Anda di halaman 1dari 22

Journal of ASTM International, Vol. 6, No.

5
Paper ID JAI102131
Available online at www.astm.org

Patricia Mariane Kavalco1 and Lauralice C. F. Canale1

Evolution of Quench Factor Analysis: A Review


ABSTRACT: One of the steps of the heat treatment process of age hardenable aluminum alloys is the
quenching process in which the alloy is cooled from the solutionizing temperature. The objective is to
quench sufficiently fast to avoid undesirable concentration of alloying elements in the defect and grain
boundary structure, while at the same time not quenching faster than necessary to minimize residual
stresses, which may lead to excessive distortion or cracking. Various studies have been conducted to
predict the relative quench rate sensitivity to yield different properties for age-hardenable alloys. Of these
different predictive methods, the one that showed the more realistic results is quench factor analysis since
it involves a correlation of the cooling curve time-temperature curve of the cooling process throughout the
quenching cycle for the desired cross-section size of interest with a C-curve time-temperature-property
curve for the specific alloy of interest. The quench factor analysis numerical procedure has evolved since
its original introduction. A review of the basic assumptions of the classical quench factor analysis model will
be provided here which will include discussion of the various improvements to the classical model that have
been proposed over the intervening years since its introduction.
KEYWORDS: quench factor analysis, aluminum, modeling, solution treatment, ageing, precipitation

Introduction
When aluminum alloys are solutionized at elevated temperatures 399 C, alloying elements are dissolved
in the aluminum lattice structure of the solid solution and the objective is to maximize the concentration
of hardening elements including: copper, zinc, magnesium, and silicon in the solid solution 1. The
concentration and rate of dissolution of these elements in the solid solution increases with increasing
temperature.
After solutionizing, the aluminum alloy is cooled. If the cooling process is too slow, the alloying
elements will diffuse through the solid solution and concentrate at grain boundaries, large voids, undissolved particles, or other defect locations. The diffusion process is slower for some alloys than others,
thus permitting slower cooling rates during quenching. For optimal properties, it is desirable to retard this
diffusion process and to maintain the alloying elements in solid solution. For quench hardenable wrought
alloys 2xxx, 6xxx, and 7xxx, and casting alloys such as 356, this is accomplished by the quenching
process. The objective is to quench sufficiently fast to avoid the undesirable concentration of the alloying
elements in the defect and grain boundary structure, while at the same time not quenching faster than
necessary to minimize residual stresses, which may lead to excessive distortion or cracking 2,3. After
quenching, the aluminum alloy is then aged. During the aging process, hardening elements precipitate in
localized areas, which significantly increases the strength of the part.
An illustration of this effect of cooling times after solutionizing on the properties after aging is the
effect of cooling on corrosion of AA2024-T4 See Table 1 for composition and Table 2 summary of
terminology related to tempering treatment where the critical temperature range and time s for the
transition of pitting to intergranular corrosion is shown by the so-called TTP time-temperature-property
or C-curve in Fig. 1 4. In this case, the C-curve for 2024-T4 is used to show the change in corrosion
behavior by correlating the critical temperature range where precipitation was fastest. Figure 1 shows that
increased intergranular corrosion for 2024-T4 will be favored if cooling rates are excessively slow during
quenching. Similar behavior can be shown for other aluminum alloys 5. Therefore, it is important that
cooling rates during quenching be sufficiently fast to avoid this undesirable behavior.
Manuscript received October 1, 2008; accepted for publication March 1, 2009; published online April 2009.
1
University of So Paulo, Department of Materials, Aeronautical and Automobile Engineering, School of Engineering, So
Carlos, SP, Brazil.
Copyright 2009 by ASTM International, 100 Barr Harbor Drive, PO Box C700, West Conshohocken, PA 19428-2959.

Copyright by ASTM Int'l (all rights reserved); Sun Mar 7 19:29:31 EST 2010
Downloaded/printed by
George Totten (GE Totten Associates Inc) pursuant to License Agreement. No further reproductions authorized.

2 JOURNAL OF ASTM INTERNATIONAL


TABLE 1Nominal compositions of selected heat treatable aluminum alloys.

2024a
0.5
0.5
3.44.9
0.30.9
1.21.8
0.1
0.25

0.15

Remainder

Element
Si
Fe
Cu
Mn
Mg
Cr
Zn
V
Ti
Bi
Ga
Pb
Zr
Al

2219a
0.2
0.3
5.86.8
0.20.4
0.02

0.1
0.050.15
0.020.10

0.100.25
Remainder

Alloy Designation
Element wt. %
6061a
7010b
7050b
0.40.8
0.12
0.12
0.7
0.15
0.15
0.150.40
1.52.0
2.02.6
0.15
0.1
0.1
0.181.2
2.12.6
1.92.6
0.040.35
0.05
0.04
0.25
5.76.7
5.76.7

0.15
0.06
0.06

0.100.16
0.080.15
Remainder
Remainder
Remainder

7075a
0.4
0.5
1.22.0
0.3
2.12.9
0.180.28
5.16.1

0.2

Remainder

356c
6.57.5
0.6
0.25
0.1
0.250.45

0.1

0.20

Remainder

Aluminum composition data, except where otherwise noted here, available from: http://en.wikipedia.org/wiki/Aluminiumalloy
See Ref. 6
c
AA356 composition obtained from: http://www.premieraluminum.com/capabilitiesaluminumalloys.html
b

Fink and Willey performed an extensive study on the effects of quenching on the strength of 7075-T6
5 using isothermal quenching techniques. This was done by constructing C-curves shown in Fig. 2 which
are plots of times required to precipitate sufficient alloy content to change the strength by a certain amount.
The critical temperature range, which is that temperature range that provided the highest precipitation
rates, was identified 4.
The characteristic C-shapes of the TTP curves such as the curves shown for 7075 in Fig. 2 indicates
that long times are typically required for a given amount of solute precipitation at elevated temperatures in
the range of 370 425 C, because the amount of solute supersaturation is low, and therefore, the thermodynamic driving force for precipitate formation is low. At temperatures in the range of 275 350 C, the
undercooling and thermodynamic driving force are high, and the time required to achieve a particular
amount of solute precipitation is relatively lowtypically only a few seconds in precipitation hardenable
aluminum alloys. In this temperature range, the diffusion coefficients for the solute species are reasonably
high, and precipitates can nucleate and grow in short times. As discussed previously, nucleation and
precipitation usually occur first along grain boundaries because less lattice strain is necessary therein and
diffusion rates are higher than within grains. At still lower temperatures, below 200 C, the time required
for a given amount of precipitation increases to hundreds or thousands of seconds because solute diffusion
coefficients are low. The thermodynamic potential for precipitation is high because of the high degree of
solute supersaturation but the rate of precipitate formation is low because of the inability of atoms to
diffuse and form the precipitating species 8.
TABLE 2Tempering designations for aluminum alloys.
Temper Designation
T4
T6
T7
T73
T76

T8

Tempering Process
Solution heat treated and naturally aged
Solution heat treated and artificially aged
Solution heat treated and stabilized over-aged
Solution heat treated and then artificially overaged in
order to achieve the best stress corrosion resistance
Solution heat treated and then artificially overaged in
order to achieve a good exfoliation corrosion
resistance
Solution heat treated, cold worked, and artificially
aged

See Ref. 7. Specific heat-treatment processes time and temperature for each wrought, forged or and extruded alloy are described in AMS 2770.
For cast alloys, see AMS.

Copyright by ASTM Int'l (all rights reserved); Sun Mar 7 19:29:31 EST 2010
Downloaded/printed by
George Totten (GE Totten Associates Inc) pursuant to License Agreement. No further reproductions authorized.

KAVALCO AND CANALE ON REVIEW OF QUENCH FACTOR ANALYSIS 3

FIG. 1C-curve indicating cooling rate dependent corrosion attack on 2024-T4 sheet.
It is important to note that although critical solute temperature may be the identified from a C-curve
for an alloy, for example 340 C for 7075-T73, the specific C-curve obtained will vary depending on the
property being measured as illustrated in Fig. 3 9.
Various studies were conducted after Fink and Willeys work to determine the relative quench rate
sensitivity to yield different properties for various alloys. Figure 4 illustrates the effect of cooling rate on
tensile strength for different aluminum alloys and tempers 4. The average cooling rate has been
traditionally defined to be the time s-s to cool from 400 C to 290 C which will yield an average
cooling rate in C/s. This is the classical approach upon which most standards for the heat treatment of
aluminum alloys are written. The maximum attainable strength properties are dependent on the average
cooling rate. Generally, faster cooling rates provide greater strengths up to a limit.
The traditional approach for modeling quench sensitivity which refers to the amount age hardening
response is reduced by slow quench cooling to quantitatively correlate quenching cooling rate with
mechanical properties of aluminum alloys presumes a linear, well-behaved cooling process between
400 C and 290 C. However, this is seldom the case and therefore, the classical method of determining the
average cooling rate of aluminum alloys can provide only an approximation of the effect of a quenching
process on properties and, in fact, may fail when the quenching process is nonlinear, such as an interrupted

FIG. 2C-curves illustrating the effect of alloy precipitation on tensile strength for 7075-T6 generated by
Fink and Willey.
Copyright by ASTM Int'l (all rights reserved); Sun Mar 7 19:29:31 EST 2010
Downloaded/printed by
George Totten (GE Totten Associates Inc) pursuant to License Agreement. No further reproductions authorized.

4 JOURNAL OF ASTM INTERNATIONAL

FIG. 3C-curves for proof stress, Vickers hardness (HV20) and electrical conductivity (% IACS
International Annealed Copper Standard) for 7075-T73.
or delayed quenching process. Therefore, it is desirable to utilize an alternative process, quench factor
analysis QFA, which correlates actual cooling pathway by using a cooling curve time-temperature
curve of the actual cooling process throughout the quenching cycle for the quenching process and crosssection size being used with a C-curve time-temperature-property curve for the specific alloy of interest
10,11.
QFA is now used routinely by many researchers in the aluminum thermal processing industry. Staley
et al., one of the pioneers in developing the QFA model and its application, has provided numerous
references on experimental model validation 1113. Bates successfully used QFA to model strength for
AA356 14 and 7075 and 2024 8. Rometsch et al. have evaluated QFA model predictions against
published data and have recommended various model refinements 15. Liu, et al. obtained excellent
results using quench factor analysis to predict hardness of a 7055 aluminum alloy 16. Mudawar and
co-workers have used QFA to spray quench 2024-T6 aluminum extrusions and have experimentally validated their results 1719. Flynn and Robinson obtained excellent results using QFA to model properties
of a 7010 aluminum alloy 20. Kim used QFA to study the properties obtained by spray quenching of

FIG. 4Tensile strength as a function average cooling rate.


Copyright by ASTM Int'l (all rights reserved); Sun Mar 7 19:29:31 EST 2010
Downloaded/printed by
George Totten (GE Totten Associates Inc) pursuant to License Agreement. No further reproductions authorized.

KAVALCO AND CANALE ON REVIEW OF QUENCH FACTOR ANALYSIS 5


TABLE 3Nominal compositions of steel alloys.
Alloy Designation
Element wt. %
AISI 4140a
0.380.43
0.751.00
max 0.035

max 0.04
0.81.1
0.150.35
Remainder

AISI 1045a
0.430.5
0.60.9
max 0.04
max 0.05
0.150.35

Remainder

Element
C
Mn
P
S
Si
Cr
Mo
Fe

SAE 5140a
0.45
0.77
0.015
0.013
0.35
0.72
0.25
Remainder

American Iron and Steel Institute AISE

aluminum alloys 21. In addition to aluminum alloys, QFA has been successfully applied to predict
as-quenched hardness for AISI 1045 and 4140 steel alloys 21 and 5140 steel 22 see Table 3.
Over the approximately 25 years since QFA was introduced, the original model has undergone considerable evolution and improvement. The objective of this paper is review the basic assumptions of the
classical model and to highlight its more significant potential deficiencies. This discussion will be followed
by a summary of the various improvements to the classical QFA model that have been proposed over the
intervening years since its introduction.
Discussion
Calculation of Quench Factors from Precipitation Kinetic Data
Avrami used isothermal kinetics to study transformation during continuous cooling processes 23,24. The
Avrami equation, commonly designated as the JohnsonMehlAvramiKolmogorov JMAK equation, is
typically used to calculate a property evolution as a function of time Ft and takes the following form:
Ft = 1 exp Ktn

where:
t time s,
K temperatur dependent rate constant s1, and
n 14 Avrami exponent.
The Avrami equation 1 may be written in the form of volume fraction developed V f at time t:
Ft = V f = 1 X

where:
X the volume fraction transformed.
The volume fraction V f may be written in the form of a property :
Vf =

t i
f i

where:
t property after time t,
i initial value of the property being measured, and
f final value of the property being measured.
Copyright by ASTM Int'l (all rights reserved); Sun Mar 7 19:29:31 EST 2010
Downloaded/printed by
George Totten (GE Totten Associates Inc) pursuant to License Agreement. No further reproductions authorized.

6 JOURNAL OF ASTM INTERNATIONAL


TABLE 4Coefficients for calculating quench factors at 99.5 % of attainable yield strength.

Alloy
7010-T76
7050-T76
7075-T6
7075-T73
7175-T73
2017-T4
2024-T6
2024-T851
2219-T87
6061-T6
356-T6
357-T6
Al-2.7Cu1.6Li-T8

k 1a
0.005 01
0.005 01
0.005 01
0.005 01
0.005 01
0.005 01
0.005 01
0.005 01
0.005 01
0.005 01
0.006 6
0.006 2
0.005 0

k2
s
5.6 1020
2.2 1019
4.1 1013
1.37 1013
1.8 109
6.8 1021
2.38 1012
1.72 1011
0.28 107
5.1 108
3.0 104
1.1 1010
1.8 108

k3
J/mol
5780
5190
1050
1069
526
978
1310
45
200
412
61
154
1520

k4
K
897
850
780
737
750
822
840
750
900
750
764
750
870

k5
J/mol
1.90 105
1.8 105
1.4 105
1.37 105
1.017 105
2.068 105
1.47 105
3.2 104
2.5 104
9.418 104
1.3 105
1.31 105
1.02 105

Calculated
Range
C
425150
425150
425150
425150
425150
425150
425150
425150
425150
425150
425150
425150
425150

Reference
17
4
4
9
27
39
38
24
25
39
37
37
37

k1 is a unitless value which corresponds to the unprecipitated fraction. For this analysis, it is usually 0.995 and ln 0.995 is 0.005 01.

The properties of aluminum alloys are dependent on the amount of alloy precipitation that occurs
during cooling, irregardless of the cooling pathway. The Avrami equation for isothermal precipitation
kinetics is 4:

= 1 exp
where:

t
k

fraction of precipitation which has occurred in time t during the quench, and
k temperature-independent constant s.

The value of k depends on the degree of supersaturation and the rate of diffusion, and is estimated from
the following equation developed by Evancho and Staley 4:

k3 k24
CT
k5
k=
= k2 exp
2 exp
k1
RTk4 T
RT

where:
CT critical time s required to precipitate a constant fraction the equation of the C-curve,
k1 constant which equals the natural logarithm of the fraction untransformed 1 fraction
defined by the C-curve; if 0.5 % is untransformed, then k1 = ln / max = 0.005 013,
k2 constant related to the reciprocal of the number of nucleation sites s,
k3 constant related to the energy required to form a nucleus J/mol,
k4 constant related to the solvus temperature K,
k5 constant related to the activation energy for diffusion, J/mol,
R ideal gas constant= 8.3143 J K1 mol1, and
T temperature K.
Numerical values of each of the constants k1 , k2 , . . . , k5 define the C-curve for the alloy composition
and temper condition. Values for k1 , k2 , . . . , k5 for selected alloys for prediction of yield strength that have
reported to date are provided in Table 4 25. Additional C-curves for alloys not shown in Table 4 that have
been published, but whose k-values have not, are shown in Ref 26.
The earliest references to deriving k-values reported using multiple linear regression analysis to fit the
k2 k5 values to the CT equation and repeat the iterations until minimum computational error is obtained 4.
Using this method, k values are readily attainable from published C-curves 4,27,28. However, Shuey
et al. 27 and Tiryakiolu 28 have studied this approach and found that although C-curves can indeed be
Copyright by ASTM Int'l (all rights reserved); Sun Mar 7 19:29:31 EST 2010
Downloaded/printed by
George Totten (GE Totten Associates Inc) pursuant to License Agreement. No further reproductions authorized.

KAVALCO AND CANALE ON REVIEW OF QUENCH FACTOR ANALYSIS 7

FIG. 5Determination of quench factor Q by the combination of a quenchant cooling curve and a
C-curve.
fitted, their physical meaning was often questionable. Instead, they recommended, when possible, that the
number of variables k-values being fitted should be reduced by using independently reported data such as
solvus temperature, solute diffusivity, and enthalpy of precipitation.
From these relationships, it is possible to redefine the equation for the amount of solute precipitated
during the quench which is called the quench factor and is typically designated by Q. Cahn showed
that transformations that nucleate heterogeneously, such as aluminum alloys, obey the Rule of Additivity,
which states that reactions are additive and therefore transformation kinetics for non-isothermal conditions,
such as those that would be present during a typical quenching process may be described by 29:
Q=

tf

t0

1
dt
CT

where:
Q
CT
t
t0
tf

measure of the amount transformed quench factor,


critical time s from the C-curve and is calculated from Eq 5,
time from the cooling curve s,
time at the start of the quench s, and
time at the finish of the quench s.

When Q = 1, the fraction transformed equals the fraction represented by the C-curve.
By the Rule of Additivity, the temperature-time curve is discretized into a series of isothermal steps as
described by Evancho and Staley 4. For each step, the volume fraction of the new phase that is formed
during this time step is calculated by isothermal kinetics Avrami equation. This is shown in Fig. 5, where
the quench factor Q is obtained by combining the cooling curve for the quenching process with the C-curve
and the value for Q is obtained by 4,21:
t1 t2
tn1
Q=
+
+ +
=
CT1 CT2
CTn1

n1

Ci
1

Ti

In the original classical approach, properties were predicted from the quench factor by 4,8:
y = maxeK1Q
where:
y
max
e
K1
Q

predicted yield strength MPa,


yield strength after an infinite quench and aging cycle MPa,
base of the natural logarithm,
ln0.995 = 0.00501, and
quench factor.

Copyright by ASTM Int'l (all rights reserved); Sun Mar 7 19:29:31 EST 2010
Downloaded/printed by
George Totten (GE Totten Associates Inc) pursuant to License Agreement. No further reproductions authorized.

8 JOURNAL OF ASTM INTERNATIONAL

FIG. 6Yield strength of aluminum 7075-T73 as a function of quench factor of the material.

The relationship between quench factor and yield strength for 7075-T73 is shown in Fig. 6 8.
In their 1974 paper, Evanco and Staley showed that from the Avrami equation, the attainable strength
is a function of the amount of solute remaining in solution after the quench and is related to k1 as follows
4:

k1 = ln

x min
max min

where:
x 0.995 max,
max maximum level of , and
min minimum level of .
When Evancho and Staley analyzed 7075-T6 and 2024-T4 interrupted quenching data previously
published by Fink and Willey, they obtained a linear correlation for log10 / min vs isothermal holding
time and the slope was equal to 1. On the basis of this observation, they concluded that the Avrami
exponent was equal to 1 n = 1. Therefore, Evancho and Staley set min = 0 in Eq 9 to facilitate the use of
Cahns model to predict the extent of transformation in their early work 4,20. In their 1974 paper, Staley
and Evancho further noted that the simplified equation for k1 could be used for high-strength alloys because
min  max 4. Staley et al. also subsequently stated that this assumption for min was made because the
strength of aluminum alloys held for infinitely long times below the solvus temperature would be zero 12.
Furthermore, this simplification was validated by the excellent fit that the simplified value for k1 provided
for C-curve data developed for 7075-T6 and 2024-T4 4. For many industrial applications this provides
adequate results when the loss in properties is less than 10 % 20. However, Swartzendruber et al.
subsequently recommended that Eq 9 was necessary to minimize predictive errors 30.
Schwartzendruber et al. also determined k-values for C-curves developed based on different
mechanical properties which are shown in Table 5. These data show that in addition to the use of Eq 9 it
is necessary to obtain the correct C-curves for the property of interest if reliable predictions are to be made.
It is not correct to use one C-curve and simply vary the property max and min using the same quench
factor Q to obtain the most reliable prediction with minimum error.
From these equations, a number of observations regarding the quenching process showing the power
of QFA in quench process design and analysis:
1. Low values of Q are associated with high quench rates, minimum precipitation during cooling and
high yield strengths.
Copyright by ASTM Int'l (all rights reserved); Sun Mar 7 19:29:31 EST 2010
Downloaded/printed by
George Totten (GE Totten Associates Inc) pursuant to License Agreement. No further reproductions authorized.

KAVALCO AND CANALE ON REVIEW OF QUENCH FACTOR ANALYSIS 9


TABLE 5C-curve parameters for aluminum alloy 2219-T87a,b.

Property
Hardness, HRB
Yield Strength,
0.2 % set in ksi
Tensile Strength
ksi
Conductivity
% IACSd

k 1c
0.005 01
0.005 01

k2
s
0.86 1010
0.78 1010

k3
J/mol
320
320

k4
K
900
900

k5
J/mol
3.2 104
3.2 104

max
79
55.3

min
46
35.4

0.005 01

0.79 1010

320

900

3.2 104

69

45.1

0.005 01

0.59 1010

320

900

3.2 104

33.2

37.7

T-87 indicates that a 10 % stretch was applied prior to thermomechanical treatment. Usually, for T-87, a 5 % or 7 % stretch is applied.
These k-values were determined from 2219 plate with a thickness of 0.625 cm 1 / 4 in., width 1.22 m 4 ft, length of 1.83 m 6 ft. Before
thermal treatment, a 0.6 m wide sample was removed from each edge and the middle for analysis. The remaining 0.625 cm thick plate was cut into
2.5 cm wide by 17 cm long test bards. The test bards were solutionized at 535 C for 75 min and then transferred directly to a salt bath and heated
from 2 3600 s, at which time the test bar was water quenched, mechanical stretch to 5 % and aged for 16 h at 172 C.
c
k1 is a unitless value which corresponds to the unprecipitated fraction. For this analysis, it is usually 0.995 and ln 0.995 is 0.005 01.
d
% IACS is used to describe percent of conductivity relative to copper and IACS indicates International Annealed Copper Standard.
b

2. Conversely, higher Q-values are obtained with slower quench rates and are associated with lower
strength values.
3. An alloy with a low rate of precipitation will produce a lower quench factor Q than an alloy with
a high precipitation rate at the same cooling rate.
4. Quench factors calculated for different alloys might be different even if similar section sizes are
cooled in the same quenchant, because quench factors take into account individual alloy precipitation kinetics by means of the equation describing the C-curve CT function for each alloy.
5. Solute elements are precipitated during cooling from the solution treating temperature at high
Q-values. As a consequence, an improperly quenched alloy may not properly harden during aging,
and it may be susceptible to intergranular corrosion, stress corrosion or exfoliation.
Staley et al. subsequently found that min varied strongly with the isothermal hold temperature. Therefore, an alternative property predictive equation was developed which includes the temperature dependence of min and is therefore more appropriate for use with continuous cooling processes as ti 0 the
equation for d / dt becomes 12:
1
d
= minT
dt
CT

10

where:
CT is defined in Eq 5, and
strength after time t at temperature T.

Future Work
Property Predictions
Recently, improvements over the initial approach reported by Evancho and Staley 4 have extended the
ability of QFA to predict post-quenched physical properties of wrought and cast aluminum alloys. The
classical approach was limited to the ability to predict properties when the maximum loss did not exceed
1015 % 20.
As discussed previously, the classical QFA approach was based on the use of isothermal kinetics to
model transformations occurring under nonisothermal conditions. However, Cahns additive model has
been shown to be unsuitable to model all quench paths used for the thermal processing of aluminum alloys
because both diffusion controlled growth following site saturation and dissolution of single particles
during reheating satisfy the additivity condition used to model the isokinetic reaction if long, soft impingement effects can be ignored. The process becomes even more complicated if nucleation and growth
processes are considered as well 31. This means that the precipitation rate is a function of the amount of
Copyright by ASTM Int'l (all rights reserved); Sun Mar 7 19:29:31 EST 2010
Downloaded/printed by
George Totten (GE Totten Associates Inc) pursuant to License Agreement. No further reproductions authorized.

10 JOURNAL OF ASTM INTERNATIONAL

solute remaining in the matrix and therefore the amount of precipitation that can occur at that temperature.
Cahns model assumes that the precipitation rate is dependent only on the temperature and the amount
transformed 20. This problem was addressed by Staley and Tiryakiolu and using a nonisokinetic model
32.
Cahns model assumes that min = 0, which means that QFA predicts properties in the upper portion of
the C-curve approaching 0 as the fraction of the maximum property value decreases. This limitation was
addressed with a nonisokinetic model where min is related to the slope of the solvus 20. This model is
based on the assumption that strength and solute content exhibit a linear relationship and that the ability of
an alloy to produce properties decreases incrementally j over each individual time interval t j.

j = j1 minT j 1 exp

t j
CtT

11

where:
j
t j
CtT
minT
j1

loss of incremental strengthening capability MPa,


time interval s,
criticaltime s,
minimum Rp0.2 MPa, and
Eq 12 MPa.
n=j1

j1 = maxT j = max

12

n=1

The Rp0.2, strength after quenching is defined as max sum of incremental losses as shown in Eq
13 20:
j=n

= max j

13

j=1

Since it is assumed that Rp0.2 and the solute content exhibit a linear relationship, the Rp0.2 of the alloy
will be equivalent to the intrinsic annealed Rp0.2 of the alloy plus a strength factor related to the solute
content s:
= int + As

14

The temperature below, which is insufficient for the driving force for precipitation is Tint and the solute
content at this temperature, is sint and is related to min by 20:
min = int + As sint

15

The change in solubility of the solute s with respect to temperature T is defined by Eq 16:

s = exp
where:
H0
S
R
T

H0
S
exp
RT
R

16

standard enthalpy J/mol,


standard entropy J/mol K,
universal gas constant J/mol K, and
temperature K.

Flynn and Robinson showed that Eq 17 could be derived by combining Eq 15 and Eq 16 20:

min = int + k6 exp

k7
k7
exp
T
Tint

17

where k6 and k7 are constants.


If values for min are obtained for experimentally from Rp0.2 for different isothermal hold temperatures
followed by cold water quenching and then aging, it is then possible to plot min versus temperature as
Copyright by ASTM Int'l (all rights reserved); Sun Mar 7 19:29:31 EST 2010
Downloaded/printed by
George Totten (GE Totten Associates Inc) pursuant to License Agreement. No further reproductions authorized.

KAVALCO AND CANALE ON REVIEW OF QUENCH FACTOR ANALYSIS 11

FIG. 7Variation of min with temperature (7010-T76).


shown in Fig. 7 20. The slope of this data follows the same trend as the equilibrium diagram. From these
data, it is possible to obtain values for k6 and k7 in addition to Tint. Then with these data, it is possible to
calculate min at each incremental time step t j. If these relationships for min as a function of temperature
are with in Eq 5, the k1 is not used since the C-curve is no longer representative of an isothermal process.
The application of this nonisokinetic model is especially significant since it extends the capability of
the use of QFA for property prediction from up to a loss of 1015 % reported by Evanch and Staley 4 up
to approximately 70 % loss of properties 20.
C-Curve Generation: Derivation from Jominy Curve Data
One of the reasons for relatively limited use of quench factors up to the present time is the relative
unavailability of C-curves for the alloy of interest. The principal reason for this is that experimental
determination of C-curves is a time-consuming process. This is illustrated by the following example of the
experimental procedure used by Robinson et al. to obtain C-curves for 7010-T6 forgings involved heating
a test specimen, typically multiple as many as ten test specimens of the alloy of interest in a forced air
furnace at the appropriate solutionizing temperature and soaking time with respect to section size of the
alloy being studied 33. A test specimen could be a tensile bar obtained from a forging or casting of the
alloy of interest. Typically, the test specimen is sufficiently small to assure temperature uniformity during
isothermal heat treatment. After solution treatment for 2 h at For 7010 forgings, the test specimens were
solutionized for 2 h at 475 C, at which point the test specimen is rapidly transferred, within approximately 5 s, to a salt bath composed of potassium nitrate and sodium nitrite mixture and maintained within
5 C of the desired isothermal holding time and temperature. The isothermal holding time begins when
the temperature is within 5 C of the desired aging temperature. For 7010 forgings, isothermal heat
treatments were performed between 210 C and 450 C. These are a series of isothermal anneals or
pre-aging treatments 34. The test specimens are manually agitated in the molten salt bath.
After the test specimens have been held at the isothermal holding temperature for the desired time, it
is quenched to room temperature. Cooling rates are determined on selected test specimens by inserting a
Type K thermocouple to the geometric center of the test bar and data acquisition was performed at a 1 Hz
sampling rate over the duration of the test. After quenching, the test specimens were aged as required for
the temper designation of interest. For example, for an artificial aging T6 process, the test specimen is
heated for 24 h at 120 C in a forced-air oven. After cooling, the test specimens were tested for the
property of interest in order to construct a C-curve. If the test specimen could not be tested immediately,
it was stored at 20 C to delay natural aging 33.
This example illustrates the time-consuming and labor-intensive nature of the experimental procedure
Copyright by ASTM Int'l (all rights reserved); Sun Mar 7 19:29:31 EST 2010
Downloaded/printed by
George Totten (GE Totten Associates Inc) pursuant to License Agreement. No further reproductions authorized.

12 JOURNAL OF ASTM INTERNATIONAL

FIG. 8Schematic illustration of the Jominy end-quench test rig [36].

used to obtain C-curves to date. Therefore, it would be highly desirable to identify an alternative procedure
to acquire C-curves. This is particularly important the C-curve produced is alloy composition dependent.
Fortunately, in addition, to determining C-curve constants experimentally 33, an alternative procedure
has been recently developed which permits the determination of C-curves experimentally by modeling
Jominy data obtained for the aluminum alloy 25,3537. A brief, but reasonably detailed, overview of this
procedure will be outlined in this section.
MacKenzie 25, Tanner and Robinson 38, and Ma et al. 36 have described Jominy end-quench test
procedures to determine hardness and cooling rate as a function of distance from the quenched end of the
Jominy test specimen. The Jominy end-quench test specimen was typically a round bar which was machined to 25 mm diameter by 100 mm long with a head at the top end of the bar of 31 mm diameter by
3 mm thick. Flats of 1 mm depth were machined on opposite sides of the test bar prior to conducting the
test. A standard Jominy end-quench test rig such as that described in ASTM A255-07 Standard Test
Methods for Determining Hardenability of Steel. The head of the test bar is supported by a plate in the
Jominy test rig, a schematic of which is shown in Fig. 8 36. The water spray is emitted from a 12.7 mm
diameter orifice which is located 12.7 mm from the end of the Jominy test bar. The heat transfer occurring
in the test bar is one-dimensional because only one end of the bar is being quenched.
After solutionizing, the bar was transferred to the Jominy end-quench test rig within 5 s and quenched
with water for 5 min, at which time the bar was removed from the test rig, dried, and cooled in liquid
nitrogen at 196 C to prevent any aging from occurring. After stabilization, the test bar was refrigerated
at 22 C for a minimum of 12 h before testing.
Hardness measurements were taken on a cold 22 C aluminum slab so that the temperature of the
Jominy bar did not exceed 0 C during hardness testing. Hardness measurements were taken at 2 mm from
the edge of the flats and at 4 mm increments from the quenched end. After hardness measurements were
taken, the test bar was refrigerated again.
Another Jominy test bar was machined with a 1.5 mm diameter holes at the center in which 1.5 mm
diameter. Type K thermocouples were inserted as shown in Fig. 9. These thermocouples were used to
obtain cooling curve data. The data acquisition rates for temperature was 3.3 Hz and one set of data was
obtained at 10 Hz for the thermocouple located 3 mm from the quenched end. The heating and quenching
process was the same as that described for the test specimen used for hardness measurements 35.
Ma et al. 36 used Jominy test bar with more type K thermocouples than the bar used by Tanner and
Robinson 35. Thermocouple positions used in Mas study at positions shown in Table 6. The 356
aluminum casting alloy used in their study was solutionized at 540 C for 4 h. The time-temperature
cooling curve and cooling rate curves obtained are shown in Fig. 10. At 3.2 mm from the quench end, the
maximum cooling rate is approximately 150 C / s, which is equivalent to a water quench. The maximum
cooling rate decreases to approximately 5 C / s at 63.5 mm from the quench, end which is similar to the
Copyright by ASTM Int'l (all rights reserved); Sun Mar 7 19:29:31 EST 2010
Downloaded/printed by
George Totten (GE Totten Associates Inc) pursuant to License Agreement. No further reproductions authorized.

KAVALCO AND CANALE ON REVIEW OF QUENCH FACTOR ANALYSIS 13

FIG. 9Dimensions of Jominy test specimen used by Tanner and Robinson [35].

cooling rate attainable from an air quench. These data show that with one experiment using a single
thermocouple-instrumented Jominy test probe, it is possible to obtain a wide range of cooling rates.
TannerRobinsons approach to obtaining cooling curves using fewer thermocouples can also be used
to obtain cooling curves and cooling rate curves at any point on the Jominy bar by using a finite element
analysis FEA 35. In this case, heat transfer model of the Jominy end quench bar is constructed using a
platform such as AQAQUS, The symmetry of Jominy end-quench test bar may be modeled using one
quarter of the test bar with heat diffusion elements of type DC3D4 four-noded linear tetrahedron for the
head of the sample and DC3D8 eight-noded quadratic brick elements for the main shaft of the sample
35. Properties for thermal conductivity, density, and specific heat were obtained from the literature. The
main boundary condition was established from cooling curves measured at 3 mm from the end of the
Jominy end-quench bar. Radial heat transfer from the unquenched sides of the specimen was ignored as
previous work has indicated that any heat transfer that occurs to the surrounding air does not affect the
hardness measured 35,39.
The next step is to develop a correlation between strength and hardness. Although good linear correlations between hardness and strength may be obtained if the same aging treatment is used after different
continuous cooling quench rates 15. However, since QFA was originally derived based on actual strength
measurements, it has been recommended that if any property other than strength is used that there exist a
linear correlation between that property and strength 36,40.
An often used alternative to strength is hardness which are typically based on the depth of indentation.
These hardness values, like Rockwell hardness, are based on an arbitrary number. However, an alternative
is Meyer hardness which may be referred to as mean pressure is proposed to possess a physical
meaning and is defined as the work to create a unit volume of elastoplastic residual impression 40.
Meyer hardness P is defined numerically as:

TABLE 6Thermocouple positions in the instrumented Jominy test bar described by Ma et al. [36].

mm
in.
1 / 16

3.20
2

6.40
4

9.50
6

12.7
8

Thermocouple Position
15.8
22.2
10
14

31.7
20

38.1
24

50.8
32

63.5
40

FIG. 10Cooling curves (a) and cooling rate curves (b) at different locations of a Jominy end quench bar
prepared from cast aluminum alloy A356 [36].
Copyright by ASTM Int'l (all rights reserved); Sun Mar 7 19:29:31 EST 2010
Downloaded/printed by
George Totten (GE Totten Associates Inc) pursuant to License Agreement. No further reproductions authorized.

14 JOURNAL OF ASTM INTERNATIONAL

FIG. 11Meyer hardness as a function of distance from the quenched end of a Jominy bar of a cast
aluminum 356 alloy [36].
P = 4L
d2

18

where:
P Meyer hardness kg/ mm2,
L load on the indentor kg, and
D diameter of the indentation mm.
The correlation between Rockwell B hardness RHB and Meyer hardness was developed by
Tiryakiolu and Campbell 40:
d = 1263 5270 103 RHB

19

Equations 18 and 19 permit the calculation of Meyer hardness from Rockwell B hardness.
For the work described by Ma et al., Meyer hardness was calculared from RHB measurements and then
plotted as a function of distance from the quenched end of the Jominy bar to yield the curve shown in
Fig. 11 36.
Ma et al. used Eq 20 to determine the relationship between the quench factor Q and Meyer hardness
36:
min
= expk1Q
max min

20

The value for max was taken as the Meyer hardness at the quenched end of the Jominy bar since only
limited precipitation will occur under these conditions and, as described above, Meyer hardness is known
to exhibit a correlation with strength. The value for min was established by solutonizing the Jominy bar at
540 C in a conventional furnace and then transferring the bar to a preheated fluidized furnace also at
540 C at which time the burners are turned off but the blowers left on and then cooling for 20 h to allow
precipitation to reach equilibrium and then the bar is quenched in water followed by aging at 165 C for
6 h. The bar is sectioned and ten hardness readings were taken across the section and averaged. This value
was used as the minimum Meyer hardness. Values for are the incremental Meyer hardnesses along the
Jominy bar.
At this point, the natural logarithm is taken on both sides of Eq 20, which yields Eq 21.

ln

P P
1
min
ln
k1
P P
max
min

= n ln Q

Copyright by ASTM Int'l (all rights reserved); Sun Mar 7 19:29:31 EST 2010
Downloaded/printed by
George Totten (GE Totten Associates Inc) pursuant to License Agreement. No further reproductions authorized.

21

KAVALCO AND CANALE ON REVIEW OF QUENCH FACTOR ANALYSIS 15

FIG. 12Fitting of experimental yield strength data to Eq 21 for aluminum cast alloys 356 (triangles) and
357 (circles). The y-intercept yields the Avrami exponent.
It was previously demonstrated that a plot of the left side of Eq 21 using yield strength instead of
Meyer hardness versus the lnQ from the calculated quench factor using measured cooling curves with the
CT constants with Eqs 5 and 6 would yield a linear relationship for aluminum cast alloys 356 and 357. The
slope or Avrami constants were determined to be n = 45 and n = 63 for aluminum alloys 356 and 357,
respectively 41. Although there was considerable scatter in the data, the linearity of this relationship is
shown in Fig. 12 41. Since Meyer hardness was shown to correlate with strength, it is expected that a
similar linear correlation would result.
As discussed previously, multiple linear regression analysis is used to iteratively determine the best set
of values for k2 k5. The value for k1 was equal to the logarithm of the percent transformed 0.005 01,
which corresponds to assumed to be 0.5 % transformed. Ma et al. minimized the difference between the
predicted and measured property using a least-squares routine to obtain the best linear relationship between
a function of experimentally measured properties and calculated quench factors Eq 21 36. Quench
factors were calculated according to Eq 5 using the cooling curves shown in Fig. 10, which correspond to
the different Jominy position for which Meyer hardnesses were obtained. In this study, initial values for
k2 k5 were assumed and then iteratively varied until a minimum error was obtained. A plot of the data as
shown in Fig. 13 was obtained and this process was repeated until the fitted line passes as close as possible
through the origin as shown in Fig. 13 36.
Using these computational methods and Jominy end-quench experiments, it is possible to utilize QFA
to identify the best alloy and quench rate for a given application. Using either a database of cooling curves

FIG. 13An example best-fit curve for quench factor analysis of cast aluminum alloy A356 [36].
Copyright by ASTM Int'l (all rights reserved); Sun Mar 7 19:29:31 EST 2010
Downloaded/printed by
George Totten (GE Totten Associates Inc) pursuant to License Agreement. No further reproductions authorized.

16 JOURNAL OF ASTM INTERNATIONAL

for different section sizes or by utilizing FEA simulations as discussed here, it is possible to calculate
cooling curves at any point in a section or to examine section size effects when coupled with QFA. Finally,
by obtaining Jominy data as discussed here for different alloy chemistries, it is possible to examine the
effect of compositional variation of a particular alloy on properties using QFA.
Limitations of QFA
Although QFA has been, and continues to be, used successfully for property prediction of heat treatable
aluminum alloys, there are limitations to the predictive accuracy which are attributable to the assumptions
made in the theoretical development 15. Three of these limitations, as addressed by Rometsch et al.,
include:
1. Classical QFA assumes that strength varies linearly with solute concentration. However, it is now
known that strengthening contributions are due to both shearable and nonshearable precipitates
that are proportional to the square root of the volume fraction, which is contrary to the original
QFA assumption. Rometsch et al. proposed that Eq 5 from the classical QFA approach should be
replaced by Eq 22 to account for this nonlinearity 15;
min
= expk1Q1/2
max min

22

2. Classical QFA assumes that transformational kinetics are described by a special case of the Avrami
equation where the Avrami coefficient is n = 1. When Evancho and Staley analyzed the 7075-T6
and 2024-T4 interrupted quenching data of Fink and Willey, they obtained linear correlations with
slope of 1 for plots of log10 / max versus isothermal hold time. However, diffusion controlled
nucleation and growth theories indicate that n 1.5 is not possible for reactions that involve
growth in three dimensions precipitates nucleating and growing within grains 15. Therefore
n = 1 is mostly related to formation of non-hardening precipitates within grains during quenching.
Rometsch et al. have proposed using Eq 23, a form of the StarinkZahra equation as an alternative
to the Avrami approach to describing transformation kinetics to accommodate values of n 1, as
an alternative to Equation 22 15:

min
k1Qn
=
+1
max min
i

i/2

23

where:
n Avrami coefficient, and
i impingement factor 42.
3. The minimum strength in the Avrami equation. By neglecting minimum strength min the ability
to successfully predict strength loses accuracy. As discussed above, Staley showed that at lower
values of / max the predictions are improved by introducing min as a constant related to strength
in the absence of hardening precipitates. However, Rometsch developed an alternative nonalloyspecific approach to address this problem based on temperature-dependent solubility to describe
the variation of min with temperature. The reader is directed to Ref. 15 for a rigorous treatment
of this approach which permits the inclusion of aging temperature variation, solution treatment
temperature and alloy composition to be included in the model.
Tiryakiolu and Shuey have developed a new QFA model that addresses many of the deficiencies of
the classical model or even the various improvements such as establishing process-structure-property
relationships, includes multiple quench precipitates, and utilizes thermodynamic and calculated data for
some coefficients 43. In this new model, the amount of each quench precipitate is represented by a
unitless microstructural state variable S volume fraction. For an arbitrarily long isothermal hold, in the
absence of competing precipitates, the limiting value Seq is defined as:

SeqT = 1 exp

H 1 1

R k4 T

T k4

where:
Copyright by ASTM Int'l (all rights reserved); Sun Mar 7 19:29:31 EST 2010
Downloaded/printed by
George Totten (GE Totten Associates Inc) pursuant to License Agreement. No further reproductions authorized.

24

KAVALCO AND CANALE ON REVIEW OF QUENCH FACTOR ANALYSIS 17

k4 solvus temperature K,
H precipitation enthalpy for quench precipitate J/mol, and
Seq limiting value of a unitless microstructural variable specifying the amount of quench
precipitate S at equilibrium after an arbitrarily long isothermal holding time at a
given temperature. Note: A state variable is an element of the set of variables that
describe the state of a dynamic system. Typical state variables in a thermodynamic system
include temperature, pressure, enthalpy, entropy, and specific volume.
The evolution of quench precipitates dS / dt is modeled by:
dS Seq S
=
dt
tc

25

where:
tc critical time s to nucleate and grow quench precipitates and is determined from:

tc = k2 exp

k3k24
k5
2 +
RTk4 T
RT

26

where:
k2
k3
k5
R
T

coefficient related inversely to number of nucleation sites s,


coefficient related to energy barrier to heterogeneous nucleation J/mol,
stoichiometrically averaged activation energy for diffusion J/mol,
gas constant, 8.3143 J / mol K, and
temperature K.

The change in the precipitates that are formed is calculated from:

Si = Seq Si1 1 exp

ti
tc

27

At the conclusion of the quenching process, the total precipitates formed is determined by summing the
incremental amounts:
S = Si

28

y = max k jS j

29

The yield strength is calculated from:

where:
j quench precipitates modeled,
max strength obtained at an infinite quench rate ksi, and
k j strength coefficient.
Equation 29 assumes that yield strength depends linearly on the amount of solute available after
quench, which is consistent with the classical model 4,12. This new Tiryakiolu and Shuey model is an
expansion of the classical versions which considered only one type of quench precipitate same
stoichiometry and site by using similar equations 43. If only one precipitate is considered with a single
quench precipitate, Eqs 2427 and 29 are equivalent to the classical equations discussed above.
Conclusions
To achieve the ideal quench, low values of Q are desirable, because they are associated with high quench
rates, minimum precipitation during cooling and high yield strengths. High values of Q can lead to an
Copyright by ASTM Int'l (all rights reserved); Sun Mar 7 19:29:31 EST 2010
Downloaded/printed by
George Totten (GE Totten Associates Inc) pursuant to License Agreement. No further reproductions authorized.

18 JOURNAL OF ASTM INTERNATIONAL

improperly quenched alloy that may not properly harden during aging and may be susceptible to intergranular corrosion, stress corrosion, or exfoliation.
Some improvements in the modeling of the QFA have been developed although classical QFA has
been used successfully for property prediction of heat treatable aluminum alloys. However, there are
limitations to the predictive accuracy of the classical equations which are attributable to the assumptions
made in the theoretical development. An overview of newer predictive methodologies that address these
limitations and expand the potential utility of QFA have been provided here.

References
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19

Croucher, T., Quenching of Aluminum Alloys: What This Key Step Accomplishes, Heat Treating,
Vol. 14, 1982, p. 2021.
Chevrier, J. C., Simon, A., and Beck, G., Optimal Cooling Rate and Process Control in Metallic
Parts Heat Treatment, Heat and Mass Transfer in Metallurgical Systems, Taylor & Francis, London,
1981, Vol. 9, pp. 535544.
Archambault, P., Bauvaist, J., Chevrier, J. C., and Beck, G., A Contribution to the 7075 Heat
Treatment, Mater. Sci. Eng., Vol. 43, 1980, pp. 16.
Evancho, J. W. and Staley, J. T., Kinetics of Precipitation in Aluminum Alloys During Continuous
Cooling, Metall. Trans., Vol. 5, 1974, pp. 4347.
Fink, W. L. and Willey, L. A., Quenching of 75S Aluminum Alloy, Trans. AIME, Vol. 175, 1948,
pp. 414427.
ASM Specialty Handbook: Aluminum and Aluminum Alloys, ASM International, Materials Park, OH,
1993, p. 23.
Howard, R., Bogh, N., and MacKenzie, D. S., Heat Treating Processes and Equipment, in Handbook of Aluminum, Totten, G. E. and MacKenzie, D. S. Eds., CRC Press, Boca Raton, FL, 2003, Vol.
1, pp. 881970.
Bates, C. E. and Totten, G. E., Procedure for Quenching Media Selection to Maximize Tensile
Properties and Minimize Distortion in AluminumAlloy Parts, Heat Treat. Met., Vol. 16, No. 4,
1988, pp. 8998.
Dolan, G. P., Robinson, J. S., and Morris, A. J., Quench Factors and Residual Stress Reduction in
7175-T73 Plate, Proceed. Materials Solution Conference, 58 November 2001, Indianapolis, IN,
ASM International, Materials Park, OH, USA, 2001, pp. 213218.
Quench Factor Analysis in Aluminum Properties and Physical Metallurgy, Hatch, J. E., Ed., ASM
International, Materials Park, OH, 1984, pp. 159164.
Staley, J. T., Modeling Quenching of Precipitation Strengthened Alloys: Application to an
Aluminum-Copper-Lithium Alloy, Ph.D. Thesis, Drexel University, 1989.
Staley, J. T., Doherty, R. D., and Jaworski, A. P., Improved Model to Predict Properties of Aluminum Alloy Products after Continuous Cooling, Metall. Trans. A, Vol. 24A, 1993, pp. 24172427.
Staley, J. T., Quench Factor Analysis of Aluminum Alloys, Mater. Sci. Technol., Vol. 3, 1987, pp.
923935.
Bates, C. E., Quench Optimization for Aluminum Alloys, AFS Trans., Vol. 102, 1994, pp. 1045
1054.
Rometsch, P. A., Starink, M. J., and Gregson, P. J., Improvements in Quench Factor Modeling,
Mater. Sci. Eng., Vol. A339, 2003, pp. 255264.
Liu, S. D., Zhang, X. M., Huang, Z. B., and You, J. H., Prediction of Hardness of Aluminum Alloy
7055 by Quench Factor Analysis, Mater. Sci. Forum, Vol. 546549, 2007, pp. 881884.
Hall, D. D. and Mudawar, I., Prediction of Mechanical Properties of Complex-Shaped Aluminum
Alloy Parts Subjected to Spray Quenching, ASME Thermal Processing of Materials: ThermoMechanics, Controls and Composites, HTD-Vol. 289, ASME, New york, 1994, pp. 719.
Bernardin, J. D. and Mudawar, I., Validation of the Quench Factor Technique in Predicting Hardness
in Heat Treatable Aluminum Alloys, Int. J. Heat Mass Transfer, Vol. 38, No. 5, 1995, pp. 863873.
Rozzi, J. C., Klinzing, W. P., and Mudawar, I., Effects of Spray Configuration on the Uniformity of
Cooling Rate and Hardness in the Quenching of Aluminum Parts with Nonuniform Shapes, J.
Mater. Eng. Perform. Vol. 1, No. 1, 1992, pp. 4960.

Copyright by ASTM Int'l (all rights reserved); Sun Mar 7 19:29:31 EST 2010
Downloaded/printed by
George Totten (GE Totten Associates Inc) pursuant to License Agreement. No further reproductions authorized.

KAVALCO AND CANALE ON REVIEW OF QUENCH FACTOR ANALYSIS 19

20 Flynn, R. J. and Robinson, J. S., The Application of Advances in Quench Factor Analysis Property
Prediction to the Heat Treatment of 7010 Aluminium Alloy, J. Mater. Process. Technol., Vol.
153154, 2004, pp. 674680.
21 J-S., Kim, Hoff, R. C., and Gaskell, D. R., A Quench Factor Analysis of the Influence of Water
Spray Quenching on the Age-Hardenability of Aluminum Alloys, in Materials Processing in the
Computer Age, Voller, R., Stachowicz, M. S., and Thomas, B. G., Eds., The Minerals, Metals and
Materials Society, Warrendale, PA, 1991, pp. 203221.
22 Bates, C. E., Predicting Properties and Minimizing Residual Stress in Quenched Steel Parts, J.
Heat Treatment, Vol. 6, 1988, pp. 2745.
23 Bates, C. E. and Totten, G. E., Quench Severity Effects on the As-Quenched Hardness of Selected
Alloy Steels, Heat Treat. Met., Vol. 2, 1992, pp. 4548.
24 Avrami, M., Kinetics of Phase Change, J. Chem. Phys., Vol. 7, 1939, pp. 11031112.
25 MacKenzie, D. S., Quench Rate and Aging Effects in Al-Zn-Mg-Cu Aluminum Alloys, Ph.D.
Thesis, University of Missouri-Rolla, Rolla, MO, 2000.
26 Totten, G. E., Webster, G. M., and Bates, C. E., Quench Factor Analysis: Step-By-Step Procedures
for Experimental Determination, in Proceedings of the 1st International Non-Ferrous Processing
and Technology Conference, Bains, T. and MacKenzie, D. S., Eds., ASM International, Materials
Park, OH, March 1012, 1997, pp. 305312.
27 Shuey, R. T., Tiryakiolu, M., and Lippert, K. B., Mathmatical Pitfalls in Modeling Quench Sensitivity for Aluminum Alloys, Proceed. 1st International Symposium on Metallutgical Modeling for
Aluminum Alloys, 1315 October 2003, Pittsburgh, PA, USA, ASM International, Materials Park,
OH, 2003, pp. 4753.
28 Tiryakiolu, M. and Shuey, R. T. Metallurgical Modeling for Aluminum Alloys, Proceedings
Materials Solutions Conference 2003: 1st International Symposium on Metallurgical Modeling for
Aluminum Alloys, ASM International, Materials Park, OH, 2003, pp. 3945.
29 Cahn, J. W., The Kinetics of Grain Boundary Nucleated Reactions, Acta Metall., Vol. 4, 1956, pp.
449459.
30 Swartzendruber, L., Boettinger, W., Ives, L., Coriell, S., Ballard, D., Laughlin, D., Clough, R.,
Biancanieilo, F., Blau, P., and Cahn, J., Nondestructive Evaluation of Nonuniformities in 2219
Aluminum Alloy PlateRelationship to Processing, US Department of Commerce, National Bureau
of Standards Technical Report NBSIR 802069, December 1980.
31 Grong, O. and Myhr, O. R., Additivity and Isokinetic Behaviour in Relation to Diffusion Controlled
Growth, Acta Mater., Vol. 48, 2000, pp. 445452.
32 Staley, J. T. and Tiryakiolu, M. The Use of TTP Curves and Quench Factor Analysis for Property
Prediction in Aluminum Alloys, Proceed. of the Materials Solutions Conference of ASM International, ASM International, Materials Park, OH, 2001, pp. 614.
33 Robinson, J. S., Cudd, R. L., Tanner, D. A., and Dolan, G. P., Quench Sensitivity and Tensile
Property Inhomogeneity in 7010 Forgings, J. Mater. Process. Technol., Vol. 119, 2001, pp. 261
267.
34 Swartzendruber, L. J., Boettinger, W. J., Ives, L. K., Coriell, S. R., and Mehrabian, R., Relationship
Between Process Variables, Microstructure and NDE of a Precipitation-Hardened Aluminum Alloy,
in Proceedings of the Nondestructive Evaluation of Microstructure and Reliable Strategies, Buck, O.
and Wolf, S. M., Eds., 1981, pp. 253271.
35 Tanner, D. A. and Robinson, J. S., Effect of Precipitation During Quenching on the Mechanical
Properties of the Aluminum Alloy 7010 in the W Temper, J. Mater. Process. Technol., Vol. 153
154, 2004, pp. 9981004.
36 Ma, S., Maniruzzaman, M. D., MacKenzie, D. S., and Sisson, R. D., A Methodology to Predict the
Effects of Quench Rates on Mechanical Properties of Cast Aluminum Alloys, Metall. Mater. Trans.
B, Vol. 38B, 2007, pp. 583589.
37 Ma, S. H. and Sisson, R. D., Modeling Heat Treatment of Age Hardenable Cast Aluminum Alloys,
Int. Heat Treatment Surf. Eng., Vol. 1, No. 2, 2007, pp. 8187.
38 Tanner, D. A. and Robinson, J. S., Residual Stress Magnitudes and Related Properties in Quenched
Aluminium Alloys, Mat. Sci. Technol., Vol. 22, No. 1, 2006, pp. 7785.
39 Newkirk, J. and MacKenzie, D., The Jominy End Quench for Light-Weight Alloy Development, J.
Copyright by ASTM Int'l (all rights reserved); Sun Mar 7 19:29:31 EST 2010
Downloaded/printed by
George Totten (GE Totten Associates Inc) pursuant to License Agreement. No further reproductions authorized.

20 JOURNAL OF ASTM INTERNATIONAL

Mater. Eng. Perform., Vol. 9, No. 4, 2000, pp. 408415.


40 Tiryakiolu, M. and Campbell, J. On Macrohardness Testing of Cast Al-Si-Mg Alloys: I: Geometrical and Mechanical Aspects, Mater. Sci. Eng., A, Vol. A361, 2003, pp. 232239.
41 Rometsch, P. A. and Schaeffer, G. B., Quench Modeling of Al-7Si-Mg Casting Alloys, Int. J. Cast
Metals, Vol. 12, 2000, pp. 431439.
42 Starink, M. J., On The Meaning of The Impingement Parameter in Kinetic Equations for Nucleation
and Growth Reactions, J. Mater. Sci., Vol. 36, 2001, pp. 44334441.
43 Tiryakiolu, M. and Shuey, R. T., Quench Sensitivity of an Al-7 Pct Si-0.6 Pct Mg Alloy: Characterization and Modeling, Metall. Mater. Trans. B, Vol. 38B, 2007, pp. 575582.

Copyright by ASTM Int'l (all rights reserved); Sun Mar 7 19:29:31 EST 2010
Downloaded/printed by
George Totten (GE Totten Associates Inc) pursuant to License Agreement. No further reproductions authorized.

Erratum for JAI102131, Journal of ASTM International, Volume 6, Issue 5 (May 2009)
EVOLUTION OF QUENCH FACTOR ANALYSIS: A REVIEW, Patricia Kavalco and Lauralice Canale;

STP 1523, Quenching and Cooling, Residual Stress and Distortion Control.
Corrections:
1. Reference [21], the correct name is Kim, J-S., and not J-S., Kim
2. All the references from table 4 are incorrect. The correct ones are shown in the table below. Note that there
are 3 new references that must be added.
3.

The Alloy 2017-T4 (in the table) should be replaced by 2017A-T4.


TABLE 4 - Coefficients for calculating quench factors at 99.5% of attainable yield strength.

Alloy

k1

k2

k3

k4

k5

Calculated

(s)

(J/mol)

(K)

(J/mol)

Range

Reference

(C)
-20

5780

897

1.90x10

425-150

[20]

-19

5190

850

1.8x10

425-150

[4]

425-150

[4]

7010-T76

-.000501

5.6x10

7050-T76

-0.00501

2.2x10

7075-T6
7075-T73
7175-T73
2017-T4
2024-T6

-0.00501
-0.00501
-0.00501
-0.00501
-0.00501

2024-T851

-0.00501

2219-T87

-0.00501

6061-T6
356-T6
357-T6
Al-2.7Cu-

-0.00501
-0.0066
-0.0062
-0.0050

-13

4.1x10

-13

1.37x10

-9

1.8x10

-21

6.8x10

-12

2.38x10
1.72x10

1069
526
978
1310

780
737

1.4x10

425-150

[44]

1.017x10

425-150

[9]

822

2.068x10

425-150

[44]

840

1.47x10

425-150

[45]

425-150

[46]

425-150

[30]

425-150

[44]

750

1.37x10

-11

45

750

3.2x10

-7

200

900

2.5x10

0.28x10

-8

5.1x10

-4

3.0x10

-10

1.1x10

-8

1.8x10

1050

412
61
154
1520

750
764
750
870

9.418x10

1.3x10

425-150

[39]

425-150

[39]

425-150

[39]

1.31x10
1.02x10

1.6Li-T8
a. k1 is a unitless value which corresponds to the unprecipitated fraction. For this analysis, it is usually 0.995 and ln 0.995 is
-0.00501.

[44] Dolan, G. P., and Robinson, J. S., Residual Stress Reduction in 7175-T73, 6061-T6 and 2017A-T4
Aluminum Alloys Using Quench Factor Analysis, Journal of Materials Processing Technology, 2004, Vol. 153
154, 2004, p. 346351.
[45] Hall, D. D., and Mudawar, I., Optimization of Quench History of Aluminum Parts for Superior Mechanical
Properties, Int. J. Heat Mass Transfer, 1996, Vol. 39, No. 1, p. 81-95.

Copyright by ASTM Int'l (all rights reserved); Sun Mar 7 19:29:31 EST 2010
Downloaded/printed by
George Totten (GE Totten Associates Inc) pursuant to License Agreement. No further reproductions authorized.

[46]

Ives,

L.

K.,

Swartzendruber,

L.,

Boettinger,

W.

J.,

Rosen,

M.,

and

Ridder,

S.

D.,

Processing/Microstructure/Property Relationships in 2024 Aluminum Alloy Plates, US Department of Commerce,


National Bureau of Standards Technical Report NBSIR 83-2669, January 1983.

Copyright by ASTM Int'l (all rights reserved); Sun Mar 7 19:29:31 EST 2010
Downloaded/printed by
George Totten (GE Totten Associates Inc) pursuant to License Agreement. No further reproductions authorized.

Anda mungkin juga menyukai