or thermal conductivity and excellent
fracture toughness as well as sufficient fatigue strength are required for a fail-safe
operation.

Figure 15 (a) EBSD band-contrast (b) orientation map of the microstructure of rotary swaged
CuNi3Si1Mg showing grain sizes of 200–500 nm in diameter (see online version for
colours)

=2 µm; Map18; Step=0.02 µm; Grid258x176 =2 µm; Copy of NewEuler; Step=0.02 µm; Grid258x176

(a) (b)

Figure 16 Grain boundaries and misorientation angle profile of rotary swaged CuNi3siMg as
obtained by EBSD, revealing high angle grain boundaries (black) and a few low angle
grain boundaries (red) (see online version for colours)

=2 µm; Copy of Copy of Map17; Step=0.02 µm; Grid258x176


140 I. Altenberger et al.

Figure 17 Nanoscale ni2si-precipitates (arrows) situated at grain boundaries of UFG CuNi3Si1Mg


(backscatter sem image, channelling (ECCI)-contrast)

7 UFG copper alloys

The Hall-Petch-relation tells us that the yield strength of crystalline, metallic materials
can be enhanced significantly by reducing the grain size (Hall, 1951). Although, today, it
is well known that for extremely small grain sizes (around 10 nm or smaller) the Hall-
Petch relation is not valid anymore (Masumura et al., 1998), grain sizes around 100–500
nm (thus classifying as UFG materials) appear to be industrially more relevant, since a
number of methods have been developed to generate components or semi-finished
products in dimensions > 5–10 mm consisting of UFG alloys. The most well known
methods for producing UFG materials are equal channel angular pressing (ECAP),
accumulative roll bonding (ARB) and high pressure torsion (HPT). However, as of 2013,
numerous other continuous or discontinuous methods have been developed, using
identical or similar physical principles (Valiev et al., 2000; Langdon, 2011). Some of
these methods have entered industrial practice already decades ago, essentially without
knowing their full potential for generating UFG microstructures. Akin to ECAP, swaging
(or rotary swaging) is a very simple method for producing UFG materials and is also
suitable for producing continuous semi-finished materials such as wires. Figures 14 and
15 show the grain-structure of UFG copper and low alloyed copper, respectively, as
obtained by orientation imaging (OIM) using electron backscatter diffraction (EBSD) in
the SEM. Both methods yield UFG microstructures exhibiting grain sizes of 200–500 nm.
It can be confirmed by misorientation measurements using EBSD that the grain
boundaries in CuNi3Si1Mg are indeed high-angle grain boundaries (misorientation angle
between adjacent grains > 10°) (Figure 16). Moreover, within the ultrafine grains some
low-angle grain boundaries (red colour; Figure 16) can be identified by EBSD. By a
combination of swaging and optimised precipitation hardening, a UFG precipitation
hardened grain structure can be formed with orthorhombic semicoherent nanoscale Ni2Si-
precipitates pinning the grain boundaries (Figure 17) and thus providing enhanced
thermal stability at elevated temperature service (For this condition, annealing at 300°C
for 100 hours showed no significant hardness decrease).
Material properties of high-strength beryllium-free copper alloys 141

Figure 18 Microhardness and plastic work of Cu47Ti34Zr11Ni8 as a function of annealing


temperature (see online version for colours)

The mechanical properties of UFG CuNi3Si1Mg are significantly superior to coarse-


grained conditions, i.e., a striking increase of the elongation to fracture from 7% to 14%
was observed in the UFG-variant of CuNi3SiMg1 as compared to the coarse-grained
condition. In addition, tensile strengths of 900–1,000 MPa are feasible, if subsequent
swaging, precipitation hardening and cold work (by rolling or drawing) are applied. Here,
in contrast to high alloyed copper alloys (e.g., produced by spray forming), the electrical
conductivity remains attractive and typically exhibits values of about 30%–35% IACS.

Figure 19 Decomposition of the amorphous state and nanocrystal formation in Cu47Ti34Zr11Ni8


after annealing above the crystallisation temperature Tx (backscatter SEM image
(ECCI-contrast), (a) 500°C/1 h (b) 750°C/1 h

(a) (b)
142 I. Altenberger et al.

Figure 20 Microstructure of CuNb15 wire (diameter 0.5 mm), SEM backscatter contrast,
white: Nb, grey: Cu (cross section)

8 Copper-based metallic glass

A completely different strengthening mechanism and concept is exploited in metallic


glasses where crystalline structure and hence dislocation movement are absent. A variety
of copper-based metallic glasses have been investigated in the past (Zhang et al., 2001).
We will confine our elaborations on metallic glass on the composition Cu47Ti34Zr11Ni8
since it does not contain any hazardous elements (such as beryllium) or extremely
expensive alloying elements. Copper-based metallic glasses typically exhibits extremely
high strength, but poor electrical and thermal conductivity (about 1% IACS). Therefore,
it is not a suitable material for current-bearing components such as electromechanical
connectors. However, the stored elastic energy and strength is superior to any crystalline
copper-based material making it attractive as structural material for applications as
springs. Tensile strengths of copper based metallic glasses can be as high as 2,000–2,500
MPa. Admittedly, copper-based metallic glasses have to stand competition to Fe-, Ni- or
Zr-based metallic glasses with similar or even higher strengths (Löffler, 2004).
Cu47Ti34Zr11Ni8 has an acceptable glass-forming capability and the synthesis of rods
or platelets of 2 mm diameter or thickness are routinely possible by mould-casting in
oxygen free atmosphere. In the as-cast condition.
Cu47Ti34Zr11Ni8 has a hardness of 640 HV0.1, which is significantly superior to
precipitation peak aged and highly cold worked CuMn20Ni20.
Material properties of high-strength beryllium-free copper alloys 143

Figure 21 Hardness and electrical conductivity of several high strength copper alloys
(see online version for colours)

If Cu47Ti34Zr11Ni8 is heat treated at temperatures above the crystallisation temperature


(444°C), a nanostructured material consisting of an amorphous matrix and embedded
nanocrystalline regions can be generated. Figure 18 shows the mechanical properties
(Vickers hardness HV 0.1 and plastic work from registrated indentation testing) of
Vitreloy 101 (Cu47Ti34Zr11Ni8) as a function of annealing temperature (annealing time:
1 h) between 80°C and 800°C. Up to 350°C the hardness does not alter significantly,
however after heat treatment at T > 400°C hardness increases significantly by 30% and
reaches a maximum at T = 500°C. Conversely, the plastic work reaches a minimum after
heat treatment at 500°C, indicating significant annealing embrittlement. Figure 19 shows
the partial nanocrystallisation of Vitreloy 101 after heat treatment at 750°C. Typically the
nanocrystals are globular to rod-like and exhibit grain sizes of < 500 nm, some crystals
are even smaller than 100 nm. Obviously a decomposition of the amorphous state took
place. With increasing annealing temperature the plastic work increases again while the
volume fraction of the crystalline phase rises. One of the biggest challenges for these
‘nano-insitu-composites’ is the enhancement of ductility and damage tolerance. For non-
copper based glass compositions several ductility-enhancing mechanisms have been
investigated and proposed (Eckert et al., 2007) while research on copper-based systems is
ongoing.
In general, critical issues for the application of metallic glasses as structural materials
are the still limited ductility of many metallic glasses as well as the rather mediocre
fatigue strength. For some compositions, however, very impressive damage tolerance has
been achieved (Demetriou et al., 2011).
144 I. Altenberger et al.

9 Composite copper-based materials

Although classical composite materials are beyond the scope of this paper it should be
pointed out that especially Cu-Nb with sufficiently high niobium content (Figure 20),
Cu-Ag- as well as Cu-Ag-Zr- (Gaganov et al., 2004; Freudenberger et al., 2006, 2010)
and Cu-Al2O3 composites (‘Glidcop’) have commercially found their way into niche
applications where material or processing costs are not the main issue. Such composite
materials offer superb combinations of strength and electrical conductivity or very
attractive high temperature properties, respectively. Moreover, composites consisting of
metallic glass and a pure copper or copper alloy with high conductivity has the potential
for high strength-high conductivity applications (Choi-Yim et al., 1998).

10 Conclusions

• Several concepts for designing high-strength copper alloys are suggested: Besides
the classical strengthening mechanisms such as precipitation hardening and cold
working, some more novel methods such as severe plastic deformation followed by
artificial aging, amorphisation as well as in-situ nanocrystallisation of metallic
glasses appear to be promising methods. Preferential hardening of surface zones
(e.g., by mechanical surface treatments) or special processing of semi-finished billets
such as spray forming may further increase the strength and serve to further optimise
the quality and homogeneity of the desired component.
• The strengths of several copper alloys are equally high as those of steels or titanium
alloys while maintaining satisfactory conductivity (Figure 21).
• ‘Ashby-maps’ are very useful tools to explore the potential of copper alloys and
widen their applications, since strength of copper alloys is only attractive in
combination with other functional properties, such as superb electrical and thermal
conductivity, low permeability and good thermal stability (as opposed to aluminium
alloys). Since the commercial success of copper alloys depends so strongly on their
‘secondary properties’ great heed has to be taken by the marketing and development
engineer if copper alloys are to be recommended as the first choice.
• Increasing miniaturisation of electromechanical components will promote the use of
high strength copper alloys in more and more high-end applications.

References
Abächerli, V. (2005) Improvement of Workability and Superconducting Properties of High Tin-
Content (Nb, Ta, Ti)3 Sn Bronze Route Wires, PhD thesis, University Geneva.
Altenberger, I. (2003) ‘Alternative mechanical surface treatments: microstructures, residual stresses
and fatigue behaviour’, in Wagner, L. (Ed.): Shot Peening, Proc. ICSP 8, p.421.
Altenberger, I. (2005) ‘Shot peening’, in Lu, J. (Ed.): Handbook on Residual Stress, p.137, Society
for Experimental Mechanics, Bethel, USA.
Altenberger, I., Kuhn, H-A. and Hölzl, H. (2012a) ‘Mikrostrukturelle Charakterisierung von
hochfesten Cu-Ni-Si-Legierungen mittels Electron-Channeling-Rückstreukontrast im
Rasterelektronenmikroskop’, Sonderband d. Prakt. Metallographie, Vol. 44, pp.79–84.
Material properties of high-strength beryllium-free copper alloys 145

Altenberger, I., Nalla, R.K., Sano, Y., Wagner, L. and Ritchie, R.O. (2012b) ‘On the effect of deep
rolling and laser-peening on the stress-controlled low- and high-cycle fatigue behaviour of Ti-
6Al-4V at elevated temperatures up to 550°C’, Int. J. Fatigue, Vol. 44, p.292.
Altenberger, I., Müller, H.R., Zauter, R. and Kudashov, D.V. (2009) ‘Microstructure of spray-
formed copper alloys’, Proc. SDMA 2009/ICSF VII, Bremen, Germany, CD.
Ashby, M.F. (2005) Materials Selection in Mechanical Design – 3rd Edition, Elsevier, Amsterdam.
Choi-Yim, H., Busch, R. and Johnson, W.L. (1998) ‘The effect of silicon on the glass forming
ability of the CuTiZrNi bulk metallic glass forming alloy during processing of composites’,
Journal of Applied Physics, Vol. 83, No. 12, p.12.
Corson, M.G. (1927) ‘Copper hardened by a new method’, Zeitschrift für Metallkunde, Vol. 19,
p.370.
Demetriou, M.D., Launey, M.E., Garrett, G., Schramm, J.P., Hofmann, D.C., Johnson, W.L. and
Ritchie, R.O. (2011) ‘A damage-tolerant glass’, Nature Materials, Vol. 10, p.123.
Eckert, J., Das, J., Pauly, S. and Duhamel, C. (2007) ‘Processing routes, microstructures and
mechanical properties of metallic glasses and their composites’, Adv. Eng. Mater., Vol. 9,
No. 6, p.443.
Förster, F. (1937) ‘Ein neues Messverfahren zur Bestimmmung des Elastizitätsmoduls und der
Dämpfung’, Zeitschrift für Metallkunde, Vol. 29, pp.109–113.
Freudenberger, J., Botcharova, E., Gaganov, A., Lyubimova, J. and Schultz, J. (2006) ‘Neuartige
Kupferlegierungen für gepulste Hochfeldmagnete’, Metall, Vol. 60, No. 11, p.714.
Freudenberger, J., Lyubimova, J., Gaganov, A., Klauß, H. and Schultz, L. (2010) ‘Mechanical
behaviour of heavily deformed CuAgZr conductor materials’, J. Phys. Conf. Series, Vol. 240,
No. 1, p.012112.
Fujiwara, H., (2004) ‘Designing high-strength copper alloys based on the crystallographic structure
of precipitates’, Furukawa Review, Vol. 26, pp.39–43.
Gaganov, A., Freudenberger, J., Grünberger, W. and Schultz, L. (2004) ‘Microstructural evolution
and ist effect on the mechanical properties of Cu-Ag microcomposites’, Z. Metallkunde,
Vol. 95, No. 6, p.425.
Hall, E.O. (1951) ‘The deformation and ageing of mild steel: iii discussion of results’ Proc. Phys.
Soc. B, Vol. 64, No. 9, p.747.
Kuhn, H-A., Altenberger, I., Käufler, A., Hölzl, H. and Fünfer, M. (2012) ‘Properties of high
performance alloysfor electromechanical connectors’, in Collini, L. (Ed.): Intech, p.52.
Kuhn, H-A., Käufler, A., Ringhand, D. and Theobald, S. (2007) ‘C7035 – a new high performance
copper based alloy for electro-mechanical connectors’, Materialwissenschaft und
Werkstofftechnik, Vol. 38, pp.624–636.
Langdon, T.G. (2011) ‘Processing by severe plastic deformation: historical developments and
current impact’, Mater. Sci. Forum, Vol. 667–669, p.9.
Löffler, J.F. (2004) ‘Metallische Massivgläser als Strukturwerkstoff im Leichtbau’, in
Kaufmann, H. and Uggowitzer, P.J. (Eds.): Proc. 3. Ranshofener Leichtmetalltage, p.225.
Mandigo, F.N., Robinson, P.W., Tyler, D.E., Bögel, A., Kuhn, H-A., Keppeler, M. and Seeger, J.
(2004) US Patent Application 0079456 A1.
Masumura, R.A., Hazzledine, P.M. and Pande, C.S. (1998) ‘Yield stress of fine grained materials’,
Acta. Mater., Vol. 46, No. 13, p.4527.
Müller, H.R. and Zauter, R. (2003) ‘Spray-formed copper alloys – process and industrial
applications‘, Erzmetall, Vol. 56, pp.643–702.
Müller, H.R., Ohla, K., Zauter, R. and Ebner, M. (2003) ‘Porosity in spray-formed copper-alloy
billets’, Proc. SDMA 2003/ICSF V, Bremen, Germany.
Müller, H.R., Ohla, K., Zauter, R. and Ebner, M. (2004) ‘Effect of reactive elements on porosity in
spray-formed copper-alloy billets’, Mater. Sci. Eng., Vol. 38, No. 1, p.78.
146 I. Altenberger et al.

Schuler, C.R., Kent, M.S., Deubner, D.C., Berakis, M.T., McCawley, M., Henneberger, P.K.,
Rossmann, M.D. and Kreiss, K. (2005) ‘Process related risk in beryllium-sensitization and
disease in a copper-beryllium alloy facility’, American Journal of Industrial Medicine,
Vol. 47, No. 3, p.195.
Schulze, V. (2005) Mechanical Surface Treatments, Wiley-VCH, Weinheim.
Valiev, R.Z., Islamgaliev, R.K. and Alexandrov, I.V. (2000) ‘Bulk nanostructured materials from
severe plastic deformation’, Progress in Materials Science, Vol. 45, pp.103–189.
Zhang, T., Kurosaka, K. and Inoue, A. (2001) ‘Thermal and mechanical properties of Cu-based,
Cu-Zr-Ti-Y bulk glassy alloys’, Mater. Trans. JIM, Vol. 42 No. 10, p.2042.

View publication stats

Das könnte Ihnen auch gefallen