Вы находитесь на странице: 1из 45

Previous Page

(1-103)
Equation (1-103) has been plotted in Fig. 1.36.
Example (LU)
Consider butt welding of a 2mm thin plate of austenitic stainless steel with covered electrodes
(SMAW) under the following conditions:

Calculate the retention time within the critical temperature range for chromium carbide
precipitation (i.e. from 650 to 8500C) for points located at the 8500C isotherm.
Solution

If we use the melting point of the steel as a reference temperature, the parameter n3/5 becomes:

A comparison with Fig. 1.35 shows that the assumption of 1-D heat flow is justified when
Qp< 1. Hence, the total time spent in the thermal cycle from 6 = 0.43 (T = 6500C) to 0p = 0.56
(Tp = 8500C) and down again to 0 = 0.43 can be read from Fig. 1.36. Taking the ordinate 0/0p
equal to 0.76, we obtain:

which gives

and

1.10.4 Medium thick plate solution


In a real welding situation the assumption of three-dimensional or two-dimensional heat flow
inherent in the Rosenthal equations is not always met because of variable temperature gradients in the through thickness z direction of the plate.
Model (after Rosenthal14)

The general medium thick plate model considers a point heat source moving at constant speed
across a wide plate of finite thickness d. With the exception of certain special cases (e.g.
watercooling of the back side of the plate), it is a reasonable approximation to assume that the

plate surfaces are impermeable to heat. Thus, in order to maintain the net flux of heat through
both boundaries equal to zero, it is necessary to account for mirror reflections of the source
with respect to the planes of z = 0 and z = d. This can be done on the basis of the 'method of
images' as illustrated in Fig. 1.37. By including all contributions from the imaginary sources
...2q__2 , 2g_i , 2q\ , 2q2 ,...located symmetrically at distances 2id below and above the upper
surface of the plate, the pseudo-steady state temperature distribution is obtained in the form of
a convergent series*:

(1-104)

where
Note that equation (1-104) is simply the general Rosenthal thick plate solution (equation
(1-45)) summed for each source.

Fig. 1.37. Real and imaginary point sources on a medium thick plate.
*The number of imaginary heat sources necessary to achieve the required accuracy depends on the chosen values of
R0 and vd/2a.

By substituting the dimensionless parameters defined above into equation (1-104), we obtain:
(1-105)
where

It follows from equation (1-104) that the thermal conditions will be similar to those in a
thick plate close to the centre of the weld. Moreover, Rosenthal1314has shown on the basis of
a Fourier series expansion that equation (1-104) converges to the general thin plate solution
(equation (1-81) for points located sufficiently far away from the source. However, at intermediate distances from the heat source, the pseudo-steady state temperature distribution will deviate significantly from that observed in thick plate or thin plate welding because of variable
temperature gradients in the through-thickness direction of the plate. Within this 'transition
region', the thermal programme is only defined by the medium thick plate solution (equation
(1-104)).
1.10.4.1 Dimensionless maps for heat flow analyses
Based on the models described in the previous sections, it is possible to construct a series of
dimensionless maps which provide a general outline of the pseudo-steady state temperature
distribution during arc welding.20
Construction of the maps
The construction of the maps is done on the basis of the medium thick plate solution (equation
(1-105)). This model is generally applicable and allows for the plate thickness effect in a
quantitative manner. Since the other solutions are only valid within specific ranges of this
equation, they will have their own characteristic fields in the temperature-distance or the temperature-time space. The extension of the different fields can be determined from numerical
calculations of the temperature distribution by comparing each of these models with the medium thick plate solution, using a conformity of 95% as a criterion.
Similarly, when the 95% conformity is not met between the respective solutions, the fields
are marked 'transition region'. Since any combination of dimensionless temperature, operating parameter, and plate thickness locates a point in a field, it means that the dominating heat
flow mechanism can readily be read off from the maps.
Peak temperature distribution
The variation of peak temperature with distance in the \j/(^j-direction has been numerically
evaluated from equation (1-105) for different values of the dimensionless plate thickness (8 =
vdlla). The results are shown graphically in Fig. 1.38(a) and (b) for the two extreme cases of
= 0 (z = 0) and = 8 (z = d), respectively.
An inspection of the maps reveals that the temperature-time pattern in stringer bead
weldments can be classified into three main categories:

(a)

VV

Thin plate solution


(2-D heat flow) (1-D heat flow)

(b)
Thin plate solution
1-D heat flow

V"3

(2-D heat flow)

Fig. 1.38. Peak temperature distribution in transverse direction (\|/ = \\fm) of plate; (a) Upper plate surface
(^ = 0), (b) Lower plate surface (J = 8).

1. Close to the heat source, the thermal programme will be similar to that in a thick plate
(Fig. 1.38(a)), which means that the temperature distribution is determined by equation (154). For large values of the dimensionless plate thickness, the mode of heat flow may
become essentially two-dimensional. This corresponds to the limiting case of a fast moving high power source in a thick plate (equation (1-74)). Under such conditions the slope of
the Qp/n3-\ym curves in Fig. 1.38(a) attains a constant value of-2.
2. With increasing distance from the heat source, a transition from three-dimensional to
two-dimensional heat flow may occur, depending on the dimensionless plate thickness and
the operational conditions applied. Considering the upper surface of the plate (Fig. 1.38(a)),
the extension of the transition region is seen to decrease with increasing values of 8 as the
conditions for thick plate welding are approached. The opposite trend is observed for the
bottom plate surface (Fig. 1.38(b)), since a small dimensionless plate thickness generally
results in a more rapid equalisation of the temperature gradients in the t,(z) direction. When
the curves in Fig. 1.38(b) become parallel with the jc-axis, the temperature at the bottom of
the plate reaches its maximum value. Note that within the transition region, reliable predictions of the pseudo-steady temperature distribution can only be made from the medium
thick plate solution (equation (1-105)).
3. For points located sufficiently far away from the heat source, the temperature gradients in the through-thickness direction of the plate become negligible. This implies that the
temperature distribution at the upper and lower surface of the plate is similar, and can be
computed from the thin plate solution (equation (1-83)). When the conditions for onedimensional heat flow are approached (equation (1-101)), the slope of the dp/n3-\\fm curves
in Fig. 1.38(a) and (b) attains a constant value o f - 1 .
Cooling conditions close to weld centre-line
Figure 1.39 contains a plot of the cooling programme for points located on the weld centre-line
(\j/ = = 0), as calculated from equation (1-105). A closer inspection of Fig. 1.39 reveals that
the slope of the cooling curves increases gradually from -1 to -0.5 with increasing distance
from the heat source. This corresponds to a change from three-dimensional to one-dimensional heat flow.
From Fig. 1.39 it is possible to read-off the cooling time within specific temperature intervals for a wide range of operational conditions. These results are also valid for positions
outside the weld centre-line, since the cooling curves are virtually parallel in the transverse \|/
direction of the plate. A requirement is, however, that the peak temperature of the thermal
cycle is significantly higher than the actual temperature interval under consideration.
Retention times at elevated temperatures
The retention time, Axn is defined as the total time spent in a thermal cycle from a chosen
reference temperature 0 to the peak temperature dp and down again to 9. This parameter can
readily be computed from equation (1-105) by means of numerical methods. The results of
such calculations (carried out in position = 0) are shown graphically in Fig. 1.40 for 9 = 0.5
An inspection of Fig. 1.40 reveals a complex temperature-time pattern. In this case it is not
possible to determine the exact field boundaries between the respective solutions, since the

e/n3

e/n3

Thin plate solution


(2-D heat flow) (1-D heat flow)

T= ^

e/n 3

X
Fig. 1.39. Cooling programme for points located on the weld centre-line (\\f = =0).

AXf

Fig. 1.40. Total time spent in a thermal cycle from 9 through 9p to 9 for a chosen reference temperature of
9 = 0.59p.

mode of heat flow may vary within a single thermal cycle. Hence, the extension of the different fields is not indicated in the graph. The results in Fig. 1.40 provide a systematic basis for
calculating the retention time within specific temperature intervals under various welding conditions.
Isothermal contours
Because of the number of variables involved, it is not possible to present a two-dimensional
plot of the isotherms without first specifying the dimensionless plate thickness. Examples of
calculated isotherms in different planes are shown in Figs. 1.41 and 1.42 for 8 equal to 0.5 and
5, respectively. It is evident that an increase in the dimensionless plate thickness from 0.5 to 5
has a dramatic effect on the shape and position of the isothermal contours. However, in order
to explain these observations in an adequate manner, it is necessary to condense the results into
a two-dimensional diagram. As shown in Fig. 1.43, this can be done by plotting the calculated
field boundaries in Fig. 1.38(a) at maximum width of the isotherms vs the parameters Qp/n3

and vdlla.
It is seen from Fig. 1.43 that a large plate thickness generally will favour three-dimensional
heat flow. With decreasing values of Qp/n3, the conditions for a fast moving high power source
are approached before the transition from the thick plate to the thin plate solution occurs. In
such cases the isotherms at the bottom of the plate will be strongly elongated in the welding
direction and shifted to positions far behind the heat source. The opposite trend is observed at
small values of vdlla, since a rapid equalisation of the temperature gradients in the throughthickness direction of the plate will result in elliptical isotherms at both plate surfaces, located
in an approximately equal distance from the heat source. In either case the temperature at
which the cross-sectional isotherms approach a semi-circle or become parallel with the XXz)axis can be obtained from Fig. 1.43 by reading-off the intercept between the line for the
dimensionless plate thickness and the respective field boundaries.
Limitations of the maps
Since the maps have been constructed on the basis of the analytical heat flow equations, it is
obvious that they will apply only under conditions for which these equations are valid. The
simplifying assumptions inherent in the models can be summarised as follows:
(a)

The parent material is isotropic and homogeneous at all temperatures,


and no phase changes occur on heating.

(b)

The thermal conductivity, density, and specific heat are constant and
independent of temperature.

(c)

The plate is completely insulated from its surroundings, i.e. there are no
heat losses by convection or radiation from the boundaries.

(d)

The plate is infinite except in the directions specifically noted.

(e)

The electrode is not consumed during welding, and all heat is concentrated in a zero-volume point or line.

(a)

(b)

(C)

C
Fig. 1.41. Computed isothermal contours in different sections for 8 = 0.5;
(a) Front view (\j/ = \|/m),
(b) Side view (\|/ = 0),
(c) Top view ( = 0) and bottom view ( = 8).

(a)

(b)

(C)

\Fig. 1.42.Computed isothermal contours in different sections for 8 = 5;


(a) Front view (\|/ = \|/m);
(b) Side view (\|/ = 0);
(c) Top view ( = 0) and bottom view ( = 8).

ep/n3

Thick plate solution


(3-D heat flow)

Thin plate
solution

1-D heat flow


Thick plate
solution
(2-D heat flow)

5 = vd/2a
Fig. 1.43. Heat flow mechanism map showing calculated field boundaries in transverse direction (i|/
= i|/m) of plate vs Qp/n3 and 8 = vdlla.

(f)

Pseudo-steady state, i.e. the temperature does not vary with time when
observed from a point located in the heat source.

In general, the justification of these assumptions relies on a good correlation between theory
and experiments. However, since the analytical solutions ignore the important role of arc
energy distribution and directed metal currents in the weld pool, predictions of the weld thermal programme should be restricted to positions well outside the fusion zone where such
effects are of less importance (to be discussed below).
Example (1.12)

Consider stringer bead welding (GMAW) on a 12mm thick plate of aluminium (> 99% Al)
under the following conditions:

Based on Fig. 1.43, sketch the peak temperature contours in the transverse section of the
weld at pseudo-steady state.

Solution
If we neglect the latent heat of melting, the parameter n3 at the chosen reference temperature
(Tc = Tm) becomes:

Similarly, when v = 3mm s l and a = 85mm2 s"1 we obtain the following value for the dimensionless plate thickness:

Readings from Fig. 1.43 give:


ep

^, (0C)

Model System

Comments

0.50 1.0

340 660

Medium thick
plate solution

Heat flow in x and y directions,


partial heat flow in z direction

0.17 -> 0.50

130 -> 340

Thin plate solution


(2-D heat
flow)

Heat flow in x and y directions,


negligible heat flow in z direction

< 0.17

< 130

Thin plate solution


(1 -D heat
flow)

Heat flow in y direction, negligible


heat flow in x and z directions

A sketch of the HAZ peak temperature contours in the transverse section of the weld is
shown in Fig. 1.44.
Case Study (Ll)
Consider stringer bead GMA welding on 12.5mm thick plates of low alloy steel and aluminium
(i.e. a Al-Mg-Si alloy), respectively under the conditions E = 1.5 kJ mm"1 and r\ = 0.8. Details
of welding parameters for the four series involved are given in Table 1.4. Computed peak
temperature contours in the transverse section of the welds are given in Figs. 1.45 and 1.46.

Arrows indicate heat flow directions

Fusion
zone

Fig. 1.44. Schematic diagram showing specific peak temperature contours within the HAZ of an aluminium weld at pseudo-steady state (Example 1.12).

WELD A1
y(mm)

(b)

WELD A2

z(mm)

(a)

z(mm)

y(mm)

Fig. 1.45. Computed peak temperature contours in aluminium welding at pseudo-steady state (Case study
1.1); (a) Weld Al, (b) Weld A2. Black regions indicate fusion zone.

Aluminium welding
In general, the maximum width of the isotherms at the upper and lower surface of the plate can
be obtained from Fig. 1.38(a) and (b), although these maps are not suitable for precise readings. A comparison with the computed peak temperature contours in Fig. 1.45(a) and (b)
reveals a strong influence of the welding speed on the shape and position of the cross-sectional
isotherms at a constant gross heat input of 1.5 kJ mm"1. It is evident that the extension of the
fusion zone and the neighbouring isotherms becomes considerably larger when the welding
speed is increased from 2.5 to 5 mm s"1. This effect can be attributed to an associated shift
from elliptical to more elongated isotherms at the plate surfaces (e.g. see Fig. 1.43), which
reduces heat conduction in the welding direction. It follows from Fig. 1.43 that the upper left
corner of the map represents the typical operating region for aluminium welding.
Because of the pertinent differences in the heat flow conditions, the temperature-time pattern will also vary significantly between the respective series as indicated by the maps in Figs.
1.39 and 1.40. Hence, in the case of aluminium welding the usual procedure of reporting arc
power and travel speed in terms of an equivalent heat input per unit length of the bead is highly
questionable, since this parameter does not define the weld thermal programme. In general,
the correct course would be to specify both qo, v and d, and compare the weld thermal history
on the basis of the dimensionless parameters n3 and 8.
Steel welding
If welding is performed on a steel plate of similar thickness, the operating region will be
shifted to the lower right corner of Fig. 1.43. Under such conditions, the isotherms adjacent to
the fusion boundary will be strongly elongated in the x-direction even at very low welding
speeds (see Fig. 1.42). This implies that the thermal programme approaches a state where the
temperature distribution is uniquely defined by the net heat input r\E, corresponding to the

(a)

W E L D S1

z(mm)

y(mm)

(b)

WELD S2

z(mm)

y(mm)

Fig. 7.46.Computed peak temperature contours in steel welding at pseudo-steady state (Case study 1.1);
(a) Weld Sl, (b) Weld S2. Black regions indicate fusion zone.

Table 1.4 Operational conditions assumed in Case study (1.1).


qo

Series

(W)

(mras"1)

(mm)

(kJmrrr 1 )

Al-Mg-Si
alloy

Al
A2

6000
3000

5
2.5

12.5
12.5

Low alloy
steel

Sl
S2

9600
4800

8
4

12.5
12.5

Material

n3

1.5
1.5

0.36
0.09

0.50
0.25

1.5
1.5

32.6
8.2

10
5

limiting case of a fast moving high power source. As a result, the calculated shape and width
of the fusion boundary and neighbouring isotherms are seen to be virtually independent of
choice of qo and v as illustrated in Fig. 1.46(a) and (b).
1.10.4.2 Experimental verification of the medium thick plate solution
It is clear from the above discussion that the medium thick plate solution provides a systematic
basis for calculating the temperature distribution within the HAZ of stringer bead weldments
under various welding conditions. In the following, the accuracy of the model will be checked
against extensive experimental data, as obtained from in situ thermocouple measurements and
numerical analyses of a large number of bead-on-plate welds.
Weld thermal cycles
Examples of measured and predicted weld thermal cycles in aluminium welding are presented
in Fig. 1.47. It is evident that the medium thick plate solution predicts adequately the HAZ
temperature-time pattern under different heat flow conditions for fixed values of the peak
temperature. This, in turn, implies that the model is also capable of predicting the total time
spent in a thermal cycle within a specific temperature interval as shown in Fig. 1.48.
Weld cooling programme
At temperatures representative of the austenite to ferrite transformation in mild and low alloy
steel weldments, the conditions for a fast moving high power source are normally approached
before the transition from thick plate to thin plate welding occurs (see Fig. 1.39). In such
cases, it is possible to present the different solutions for Ax875 (at \|/ = = 0) in a single graph by
introducing the following groups of variables*:

Ordinate:

Abscissa:
Relevant literature data for the cooling time between 800 and 5000C are given in Fig. 1.49.
A closer inspection of the figure reveals a reasonable agreement between observed and predicted values in all cases. For welding of thick plates, the ordinate attains a constant value of
1. Similarly, in thin sheet welding, the slope of the curve becomes equal to 1.
In aluminium welding, the thermal conditions will be much more complex because of the
resulting higher base plate thermal diffusivity (see Fig. 1.39). Hence, it is not possible to
describe the weld cooling programme in terms of equations (1-74) and (1-101), which apply to
fast moving high power sources. The plot in Fig. 1.50 confirms, however, that the medium
thick plate solution is also capable of predicting the cooling time within specific temperature
intervals (e.g. from 300 to 1000C) in aluminium weldments.

These groups of variables can be obtained from equations (1-74) and (1-101).

Measured
Predicted

Temperature (0C)

GMAW: q 0 = 3872 W, v = 8.8 mm/s,


d =8 mm

Time (sec)

A tr(s), observed

Fig. 1.47. Comparison between measured and predicted weld thermal cycles in aluminium welding for
fixed values of Tp. Data from Myhr and Grong.20

P
t

A tr(s), predicted
Fig. /.4#. Comparison between measured and predicted retention times in aluminium welding for fixed
values of Tp. Data from Myhr and Grong.20

8/5
AT

n 3 [-i--4-

SMAW
SMAW
SAW
SAW
THICK PLATE
SOLUTION

THIN PLATE
SOLUTION

-2Hr + Ir 1
5 6SOO 6SOO
Fig. 1.49. Comparison between observed and predicted cooling times from 800 to 5000C in steel welding
(solid lines represent theoretical calculations). Data from Myhr and Grong.20

At

(sec), observed

GMAW (Al+2.5 wt% Mg)

T(0C)

t(sec)

At

(sec), predicted

Fig. 1.50. Comparison between observed and predicted cooling times from 300 to 1000C in aluminium
welding. Data from Myhr and Grong.20

Peak temperatures and isothermal contours


Figure 1.51 shows a comparison between measured and predicted HAZ peak temperatures in
aluminium welding. It is evident that the relative positions of the HAZ isotherms can be
calculated with a reasonable degree of accuracy from the medium thick plate solution, provided that the equation is precalibrated against a known isotherm (i.e. the fusion boundary).
Additional information on the HAZ peak temperature distribution in aluminium welding
can be obtained from the data of Koe and Lee21 reproduced in Fig. 1.52. These numerical
calculations* showed a good correlation with experimental measurements. A comparison with
the medium thick plate solution in Fig. 1.52 reveals a fair agreement between numerically and
analytically computed isothermal contours. It is interesting to note that even though the plate
thickness is small (i.e. 3.2mm), the mode of heat flow becomes essentially three-dimensional
close to the fusion boundary. This important point is often overlooked when discussing the
relevance of the analytical heat flow models in thin sheet welding.
LlOAJ Practical implications
The following important conclusions can be drawn from the results presented in Figs. 1.381.52:

Tp, 0C , observed

GMAW (Al +2.5 wt% Mg)

Tp, 0C , predicted
Fig. 1.51. Comparison between observed and predicted HAZ peak temperatures in aluminium welding.
Data from Myhr and Grong.20
Based on the finite difference method (FDM).

Z (mm)

y (mm)

Fig. 1.52. Comparison between numerically and analytically computed peak temperature contours in
GTA welding of a 3.2mm thin aluminium sheet. (Broken lines: numerical model; solid lines: analytical
model.) Data for welding parameters and material properties are given in Ref. 21.

1. Considering heat flow and temperature distribution in fusion welding, there exists no
defined plate thickness which can be regarded as 'thick' or 'thin'. Accordingly, in a real
welding situation, the mode of heat flow will vary continuously with increasing distance
from the heat source.
2. Close to the centre of the weld, the thermal conditions will be similar to those in a
heavy slab. This means that the temperature distribution is approximately described by the
Rosenthal thick plate solution.
3. At intermediate distances from the heat source, the temperature distribution will deviate significantly from that observed in thick plate welding because of variable temperature gradients in the through-thickness direction of the plate. Within this transition region,
reliable predictions of the pseudo-steady state temperature distribution can only be made
from the medium thick plate solution.
4. For points located sufficiently far away from the heat source, the temperature gradients in the through-thickness direction of the plate become negligible. Under such conditions, the weld thermal programme is approximately defined by the Rosenthal thin plate
solution.
5. In general, a full description of the weld thermal history requires that both the arc
power qo, the travel speed V, and the plate thickness d are explicitly specified. Hence, the
usual procedure of reporting welding variables in terms of an equivalent heat input per unit
length of the bead, (kJ mm"1), is highly questionable, since this parameter does not define
the weld thermal programme. An exception is welding of thick steel plates, where the
temperature distribution approaches that of a fast moving high power source because of a
low thermal diffusivity of the base metal.
6. A comparison between theory and experiments shows that the medium thick plate
solution predicts adequately both the peak temperature distribution and the temperaturetime pattern within the HAZ of stringer bead weldments for a wide range of operational
conditions (including aluminium and steel welding). A requirement is, however, that the
equation is calibrated against a known isotherm (e.g. the fusion boundary) due to the simplifying assumptions inherent in the model.

1.10.5 Distributed heat sources


In some cases it is also necessary to consider the important influence of filler metal additions,
arc energy distribution, and convectional heat flow in the weld pool on the resulting bead
morphology to obtain a good agreement between theory and experiments. Of particular interest in this respect is the formation of the so-called weld crater/weld finger, frequently observed
in GMA and SA stringer bead weldments (see Fig. 1.53). Although the nature of these phenomena is very complex, they can readily be accounted for by applying an empirical correction for the effective heat distribution in the weld pool.
1.10.5.1 General solution
Model (after Myhr and Grong22)
The heat distribution used to simulate the weld crater/weld finger formation is shown
schematically in Fig. 1.54(a). Here we consider two discreate distributions of elementary
point sources, which extend in the y- and z-direction of the plate, respectively. The contribution from each source to the temperature rise in an arbitrary point P located within the plate is
calculated on the basis of the "method of images", as shown in Fig. 1.54(b) and (c). For a heat
source displaced in the y-direction (Fig. 1.54(b)), the temperature field is given by equation (1 104):

(1-106)
where

y
Crater

HAZ"

Finger
Fusion line

Z
Fig. 1.53. Schematic diagram showing the weld crater/weld finger formation during stringer bead welding.

(a)

(b)

Z
Fig. 1.54. General heat flow model for welding on medium thick plates; (a) Physical representation of the
heat distribution by elementary point sources, (b) Method for calculating the temperature field around an
elementary point source displaced along the y-axis.

(C)

Z
Fig. 1.54. General heat flow model for welding on medium thick plates(continued): (c) Method for calculating the temperature field around an elementary (submerged) point source displaced along the z-axis.
Similarly, for a submerged point source located along the z-axis (Fig. 1.54(c))> we obtain:

(1-107)

where
and

Note that equation (1-107) correctly reduces to equation (1-106) when A approaches zero.
The total temperature rise in point P is then obtained by superposition of the temperature
fields from the different elementary heat sources, i.e.:
(1-108)
where

In practice, we can subdivide the heat distributions into a relatively small number of elementary point sources, and usually a total number of 8 to 10 sources is sufficient to obtain
good results (i.e. smooth curves). However, the relative strength of each heat source and their
distribution along the y- and z-axes must be determined individually by trial and error by
comparing the calculated shape of the fusion boundary with the real (measured) one.
Figure 1.55 shows the results from such calculations, carried out for a single pass (bead-ingroove) GMA steel weld. It is evident that the important effect of the weld crater/weld finger
formation on the HAZ peak temperature distribution is adequately accounted for by the present
model. A weakness of the model is, of course, that the shape and location of the fusion boundary must be determined experimentally before a prediction can be made.
1.10.5.2 Simplified solution
Similar to the situation described above, the point heat source will clearly not be a good
model when the heat is supplied over a large area. Welding with a weaving technique and surfacing with strip electrodes are prime examples of this kind.
Model (after Grong and Christensen19)

As a first simplification, the Rosenthal thick plate solution is considered for the limiting case
of a high arc power qo and a high welding speed D, maintaining the ratio qo Iv within a range
applicable to arc welding. Consider next a distributed heat source of net power density qo I2L
extending from -L to +L on either side of the weld centre-line in the y-direction*, as shown
schematically in Fig. 1.56. It follows from equation (1-73) that an infinitesimal source dqy
located between y and y + dy will cause a small rise of temperature in point P at time f, as:
(1-109)
where

and

Alternatively, we can use a Gaussian heat distribution, as shown in Appendix 1.4.

z (mm)

Shaded region indicates


fusion zone

y (mm)
Fig. 1.55. Calculated peak temperature contours in the transverse section of a GMA steel weldment (Operational conditions: / = 450A, U - 30V, v = 2.6mm s"1, d = 50mm).

2-D heat flow

Fig. 1.56. Distributed heat source of net power density qJ2L on a semi-infinite body (2-D heat flow).

(1-110)

where erf(u) is the Gaussian error function (previously defined in Appendix 1.3).
Peak temperature distribution
Because of the complex nature of equation (1-110), the variation of peak temperature with
distance in the y-z plane can only be obtained by numerical or graphical methods. Accordingly, it is convenient to present the different solutions in a dimensionless form. The following
parameters are defined for this purpose:
Dimensionless operating parameter:

(1-111)

Dimensionless time:
(1-112)
Dimensionless y-coordinate:
(1-113)
Dimensionless z-coordinate:
(1-114)
By substituting these parameters into equation (1-110), we obtain:

(1-115)

where 0 is the dimensionless temperature (previously defined in equation (1-9)).

The variation of peak temperature Qp with distance in the y-z plane has been numerically
evaluated from equation (1-115) for chosen values of P and 7 (i.e. P = 0, (3 = 3/4, P = I , and
7 = 0). The results are presented graphically in Figs. 1.57 and 1.58 for the through thickness
(z = zm) and the transverse (y = ym) directions, respectively. These figures provide a systematic
basis for calculating the shape of the weld pool and neighbouring isotherms under various
welding conditions.
In Fig. 1.59 the weld width to depth ratio has been computed and plotted for different
combinations of nw and 9p. It is evident that the predicted width of the isotherms generally is
much greater, and the depth correspondingly smaller than that inferred from the point source
model. Such deviations tend to become less pronounced with decreasing peak temperatures
(i.e. increasing distance from the heat source). At very large value of nw, the theoretical shape
of the isotherms approaches that of a semi-circle, which is characteristic of a point heat source.
Example (1.13)

Consider GMA welding with a weaving technique on a thick plate of low alloy steel under the
following conditions:

Sketch the contours of the fusion boundary and the Ac r isotherm (71O0C) in the y-z plane.
Compare the shape of these isotherms with that obtained from the point heat source model.
Solution
If we neglect the latent heat of melting, the operating parameter at the chosen reference temperature (Tc = Tm) becomes:

Fusion boundary
Here we have:

Readings from Figs. 1.57 and 1.58 give:

Ac j-temperature
In this case the peak temperature should be referred to 7200C, i.e.:

ep/nw

V -

ep/nw

Fig. /.57. Calculated peak temperature distribution in the through-thickness direction of the plate at different positions along the weld surface.

y m /L
Fig. 1.58. Calculated peak temperature distribution in the transverse direction of the plate at position y
(Z) = 0.

y m /z m

nw
Fig. 1.59. Effects of nw and dp on the weld width to depth ratio.

from which

Readings from Figs. 1.57 and 1.58 give:

Similarly, equation (1-75) provides a basis for calculating the width of the isotherms in the
limiting case where all heat is concentrated in a zero-volume point. By rearranging this equation, we obtain:

which gives
6.3 mm when Bp = 1(Tp = 15200C)

and
when
Figure 1.60 shows a graphical presentation of the calculated peak temperature contours.
Implications of model

It is evident from Fig. 1.60 that the predicted shape of the isotherms, as evaluated from equation (1-110), departs quite strongly from the semi-circular contours required by a point heat
source. Moreover, a closer inspection of the figure shows that inclusion of the heat distribution
also gives rise to systematic variations in the weld thermal programme along a specific
isotherm, as evidenced by the steeper temperature gradient in the v-direction compared with
the z-direction of the plate. This point is more clearly illustrated in Fig. 1.61, which compares
the HAZ temperature-time programme for the two extreme cases of z = 0 and y = 0, respectively. It is obvious from Fig. 1.61 that the retention time within the austenite regime is
considerably longer in the latter case, although the cooling time from 800 to 5000C, Af8/5, is
reasonably similar. These results clearly underline the important difference between a point
heat source and a distributed heat source as far as the weld thermal programme is concerned.
Model limitations

In the present model, we have used the simplified solution for a fast moving high power
source (equation 1-73)) as a starting point for predicting the temperature-time pattern. Since
the equations derived later are obtained by integrating equation (1-73), they will, of course,
apply only under conditions for which this solution is valid.
Moreover, a salient assumption in the model is that the heat distribution during weaving
can be represented by a linear heat source orientated perpendicular to the welding direction.
Although this is a rather crude approximation, experience shows that the assumed heat dis-

z, mm
Fig. /.60. Predicted shape of fusion boundary and Acpisotherm during GMA welding of steel with an
oscillating electrode (Example 1.13). Solid lines: Distributed heat source; Broken lines: Point heat source.

Temperature, 0C

Time, s
Fig. L61. Calculated HAZ thermal cycles in positions y = 0 and z = 0 (Example 1.13).
tribution is not critical unless the rate of weaving is kept close to the travel speed. However,
for most practical applications weaving at such low rates would be undesirable owing to an
unfavourable bead morphology.
Case Study (1.2)

Surfacing with strip electrodes makes a good case for application of equation (1-110). Specifically, we shall consider SA welding of low alloy steel with 60mm X 0.5mm stainless steel
electrodes. The operational conditions employed are listed in Table 1.5.
It is evident from the metallographic data presented in Fig. 1.62 that neither the bead penetration nor the HAZ depth (referred to the plate surface) can be predicted readily on the basis
of the present heat flow model when welding is carried out with a consumable electrode,
owing to the formation of a reinforcement. This situation arises from the simplifications made
in deriving equation (1-110). The problem, however, may be eliminated by calculating the
depth of the Ac3 and Ac1 regions relative to the fusion boundary, i.e. Azm = zm(Qp) - zm (Qp = 1),
or Aym = ym (8p) - ym (6p = 1), for specific positions along the weld fusion line, as shown by the
solid curves in Fig. 1.62 for Qp = 0.54 and 0.45, respectively. An inspection of the graphs
reveals satisfactory agreement between theory and experiments in all three cases, which implies that the model is quite adequate for predicting the HAZ thermal programme as far as strip
electrode welding is concerned. This result is to be expected, since the assumption of twodimensional heat flow is a realistic one under the prevailing circumstances.
Case Study (1.3)

As a second example we shall consider GTA welding (without filler wire additions) at various
heat inputs and amplitudes of weaving within the range from 1 to 2.5 kJ mm"1 and 0 to 15mm,
respectively. Data for welding parameters are given in Table 1.6.

Stainless steel!
y, mm

Weld S3

Fusion line

z, mm

Stainless steel
y, mm

Weld S4

z, mm
Stainless steel!
y, mm
Weld S5

z, mm
Fig. 1.62. Comparison between observed and predicted Ac3 and Ac1 contours during strip electrode welding (Case study 1.2). Data from Grong and Christensen.19
Table 1.5 Operational conditions used in strip electrode welding experiments (Case study 1.2).
Base metal/
filler metal
combination

Weld
No.

S3
Low alloy steel/
stainless steel

/
(A)

730

S4

73

S5

730

U
(V)

27
27

27

v
(mms"1)

2L
(mm)

nw
Cn = 0.7)

1.8

60

0.34

2 2

2.5

60
60

0.28
0.24

Table 1,6 Operational conditions used in GTA welding experiments (Case study 1.3).
I

2L
1

nw
"V=O-

12

WeIdNo.

(A)

(V)

(mms" )

(mm)

T] = 0.23

Bl

200

13.5

2.6

9.5

0.20

0.40

B2

200

14.0

1.1

15.0

0.20

0.40

B3

200

13.5

2.5

B4

200

12.5

1.0

Calibration procedure
In general, a comparison between theory and experiments requires that the arc efficiency factor can be established with a reasonable degree of accuracy. Unfortunately, the arc efficiency
factor for GTA welding has not yet been firmly settled, where values from 0.25 up to 0.75 have
been reported in the literature (see Table 1.3). Additional problems result from the fact that
only a certain fraction of the total amount of heat transferred from the arc to the base plate is
sufficiently intense to cause melting. This has led to the introduction of the melting efficiency
factor T]m, which normally is found to be 30-70% lower than the total arc efficiency of the
process, depending on the latent heat of melting, the applied amperage, voltage, shielding gas
composition, or electrode vertex angle.23 Consequently, since these parameters cannot readily
be obtained from the literature, the following reasonable values for r\m and r\ have been assumed to calculate nw in Table 1.6, based on a pre-evaluation of the experimental data: j \ m =
0.12 (fusion zone), TI = 0.23 (HAZ).
It should be noted that the above values also include a correction for three-dimensional heat
flow, since the assumption of a fast moving high power source during low heat input GTA
welding is not valid. Hence, both the arc efficiency factor and the melting efficiency factor
used in the present case study are seen to be lower than those commonly employed in the
literature.
Full weaving (welds Bl and B2)
The results from the metallographic examination of the two GTA welds deposited under full
weaving conditions are presented graphically in Fig. 1.63. Note that the shape of the fusion
boundary as well as the Ac3 and the Ac1 isotherms can be predicted adequately from the present
model for both combinations of E and L (an exception is the HAZ end points in position z = 0),
provided that proper adjustments of i\m and Tj are made. The good correlation obtained in Fig.
1.63 between the observed and the calculated peak temperature contours justifies the adaptation of the model to low heat input processes such as GTA welding, despite the fact that the
assumption of two-dimensional heat flow is not valid under the prevailing circumstances.
No weaving (welds B3 and B4)
For the limiting case of no weaving (Fig. 1.64), the concept of an equivalent amplitude of
weaving has been used in order to calculate the peak temperature contours from the model.
This parameter (designated Leq) takes into account the effects of convectional heat flow in the
weld pool on the resulting bead geometry, and is evaluated empirically from measurements of
the actual weld samples. At low heat inputs (Fig. 1.64(a)), the agreement between theory and

y, mm

Weld B1

z, mm
Fusion line

y, mm

Weld B2

z, mm

Fig. 1.63. Comparison between observed and predicted fusion line, Ac3 and Ac1 contours during GTA
welding under full weaving conditions (Case study 1.3). Data from Grong and Christensen.19
experiments is largely improved by inserting 2Leq = 7.5mm into equation (1-110), when comparison is made on the basis of the point source model. In contrast, at a heat input of 2.5 kJ
mm"1 (Fig. 1.64(b)), the measured shape of the HAZ isotherms is seen to approach that of a
semi-circle, and hence the deviation between the present model and the simplified solution for
a fast moving high power point source is less apparent.
Intermediate weaving
At intermediate amplitudes of weaving (2L = 5 and 7.5mm, respectively), convectional heat
flow in the weld pool will also tend to increase the bead width to depth ratio beyond the
theoretical value predicted from the present model, as shown in Fig. 1.65. The plot in Fig. 1.65

Weld B3
y, mm

z, mm

Fusion line

y, mm

Weld B4

z, mm
Fig. 1.64. Comparison between observed and predicted fusion line, Ac3 and Ac1 contours during GTA
welding with a stationary arc (Case study 1.3). Solid lines: Distributed heat source, Broken lines: Point
heat source. Data from Grong and Christensen.19
includes all data obtained in the GTA welding experiments with an oscillating arc, as reported
by Grong and Christensen.19 These results suggest that the applied amplitude of weaving must
be quite large before such effects become negligible. Consequently, adaptation of the model to
the weld series considered above would require an empirical calibration of the weaving amplitude similar to that performed in Fig. 1.64 for stringer bead weldments to ensure satisfactory
agreement between theory and experiments.
1.10.6 Thermal conditions during interrupted welding
Rapid variations of temperatures as a result of interruption of the welding operation can have
an adversely effect on the microstructure and consequently the mechanical properties of the
weldment.

Width

Width to depth ratio

Depth

Theoretical
curve

Fig. 1.65.Comparison between observed and predicted weld width to depth ratios during GTA welding
with an oscillating arc (Case study 1.3). Data from Grong and Christensen.19

T,e

t,t

Fig. 1.66. Idealised heat flow model for prediction of transient temperatures during interrupted welding.

Model (after Rykalin9)

The situation existing after arc extinction may be described as shown in Fig. 1.66. From time
t = t* there is no net heat supply to the weldment. This condition is satisfied if the real source
q0 is considered maintained by adding an imaginary source +qo and sink -qo of the same
strength at t*. The temperature at some later time t** in a given position R0 (measured from the
origin 0") is then equal to the difference of temperatures due to the positive heat sources qo
and +qo and the negative heat sink -qo. Each of these temperature contributions will be a
product of a pseudo-steady state temperature Tps, and a correction factor K1 or K2 (given by
equations (1-49) and (1-82), respectively). Hence, for 3-D heat flow, we have:
(1-116)
where
and
Similarly, for 2-D heat flow, we get:
(1-117)
where

(ro is the position of the weld with respect to the imaginary heat source at time f* in the x y
plane).
Example (1.14)

Consider repair welding of a heavy steel casting with covered electrodes under the following
conditions:

Suppose that a 50mm long bead is deposited on the top of the casting. Calculate the temperature in the centre of the weld 5 s after arc extinction.
Solution

The pseudo-steady state temperature for points located on the weld centre-line (i|/ = = 0) can
be obtained from equation (1-65). When t** - t* = 5 s, we get:

Referring to Fig. 1.67, the position of the weld with respect to the imaginary heat source at
time f* is 10mm, which gives:

Fig. 1.67. Sketch of weld bead in Example 1.14.

(a)
3-D heat flow

(b)
3-D heat flow

(C)
3-D heat flow

Fig. /.6#. Recommended correction factor/for some joint configurations; (a) Single V-groove, (b) Double V-groove, (c) T-joint.

Moreover, the dimensionless times T** and T** - x* are:

and
At these coordinates, the correction factor K1 is seen to be 1 and 0.62, respectively (Fig.
1.18). The temperature in the centre of the weld 5 s after arc extinction is thus:
which is equivalent to

LlOJ

Thermal conditions during root pass welding

During conventional bead-on-plate welding the angle of heat conduction is equal to 180 due
to symmetry effects (e.g. see Fig. 1.23). In order to apply the same heat flow equations during
root pass welding, it is necessary to introduce a correction factor,/, which takes into account
variations in the effective heat diffusion area due to differences in the joint geometry. Taking
/equal to 1 for ordinary bead-on-plate welding (b.o.p.), we can define the net heat input of a
groove weld as:9
(1-118)
Recommended values of the correction factor/for some joint configurations are given in
Fig. 1.68.
Example (Ll5)

Consider deposition of a root pass steel weld in a double-V-groove with covered electrodes
(SMAW) under the following conditions:

Calculate the cooling time from 800 to 5000C (Ar875), and the cooling rate (CR.) at 6500C
in the centre of the weld when the groove angle is 60.
Solution

The cooling time, Ar875, and the cooling rate, CR., can be obtained from equation (1-68) and
(1-71), respectively:
Cooling time, Atm

Cooling rate at 6500C

The above calculations show that the thermal conditions existing in root pass welding may
deviate significantly from those prevailing during ordinary stringer bead deposition due to
differences in the effective heat diffusion area. These results are in agreement with general
experience (see Fig. 1.69).
1.10.8 Semi-empirical methods for assessment of bead morphology
In fusion welding fluid flow phenomena will have a strong effect on the shape of the weld
pool. Since flow in the weld pool is generally driven by a combination of buoyancy, electromagnetic, and surface tension forces (e.g. see Fig. 3.10 in Chapter 3), prediction of bead morphology from first principles would require detailed consideration of the current and heat flux
distribution in the arc, the interaction of the arc with the weld pool free surface, convective
heat transfer due to fluid flow in the liquid pool, heat of fusion, convective and radiative losses
from the surface, as well as heat and mass loss due to evaporation.
Over the years, a number of successful studies have been directed towards numerical weld
pool modelling, based on the finite difference, the finite element, or the control volume approach.24"31 Although these studies provide valuable insight into the mechanisms of weld pool
development, the solutions are far too complex to give a good overall indication of the heatand fluid-flow pattern. The present treatment is therefore confined to a discussion of factors
affecting the nominal composition of single-bead fusion welds. This composition can be obtained from an analysis of the amount of deposit D and the fused part of the base material B,
from which we can calculate the mixing ratio BI(B + D) or DI(B + D). Methods have been
outlined in the preceding sections for handling such problems by means of point or line source
models. The following section gives a brief description of procedures which can be used for
predictions of the desired quantities in cases where the classic models break down, or where
the calculation will be too tedious.
1.10.8.1 Amounts of deposit and fused parent metal
The heat conduction theory does not allow for the presence of deposited metal. The rate of
deposition, dMwldt, is roughly proportional to the welding current /, and is often reported as a
coefficient of deposition, defined as:
(1-119)
Since the area of deposited metal D is frequently wanted, we may write:
(1-120)
where p is the density, and v is the welding speed.

Groove angle: <>


| = 60

Cooling time, AtQ/5(s)

Plate thickness:

Heat input, E (kJ/mm)


Fig. 1.69. Comparison between observed and predicted cooling times from 800 to 5000C in root pass
welding of steel plates (groove preparation as in Fig. 1.68(b)). Data from Akselsen and Sagmo.34
Recommended values of k'/p for some arc welding processes are given in Table 1.7. In
practice, the deposition coefficient k'/p will also vary with current density and electrode stickout
due to resistance heating of the electrode. Consequently, the numbers contained in Table 1.7
are estimated averages, and should therefore be used with care.
Example (1.16)

Consider stringer bead deposition (S AW) on a thick plate of low alloy steel under the following conditions:

Table 1.7 Average rates of volume deposition in arc welding. Data from Christensen.32
Welding Process

k'/p (mm3 A"1 s l)

SMAW

0.3-0.5

GMAW, steel

0.6-0.7

GMAW, aluminium

-0.9

SAW, steel

-0.7

Calculate the mixing ratio Bf(B + D) at pseudo-steady state.


Solution

The amount of fused parent material can be obtained from equation (1-75). If we include an
empirical correction for the latent heat of melting, the dimensionless radius vector a4m becomes:

This gives:

Similarly, the amount of deposited metal can be calculated from equation (1-120). Taking
ifc'/p equal to 0.7mm3 A"1 s"1 for SAW (Table 1.7), we get:

The mixing ratio is thus:

Example (L 17)

Consider stringer bead deposition with covered electrodes (SMAW) on a thick plate of low
alloy steel under the following conditions:

Calculate the mixing ratio BI(B + D) at pseudo-steady state.


Solution

In this particular case the conditions for a fast moving high power source are not met. Thus, in
order to eliminate the risk of systematic errors, the amount of fused parent metal should be
calculated from the general Rosenthal thick plate solution (equation (1-45)) or read from Fig.
1.21. When Tc = Tm (i.e. 8^ = 1), we obtain:

Reading from Fig. 1.21 gives:

and

Moreover, the amount of deposited metal can be calculated from equation (1-120). Taking
k'/p equal to 0.4mm3 A"1 s~l for SMAW (Table 1.7), we get:

and

The above calculations indicate a small difference in the mixing ratio between SA and
SMA welding, but the data are not conclusive. In practice, a value of BI(B + D) between 1/3
and 1/2 is frequently observed for SMAW, while the mixing ratio for SAW is typically 2/3 or
higher. The observed discrepancy between theory and experiments arises probably from difficulties in estimating the amount of fused parent metal from the point heat source model.
1.10.8.2 Bead penetration
It is a general experience in arc welding that the shape of the fusion boundary will depart quite
strongly from that of a semi-circle due to the existence of high-velocity fluid flow fields in the
weld pool.24"31 For combinations of operational parameters within the normal range of arc
welding, a fair prediction of bead penetration h can be made from the empirical equation
derived by Jackson et al.:33
(1-121)
A summary of Jackson's data is shown in Table 1.8. It is seen that the constant C in equation (1-121) has a value close to 0.024 for SAW and SMAW with E6015 type electrodes, and
about 0.050 for GMAW with CO2 -shielding gas. Penetration measurements of GMA/Ar + O2,
GMA/Ar, and GMA/He welds, on the other hand, show a strong dependence of polarity, and
shielding gas composition, to an extent which makes the equation useless for a general prediction. Such data have therefore not been included in Table 1.8.
Example (1.18)

Based on the Jackson equation (equation (1-121)), calculate the bead penetration for the two
specific welds considered in Examples (1.16) and (1.17). Use these results to evaluate the
applicability of the point heat source model under the prevailing circumstances.
Solution

From equation (1-121) and Table 1.8, we have:

Table 1.8 Recommended bead penetration coefficients for some arc welding processes. Data from
Jackson.33
Welding Process

Comments

SAW, steel

-0.024

Various types of fluxes


(Zz from 3 to 15 mm)

SMAW, steel
(E6015)

-0.024

Wide range of /, U, and v


(h from 0.7 to 5mm)

GMAW, steel
(CO2 - shielding)

-0.050

Electrode positive
(h from 6.5 to 8mm)

and

The corresponding values predicted from the point heat source model are:

and

Provided that the Jackson equation gives the correct numbers, it is obvious from the above
calculations that the point heat source model is not suitable for reliable predictions of the bead
penetration during arc welding. This observation is not surprising.
1.10.9 Local preheating
So far, we have assumed that the ambient temperature T0 remains constant during the welding
operation (i.e. is independent of time). The use of a constant value of T0 is a reasonable approximation if the work-piece as a whole is subjected to preheating. In many cases, however,
the dimensions of the weldment allow only preheating of a narrow zone close to the weld.
This, in turn, will have a significant influence on the predicted weld cooling programme, particularly in the low temperature regime where the classic models eventually break down when
T approaches T0.
Model (after Christensen n)

The idealised preheating model is shown in Fig. 1.70. Here it is assumed that the weld centreline temperature is equal to the sum of the contributions from the arc and from the field of
preheating. The former contribution is given by equation (1-45) for R = -x = Vt, provided that
the plate thickness is sufficiently large to maintain 3-D heat flow. Similarly, the temperature
field due to preheating can be calculated as shown in Section 1.7 for uniaxial heat conduction
from extended sources (thermit welding). By combining equations (1-45) and (1-22), we
obtain the following relation for the weld centre-line:
(1-122)

Temperature profile
att = 0

Weld
Preheated
zone

z
Fig. 1.70. Sketch of preheating model.
where T* is the local preheating temperature, and L* is the half width of the preheated zone.
Equation (1-122) can be written in a general form by introducing the following groups of parameters:
Dimensionless temperature:
(1-123)

Time constant:
(1-124)

Dimensionless time:
(1-125)

Dimensionless half width of preheated zone:


(1-126)
By inserting these parameters into equation (1-122), we obtain:
(1-127)

It is evident from the graphical representation of equation (1-127) in Fig. 1.71 that the
predicted weld cooling programme falls within the limits calculated for Q"-^ 0 (no preheating)
and Q /7 -> oo (global preheating). The controlling parameter is seen to be the dimensionless
half width of the preheated zone Q", which depends both on the actual width L*, the base plate
thermal properties a, X, and the net heat input qo Iv.
Example (1.19)

Consider stringer bead deposition with covered electrodes (SMAW) on a thick plate of low
alloy steel under the following conditions:

Calculate the cooling time from 800 to 50O0C (A%5), and the cooling time 1Oo measured
from the moment of arc passage to the temperature in the centre of the weld reaches 10O0C.
Solution
First we calculate the time constant to from equation (1-124):

e*-

from which we obtain

^6

Fig. 1.71. Graphical representation of equation (1-127).

Next Page
Cooling time, At8/5

The dimensionless temperatures conforming to 800 and 5000C are:

Reading from Fig. 1.71 gives:

from which

This cooling time is only slightly longer than that calculated from equation (1-68) for T0 =
200C (6.9s), showing that moderate preheating up to 1000C is not an effective method of
controlling Ar875.
Cooling time, t]00

When T=T0* = 1000C, the dimensionless temperature 9* = 1. Reading from Fig. 1.71 gives
T 6 - 10, from which:

The above value should be compared with that evaluated from the numerical data of Yurioka
et al.,35 replotted in Fig. 1.72 (see p.104). It follows from Fig. 1.72 that the weld cooling
programme in practice is also a function of the plate thickness d, an effect which cannot readily be accounted for in a simple analytical treatment of the heat diffusion process. For the
specific case considered above the parameter ^100 varies typically from 500 to 900s, depending
on the chosen value of d. This cooling time is significantly shorter than that calculated from
equation (1-127), indicating that the analytical model is only suitable for qualitative predictions.

References
1.
2.
3.
4.
5.
6.

H.S. Carslaw and J.C. Jaeger: Conduction of Heat in Solids; 1959, Oxford, Oxford University
Press.
British Iron and Steels Research Association: Physical Constants of some Commercial Steels
at Selected Temperatures; 1953, London, Butterworths.
R. Hultgren, R.L. Orr, RD. Anderson and K.K. Kelly: Selected Values of Thermodynamic
Properties of Metals and Alloys; 1963, New York, J. Wiley & Sons.
E. Griffiths (ed.): J. Iron and Steel Inst., 1946,154, 83-121.
J.E. Hatch (ed.): Aluminium Properties and Physical Metallurgy; 1984, Metals Park (Ohio),
American Society for Metals.
Metals Handbook, 9th edn., Vol. 2, 1979, Metals Park (Ohio), American Society for Metals.

Вам также может понравиться