You are on page 1of 203

In-Service

Welding of
Gas Pipelines

CRC WS Pipeline Program

Project Report
March 2000
In-Service
Welding of Gas
Pipelines
CRCWS Project 96:34 Final Report

Final Project Report


June 2000

M.J.Painter, CSIRO Manufacturing Science and Technology


P. Sabapathy, The University of Adelaide

CRCWS
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report Contents page

In-Service Welding Of Gas Pipelines


Pipeline Program Report

September 1996- September 1999

Executive Summary

1. Introduction

2. Research Strategy & Outline

3. Literature Review: Experimental Studies of In-Service Welding

4. Literature Review: Simulation of Fusion Welding

5. Appraisal of Battelle In-Service Welding Software

6. Welding on In-Service Pipelines: Industry Survey

7. An Experimental Appraisal of Hydrogen Controlled Electrodes

8. Numerical Modelling of In-Service Welding

9. Burn-through Prediction

10. Model Validation Using the Gladstone Flow Loop

11. A Proposed Method of Presenting Results for Industry Use

Mike Painter, CSIRO Manufacturing Science and Technology


Prakash Sabapathy, The University of Adelaide
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report Contents page

Acknowledgement
This is a collaborative project of the Cooperative Research Centre for Welded Structures, under the
sponsorship and guidance of the Pipeline Program Management Committee. Therefore it would not
have progressed without the guidance, endeavor and financial assistance of the following
organisations and people:

Collaborative Partners
CSIRO Manufacturing Science & Technology
The University of Adelaide
BHP
Epic Energy
AGL Gas
The Pipeline Program Management Committee
Researchers
M.A.Wahab, The University of Adelaide
Bing Feng, BHP
Prakash Sabapathy,The University of Adelaide
Industrial mentors
Paul Grace, WTIA
Hans Borek, Epic Energy

To which, I express thanks and gratitude.


CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 1

In-Service Welding of
Gas Pipelines
September 1996 – September 1999

A collaborative project of the Cooperative Research Centre for Materials Welding & Joining, under
the sponsorship and guidance of the Pipeline Program Management Committee.

Collaborative Partners
CSIRO Manufacturing Science & Technology
The University of Adelaide
BHP
Epic Energy
AGL Gas
The Pipeline Program Management Committee
Researchers
Mike Painter, CSIRO Manufacturing Science & Technology
M. A. Wahab, The University of Adelaide
Bing Feng, BHP
Prakash Sabapathy, The University of Adelaide
Industrial mentors
Paul Grace, WTIA / ex AGL Gas Sydney
Hans Borek, Epic Energy

Project Objective
To develop recommended weld procedures for the safe and effective in-service welding of
thin-wall, high-strength steel, high-pressure gas pipelines.

Executive Summary

Background
The process of welding onto a ‘live’ high-pressure pipe is frequently employed for the repair,
modification or extension of gas pipelines. This ‘in-service welding’ has significant economic
advantages for the gas transmission and gas distribution industries, since it avoids the costs of
disrupting pipeline operation, and it maintains continuity of supply to customers. In-service welding
is an essential part of ‘hot-tapping’, a technique which allows the creation of a branch connection
to a ‘live’ pipeline. In-service welding is also important for pipeline maintenance, such as the
installation of sleeves around damaged sections of pipe. Direct deposition of welds onto ‘live’
pipes has been suggested as a way of replacing wall thickness lost through corrosion or local
damage. If in-service welding is not possible then sections of a pipeline have to be sealed and de-
gassed prior to welding, and then re-purged prior to reinstatement. These are costly, wasteful, and
environmentally damaging actions, since there are large gas losses and methane is a green-house
gas. TransCanada Pipelines Ltd. has estimated that relative to a cold connection a hot-tap avoids
gross revenue losses of approximately 1M$Canadian per hot- tap.

Welding Onto a Gas Pipeline


Two factors make it difficult to weld onto a ‘live’ pipeline. Firstly, the flowing gas creates a large
heat loss through the wall of the pipe, resulting in accelerated cooling of the weld. High carbon
equivalent steels are sensitive to such rapid cooling rates, which increase hardness, and increase
the possibility of heat affected zone (HAZ) cracking. The second factor concerns the local
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 2

heating, and the reduction in pipe-wall-strength during the welding process. If this reduction in
strength is too great the pipe wall can burst under the pipe’s internal pressure. This hazardous
event is termed ‘burn-through’. Increasing the welding heat input can reduce fast cooling, but this
promotes weld penetration and increases the risk of burn-through. Suitable weld procedures must
ensure the HAZ hardness is not high enough to cause cracking, whilst heat input and penetration
are not so high that the integrity of the pipe wall is jeopardised.

In Australia, there is a significant trend towards the use of high yield strength steels for pipeline
construction. Future pipelines using X70 and X80 steels could have wall thickness as low as
3mm. Unfortunately, in-service welding is made much more difficult with such thin-walled pipes.
Thin pipe walls increase the risk of burn-through during welding, and are more easily cooled by the
flowing gas. High strength steels can also be susceptible to the generation of excessive hardness
for a given cooling rate. If the economic advantages of in-service welding are to be maintained
then technology to support the safe and effective welding of thin-walled high-strength pipelines
must be established.

In-Service Welding: Current Technology


Much of the technology associated with in-service welding was generated in the USA. This has
consisted of two approaches:

1. An experimental method of measuring the cooling capacity of the pipeline.


This data is used to establish the required heat input and weld cooling rates from laboratory
pipe welds under the same simulated cooling conditions. This method, generally known as the
EWI Test, was developed by Edison Welding Institute (EWI).

2. Numerical simulation of in-service welding.


From 1980 to 1990 researchers at Battelle Memorial Institute developed a 2D finite difference
approach to simulate sleeve and direct-branch in-service welds. For a given pipe geometry
and a set of welding parameters the weld cooling times, t8/5 (the time to cool from 800 to
500°C) and the maximum inside wall temperature could be calculated. HAZ hardness was
estimated from this predicted cooling rate and the carbon equivalent of the steel. Hardness
below 350 HVN was considered to have a low potential for cracking. Pipes were considered
safe from burn-through if the maximum inside wall temperature was below 980°C. These
limits were determined from a comparison of experimental in collaboration with EWI. These
tests were mainly on thick walled (≈6.35 mm), lower strength (X52) materials. Recommended
limits for the valid use of the Battelle approach are pipe wall thickness within the range 3.2-9.6
mm, for steel grades, up-to and including X52.

In-Service Welding: The Current Project


A survey of the Australian pipeline industry established that there was a substantial conservatism
in the in-service welding procedures currently used. Procedures generally involved a reduction of
pressure and gas flow before welding. The potential value to the pipeline industry of a research
program, which would allow the safe relaxation of such constraints, was estimated to be $2-4 M
over the period 1998-2002.

In recognition of:
• the differences between Australian conditions and those previously researched, namely,
reduced pipe wall thickness and use of higher strength steels,
• the lack of quantified information on burn-through limits for thin walled pipes,
• the importance of being able to carry out in-service welding on new , thin-walled pipelines,
• the indicated economic benefits that improved in-service welding would achieve,

the CRCMWJ Pipeline Program initiated a research project with the following aim:

Project Objective
To develop recommended weld procedures for the safe and effective in-service welding of thin-
walled, high-strength steel, high-pressure gas pipelines.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 3

Project Structure
The important features of the project were:

1. An experimental study of circumferential manual metal arc (MMA) welding in the vertical-up
and vertical-down position, with low hydrogen electrodes (E8018G and E7016, 2.5 mm & 3.2
mm diameter). These tests used a water-cooled simulation of in-service welding and a range
of typical Australian pipe grades (X42-X80).

2. The extensive development of numerical finite element simulations of in-service welding.


These numerical simulations covered 2D and 3D models, and unlike previous research work
aimed to develop numerical simulations of burn-through.

3. The transfer of technology to the pipeline industry.

Brief Project Outcomes


Appraisal of Current Knowledge
It has been suggested that the Battelle program over-estimates the t8/5 cooling times for thin-walled
pipe and under-estimates t8/5 for thick walled pipe. This leads to the possibility of choosing non-
conservative heat input values for thin pipe (≤4.8 mm). The producers of the Battelle software
considered it should be restricted to a wall thickness range of 3.2-12 mm. The calculation of
hardness relies on a relatively simple estimate of a carbon equivalent, which does not necessarily
apply to modern high-strength steel compositions, and should be restricted to grades up to X52.

It is generally accepted that in-service welds on pipes with wall thickness greater than 6.35 mm
have no significant risk of burn-through with good welding practice. The literature review confirmed
the sporadic nature of information relating to burn-through limits on pipes less than 6.35mm thick.

Recommended weld procedures have the following characteristics:

• Severe restriction on welding conditions for pipes less than 5mm wall thickness. Some
recommendations that in-service welding should be restricted to pipes thicker than 4 mm.

• Use of low hydrogen electrodes of small diameter 2-2.4 mm (restricting arc current), in the
vertical-down welding position.

• Limitations on heat input, typically ≤ 0.5 kJ/mm for 3.2 mm wall thickness, ≤ 1.4kJ/mm for
5mm wall thickness.

• Limits on pressure, typically < 6 MPa for thin pipes.

• Careful control of arc current, heat input and welding practice.

Experimental studies have generally used water flow to simulate the cooling effect of the gas, but
there is a general recognition that water gives much higher quench rates than gas flow. It therefore
generates conservative welding conditions for a required hardness. That is heat input determined
with water-flow will give slower t8/5 with gas flow. The corollary of this however, is that using water-
flow simulations may give non-conservative heat inputs for burn-through.

Outcomes from the Experimental Study of Hydrogen Controlled Electrodes.


The analysis of welding conditions and welds produced using pipe cooled with a water jacket to
generate a rapid quench, gave the following results;

• It was possible to generate a weld on 7.8 mm thick X80 pipe with a maximum hardness of 325
HVN with a weld t8/5 cooling time of 3.8 seconds.

• For the range of pipe grades examined (X42-X80), the Yurioka-1 carbon equivalent (CE)
relationship provided the best correlation between composition, hardness and t8/5 cooling time.
This correlation gave an absolute error of 5.7% for E5548-G (E8018-G) electrodes, and a
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 4

3.4% error with E7016. Based on this relationship the X80 grade of pipe used in this work
would only require a t8/5 of 0.8 seconds in order to achieve a hardness of >350 HVN.

• A small beneficial tempering effect was measured from multi-pass [3 passes] welds. This was
most noticeable for the X70 grade steel, which with a CE of 0.288 gave a hardness of 381
HVN after a single root pass. After three passes, the maximum hardness was reduced to an
acceptable 321 HVN.

• Only the two X60 steels with CE of 0.38 and 0.41 gave multi-pass hardness >350 HVN.

• A minor difference in the incidence of weld defects was observed between the two electrode
types tested. There was a greater incidence of HAZ cracking with E8018-G in keeping with its
reduced heat input. Some HAZ cracks were detected with X80 grade although the maximum
hardness was 299-310 HVN.

• There was no systematic variation in heat input as the weld progressed around the pipe.
However, whilst total weld energy remained reasonably constant the natural variation in
welding speed, since this is a manually skilled process, caused variation in the heat input. This
was significant and amounted to approximately a ±20% variation on a nominal value.

• Penetration into the run-pipe was generally greater when using the E7016 electrode in the
vertical-up position rather than with E8018-G in the vertical-down.

• Penetration into the run-pipe slightly increased with increasing heat input although this effect
was largely swamped by a significant variability at a given heat input. At a nominal heat input
of about 1 kJ/mm this variability was approximately, 0.2-0.8 mm with E8018-G and 0.5-1.0
mm with E7016.

These experiments also provided,

• empirical data to relate the deposited weld bead volume, and weld bead shape to heat input
for both E8018-G and E7016 electrodes, and

• preliminary data on measured t8/5 cooling times and measured weld penetrations in order to
validate and refine numerical models of MMA in-service welds.

Development of Numerical Simulations


Utilising commercial finite element software, a wide range of 2D and 3D models of in-service welds
has been developed. The capacity to generate stable, accurate models of all in-service joint forms
has been demonstrated.

To represent the heat loss due to the flowing gas the numerical models follow the Battelle model
and utilise the Sieder & Tate non-dimensional approach. This determines an effective heat transfer
coefficient at the pipe wall, based on the pipe diameter, gas pressure, and flow speed. By
incorporating empirical data describing weld bead volume and joint form the representation of the
welding heat input has been tailored for both vertical-up and vertical-down welding with low-
hydrogen electrodes. For circumferential fillet welds, predictions of both t8/5 cooling times and weld
penetrations have been made with acceptable accuracy.

Model development has concentrated on a 3D quasi-steady-state analysis of circumferential fillet


welds. In practice, this is the most popular joint form. It also provides a numerical analysis which
allows an acceptable, short CPU time. Pre-processing software has been produced to efficiently
construct and mesh models of varied geometry. Post-processing software has been established to
calculate the distribution of t8/5 cooling times throughout the weld zone. This can be further used to
calculate a distribution of hardness based on an empirical relationship with composition and t8/5.
Isotherms give an estimate of weld penetration and size and extent of the HAZ.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 5

Models have mainly considered single root pass welds but some multi-pass welds have been
simulated.

Various methods of assessing the risk of burn-through have been developed. These were based
on:
• the maximum temperature at the inside surface of the pipe, following Battelle,
• a thermo-elastic plastic stress analysis,
• using the thermal field in the pipe wall to calculate the reduction in wall strength.

Thermo-elastic-plastic models of circumferential welds have showed similar deformation patterns


to those observed during burn-through, namely a localised bulge in the pipe wall near the weld.
Failure can be specified as a bulge that exceeds a limiting height. Plots of internal pressure
versus bulge height have shown an effective yield pressure, which can also be used as a failure
index. Only a limited numbers of these models were studied because they were very
computationally demanding.

Estimating the reduction in pipe wall strength in the weld zone has created a novel alternative
method of assessing burn-through risk. This method determines the reduction in material strength
around the weld based on the predicted temperature field, and the known relationship between
material yield strength and temperature. This reduced strength is regarded as equivalent to a local
reduction in the thickness of the pipe-wall at constant ambient temperature. Hence, the
temperature field around the in-service weld effectively converts to a ‘cavity’ in the pipe wall. A
number of alternate strategies can be considered. The limiting pressure for safe welding can be
based on the remaining wall thickness, or based on the effective reduction in cross-sectional area.
The risk of burn-through is equivalent to the possibility of this ‘cavity’ causing rupture at the current
operating pressure. This assessment can also be easily carried out utilising the approach
specified for the evaluation of corrosion cavities in Australian Standard AS2885. This method has
produced excellent results. Although limited by lack of data, comparison between predicted safe
welding pressures and published values measured on 5 mm thick pipes has been good. The
approach provided a way of assessing burn-through potential which is in agreement with reported
behaviour. Longitudinal welds are more prone to burn-through than circumferential ones of the
same heat input. Pressure has a significant effect. The width or size of the weld is important as
well as penetration. It provides an efficient approach to in-service weld simulation since it does not
require a stress analysis and uses only thermal predictions. Unlike Battelle’s maximum wall
temperature approach, it is more realistic since it accounts for weld orientation and internal pipe
pressure.

Model Validation
The above numerical simulations were validated by comparing predicted values with:

• published result for t8/5 cooling times measured on pipes of 4.8 mm wall thickness,
• measured HAZ and fusion zone geometries from a hot-tap coupon,
• data from welds carried out on an uncooled, empty pipe,
• measured values of t8/5 and HAZ hardness obtained from a number of test welds carried out
on a flow loop at Duke Energy’s Gladstone Gate facility.

This last extensive validation used simulated circumferential fillet welds on three materials, 4.8 mm
and 5.2 mm thick X70, and 6.4 mm thick Ultrapipe X42, under a range of gas pressures and flows.
The t8/5 cooling times were measured, welds were metallographically sectioned and HAZ hardness
determined.

Predicted values of fusion zone depth, HAZ depth, t8/5 cooling time, and HAZ hardness, compared
favourably with measured values. This was an aggressive test of the models’ validity and
accuracy.

Transferring Project Knowledge to Industry.


Although finite element software is readily available, developing and using thermal models to
simulate weld processes is a specialised activity. It would be difficult for industry to attempt this
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 6

analysis without significant investment in software and the development of personnel with
appropriate expertise.

Although access to the modeling capabilities developed within this project will remain, it was felt
that results could be put in a more accessible form. One possibility has been developed to a
prototype stage. This consisted of using the finite element models to develop a database of
predicted values for a range of heat inputs, pipe wall thickness, and heat transfer conditions at the
pipe wall.

Concentrating on the circumferential fillet weld, it has been established that the t8/5 cooling time is
almost independent of pipe diameter provided the heat transfer coefficient at the pipe wall is
constant. That is, the heat transfer coefficient determines the weld cooling rate. This effectively
means that a single model can provide results for any combination of gas pressure, flow rate and
pipe diameter, which gives a constant heat transfer coefficient.

There is a smooth variation in calculated t8/5 cooling time, weld penetration, effective cavity sizes
etc. as heat input, and pipe wall thickness varies. Hence, it is feasible to interpolate values from a
data set with a reasonable degree of accuracy. Using this approach it is possible to develop a
very fast computer program which produces estimates of HAZ hardness, and burn-through risk by
simply interpolating an established database.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 7

1. In-Service Welding of Gas Pipelines:


1.1 Introduction
Metal welding processes are used for the fabrication of structures ranging from, the large and
complex to, the small and simple. The significance of welding may not be directly noticeable,
but it has an important role in the manufacture of many tools, consumer objects, and in almost
all industrial structures. Fusion welding is a significant engineering process because of the
unparalleled advantages it has over other joining methods, and it is used extensively in the
construction of Australia's gas pipeline network.

Unfortunately undesirable changes to material properties can occur during welding and these
have the capacity for generating structural weakness or premature failure. Therefore a large
amount of research has been carried out on pipeline welding to avoid failures and such
detrimental economic and environmental results.

A weld procedure defines all the weld process parameters which must be used in order to
achieve a weld with the required service properties, e.g. type of welding process, (gas metal
arc (GMA), manual metal arc (MMA)), electrode type, voltage, arc current range, welding
speed etc. The present work concerns the development of MMA welding procedures used for
the maintenance and repair of gas pipelines. These in-service welding procedures are carried
out whilst those pipelines are 'live', or in continuous service. Due to the unique conditions in-
service welds and welding procedures have particularly demanding requirements.

1.1.2 Industrial significance of In-service welding


In-service welding may be used as part of a pipeline construction technique called "hot-
tapping". This technique enables the connection of a branch pipe to a pipeline without stopping
or significantly disrupting the gas flow. The major advantage of this operation is that it avoids
the need to decommission the pipeline. That would be costly to the pipeline operator both in
terms of wasted gas and in un-serviced customers. McElligott et al (1) of TransCanada Ltd
have estimated that relative to using a cold connection, a single hot-tap can reduce gross
losses by $1 millon. In a simplified hot-tap, a pipe sleeve is initially welded to the live pipe, and
a slide valve is attached to this fitting, see Figure 1.1(a). The hot-tap drill is fitted to this valve,
see Figure 1.1(b). Next, the drill is used to cut a hole in the wall of the pipe, see Figure 1.1(c).
As the drill is extracted it carries with it the cut-out or coupon, see Figure 1.1(d). Finally the
valve is closed allowing the drill assembly to be removed, see Figure 1.1(e). The success of
the operation depends on the ability to weld the valve assembly or sleeve fitting onto the 'live'
pipeline.

Because of the difficulties associated with this welding operation many hot-taps are currently
carried out under conservative conditions which are achieved by reducing gas pressures and
flows. This can have significant impact on normal pipeline operation. Vented gas and
curtailment costs associated with such planned hot-taps in Australia from 1998-2002 have
been estimated by Venton (2) to be 4 M$. As methane is a green-house gas, purging and
venting of pipelines is potentially hazardous to the environment. TransCanada Pipelines Ltd.
estimates that their use of hot-tapping will avoid an annual emission of 603 kTonne of carbon
dioxide equivalent in 1999 and 2000(3). That represents 18% of the total emissions reduced
by their green house gas management program.

In-service welding can also be used as a technique for pipeline maintenance, to weld
circumferential sleeves at points of pipeline damage. It has also been suggested, Bruce(4),
that weld deposits made directly on to a pipeline could be used to replace pipe wall thickness
lost by corrosion.

The success of such operations depends on safe and effective in-service welding procedures.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 8

Figure 1(b)
Hot-tap drill
attached to Slide
Figure 1(a) Sleeve welded in place and valve
Slide valve attached.

Figure 1(c) Drill used to cut hole


in pipe under pressure

Figure 1(d) Drill and cut


coupon removed and Slide
valve closed

Figure 1(e) Drill removed, branch ready


for connection

Figure 1.1
An illustration of the hot-tapping process taken from IPSCO’s animation of hot-tapping on
http:/www.hottap.com.us
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 9

1.2 In-service welding problems


There are two significant problems associated with in-service welding. Firstly the high gas flow
within the pipe (up to 15 m/sec) causes the weld to cool rapidly due to the convective transfer
of heat from the pipe-wall to the flowing gas. The result of increased weld cooling rates is
greater hardness levels within the weld and in the surrounding heat affected zone (HAZ). With
the increased hardness of the microstructure in the HAZ there is an increased possibility of
hydrogen assisted cracking. The conditions needed for hydrogen assisted cracking include,
hydrogen present to a sufficient degree, tensile stresses acting on the weld, and a susceptible,
hard, HAZ microstructure.

The second problem concerns the risk of bursting the pipe wall during welding. Pressurised
natural gas (up to 15 MPa) imposes a significant stress on the pipe wall, and since the strength
of the pipe is decreased due to the localised heating during welding this can result in failure of
the pipe wall. The result can vary from a small localised bulging of the pipe wall, up to bursting
of the pipe. This is termed ‘burn-through’ and occurs when the region around the weld pool has
insufficient strength to withstand the internal gas pressure, see Figure 1.2.

Figure 1.2
Schematic illustration of burn-through, caused by localised
heating and internal gas pressure.

1.3 Australian Conditions


The recent development of high yield-strength, control-rolled, micro-alloyed steels has allowed
thinner steel pipes to have the same load capacity as earlier, low strength, thicker pipes. The
Australian Pipeline Standard AS 2885, designates that the maximum pressure allowed for
pipeline design is one giving a hoop stress equal to 72% of the yield strength. That is:

P.D = 0.72.σ y
2.t w

where P is the internal pressure, D is the pipe diameter, tw is the pipe wall thickness, and σy
is the minimum specified yield strength.

Hence for a given diameter pipe and gas pressure the tonnage of pipe required for a given
distance can be reduced as the material’s yield strength is increased. Alternatively using high
yield strength pipe permits the transmission of natural gas at higher pressures and flow rates.

The Australian pipeline industry recognises the economic advantages of using high strength
steels. Unfortunately the use of thin walled, high strength steel pipelines increases the
difficulties associated with in-service welding. With the combination of enhanced gas
transmission and diminished wall thickness the weld cooling rate for a given weld procedure
increases. Such high strength steels have a greater sensitivity to strength reduction during
welding and together with the decreased wall thickness are more prone to bulging or burn-
through.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 10

1.4 Summary
Weld procedure development is particularly difficult for in-service welding. For safety and
practicality, experimental test welds can not simply be carried out on ‘live’ pipelines. Hence,
external means of establishing weld procedures have to be used. Traditionally two approaches
have been developed. Laboratory simulation of pipe flow conditions (Edison Welding Institute,
1980-1990 (5,6)) or through the use of simple numerical calculations (Battelle Memorial
Institute, 1985(7,8)). Approximations in such approaches may lead to inaccuracies and excess
conservatism in the choice of weld parameters. Such difficulties will be increased by the
Australian pipeline industry’s use of higher strength, thin walled pipelines because the existing
technology may not apply to such new materials.

1.5 Aim of Current Research


The aim of this current research is:

To develop recommended welding procedures for the


safe and effective in-service welding of thin-wall, high-strength steel,
high pressure, gas pipelines.

The technical challenge is to develop methods of establishing welding procedures which


produce welds that are free from the risk of cracking, and do not risk bursting the pipe wall
during welding: and to confirm their application for the thin walled high strength materials that
will be used in future pipeline construction.

References
1. McElligott J. A., Delanty J., & Delanty B. Full Flow High-Pressure Hot Taps: The New
Technology and Why It’s Indispensable to Industry, Paper Presented at International
Pipeline Conference – pub. ASME v2, 1988, pp813-820.
2. Venton P., Report Prepared for Pipeline Program of Cooperative Research Centre for
Materials Welding & Joining 1996.
3. TransCanada Pipeline, http://www.transcanada.com
4. Bruce W. A., Holdren R.L., Mohr W. C., Kiefner J.F. & Swatzel J.F., Repair of Pipelines by
Weld Metal Deposition, Paper presented at PRCI 9th Symposium on Pipeline Research,
Houston Texas, September 1996.
5. Bruce W. A. & Threadgill P. L. Welding Onto In-Service Pipelines Welding Design &
Fabrication Feb 1991, pp19-24.
6. Cola M.J. & Threadgill P.L., Final Report on Criteria for Hot Tap Welding, American Gas
Association, Edison Welding Institute Project J7038, March 1988
7. Kiefner J.F & Fischer R. D. Models Aid Pipeline Repair Welding Procedure
Oil & Gas Journal March 1988, pp41-47.

8. Fischer R.D., Kiefner J.F. & Whitacre G.R., User Manual for Model1 & Model 2 Computer
Programs for the Predicting Critical Cooling Rates and Temperatures During Repair and
Hot Tap Welding on Pressurised Pipelines, Battelle Memorial Institute Report, June 1981.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 11

2. Research Strategy & Outline

2.1 Background
The development of a weld procedure is essentially a trial-and-error process. For example a
hypothetical structure may require a weld with predefined properties (e.g. penetration).
Through experimental welding trials, involving welding on a replica or similar structure, many
different weld procedures are tested and the one with the closest properties to those desired is
chosen, and replicated in the field. Often weld procedures are set by standards based on past
welding trials and as a result may lack a scientific footing. The cost of establishing a weld
procedure can be large.

In relation to hot-tap welding, the majority of research has involved welding trials, either using a
flow-loop (a diversion adjacent to an operating pipeline which allows welding trials to be carried
out on a test section of pipe without disruption to the existing pipeline) or using a laboratory
simulation. For in-service welds this process is made difficult since the heat loss due to gas
flow cannot be reliably simulated in the laboratory. The details of the normal experimental
approach will be discussed later, but it often consists of using water flow to generate high
cooling rates. Water is a more efficient coolant than gas and hence there is a basic limitation in
duplicating the heat losses due to high gas flow. In addition, with this approach only part of the
problem is addressed. Such tests examine the required penetration, bead shape and HAZ
hardness but fail to consider burn-through, since the system is not pressurised. The
determination of burn-through limits requires further testing. If carried out at all, these tests
would use pressurised, non-flowing gas, and again this represents an approximation to the real
pipeline conditions. Because of the expense and time-consuming nature of this experimental
approach there is a tendency to only examine a limited range of parameters. This can result in
a lack of understanding of the sensitivity of the process to slight variation in welding
parameters, and it may not establish how close a particular weld is to the failure limit. Lack of
knowledge of process sensitivity can cause the welding trial to be both unreliable and unstable.
The economic cost of performing in-service welding trials to this degree of accuracy would be
prohibitive.

Because of the inherent difficulties with an experimental approach some research has used
numerical methods to calculate temperatures and cooling rates. Numerical simulations of in-
service welding offer significant advantages. Firstly this approach avoids or minimises the use
of time-consuming experimentation. The wide range of fittings and pipe geometries that are
used in hot-tapping do not represent a significant problem to computer models. Similarly, the
variation of gas flows and pressures can be economically dealt with. Numerical models also
allow the alternative of determining safe pressures and appropriate gas flows for a given
welding process and this information can facilitate the management of hot-tapping and in-
service welding procedures.

To-date only 2D models have been applied in the pipeline industry. These have only treated
burn-through control in a simple fashion by suggesting that it be signified by a limit to the pipe
wall temperature.

2.2 Project Strategy


This research programme has developed 3D numerical models of in-service welding using the
Finite Element Method. 3D numerical models of the welding process, calculate the thermal
field during welding. This provides estimates of the cooling rate, and by linking this data to
appropriate empirical equations relating hardness and cooling rate the HAZ hardness is also
predicted. Burn-through has been directly examined by combining the thermal analysis with an
elastic-plastic stress analysis. The numerical analysis of burn-through has been a novel
feature of this research programme. In particular this research has developed a new method of
assessing burn-through risk which takes into account the major factors affecting burn-through,
namely, thermal field and its distribution, weld orientation, wall thickness and gas pressure.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 12

Different joint configurations are used by the pipeline industry which require individual analysis.
Australian hot-tap fittings can be broadly classified under three types,
• the full encirclement, circumferential sleeve fitting, see Figure 2.1(a),
• the direct-branch with reinforcement-saddle, see Figure 2.1(b), and
• the direct-branch-to-pipe weld, see Figure 2.1(c).

As the pipe thickness decreases the full encirclement sleeve provides the best structural
support to both the pipe and attachments. Therefore it is the most common joint configuration.
Although longitudinal welds are used to secure the sleeve around the pipe these do not directly
contact the pipeline and therefore are not critical. This research program has concentrated on
numerical simulations of circumferential sleeve welds and branch connections with
reinforcement sleeves.

(a) (b) (c)

Figure 2.1
Common in-service welding pipe configurations:
(a) full encirclement fitting, longitudinal weld to join sleeves and a circumferential fillet to the
run pipe.
(b) Reinforcing saddle around branch pipe.
(c) Directly welding the branch pipe on to the run pipe.

Numerical models of fusion welding processes always include some empirical factors to ensure
that the resulting calculated values agree with those found in practice. This means that such
models can not be created without significant experimental input, and their accuracy must be
validated.

The welding process commonly used for in-service welding in Australia is MMA welding using
hydrogen controlled electrodes. Unlike other welding processes, MMA welding requires
relatively little equipment (power supply + stick electrode) and is the traditional process for in-
field pipeline welding. An experimental assessment of the performance of two typical electrode
types in common use has been carried out using a range of pipe material grades.

Although there is a large body of work on numerical modelling of fusion welding there is little
specifically addressing MMA welding. This research has addressed that deficiency. In
particular it has developed appropriate modelling strategies, for vertical-up or vertical-down
MMA welding positions.

The current numerical simulations have been validated by comparing predicted values with:

• published results for t8/5 cooling times measured on pipes of 4.8 mm wall thickness,
• with measured HAZ and fusion zone geometries from a hot-tap coupon,
• with data from welds carried out on an uncooled, empty pipe,
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 13

• measured values of t8/5 and HAZ hardness obtained from a number of test welds carried
out on a flow loop at Duke Energies Gladstone Gate facility.

- The results of thermal analysis have been directly compared with micrographs of test
welds to ensure the correct calculation of weld penetration and HAZ geometry.

- The measured cooling rates of test welds have been compared with predicted values.

- HAZ hardness has been measured for a range of materials and test welds and has
allowed further comparisons between predicted hardness and measured values.

This last extensive validation was an aggressive test of the model’s validity and accuracy under
operational conditions on a live pipeline.

Although finite element software is readily available, developing and using thermal models to
simulate weld processes is a specialised activity. It would be difficult for industry to adopt this
approach without significant investment in software and the development of personnel with
appropriate expertise.

Although access to the modelling capabilities developed within this project will remain, it was
felt that results could be put in a more accessible form. One possibility has been developed to
a prototype stage. This consists of using the finite element models to develop a database of
predicted values for a range of heat inputs, pipe wall thickness, and heat transfer conditions at
the pipe wall. Using this approach it is possible to develop a very fast program which produces
estimates of HAZ hardness, and burn-through risk by simply interpolating an established
database.

It is anticipated that through the development of improved numerical simulations of in-service


welding, and the establishment of their accuracy and scientific credibility, more efficient weld
procedure development will result. Validated numerical models will also allow a safe
combination of welding procedure, gas pressures and gas flows to be determined for a given
pipe geometry, and this in itself will facilitate more efficient management and control of in-
service procedures.

2.3 Outline
Section 3 will discuss the experimental work that has examined in-service welding. The areas
of post weld hardness and the possibility of pipe wall failure during in-service welding will be its
foci. Section 4 will introduce the concepts related to the numerical simulation of welding
processes. Sections 5 & 8 will concentrate on the computer simulation of in-service welding,
and on work related to the development of a numerical approach to the prediction of safe
welding procedures.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 14

3. Literature Review:
Experimental Studies of In-Service Welding
3.1 Introduction
This section reviews the past research on in-service welding. In particular it will identify the need for
additional research and the reasons for the current research activities.

The research work into in-service welding can be grouped into three areas, namely:
• experimentally determined weld procedures,
• experimental studies of burn-through and investigations of weld repair, and,
• the development and application of numerical simulation.

This list does not imply that there is a large amount of information available, for published research
work and data related to in-service welding is rare. Most work has been carried out by the American
Gas Association’s Pipeline Research Committee, in sponsored work at the Edison Welding Institute
(EWI), by Cola & Threadgill (1) and Bruce & Threadgill (2,3), and at the Battelle Memorial Institute
(BMI), by Kiefner & Fischer (4) and Fischer et al (5). The results of that work form the basis of the
common methods used to establish appropriate in-service welding procedures.

EWI (2) developed a method for experimentally establishing in-service welding procedures, following
similar work by British Gas. At BMI, Fischer, Kiefner & Whitacre (5) developed a numerical approach,
and produced commercial software to predict weld cooling rates and the possibility of the pipe bursting
(burn-through) during in-service welding. In collaboration with the EWI, their numerical approach was
validated and connections between the two approaches established (6). Other literature is sporadic,
generally relating to limited experimental studies of particular hot-tap welding conditions. Phelps et al
(7), Wade (8,9,10) and Bruce et al (11,12), have produced data on burn-through limits.

3.2 Experimental Studies of In-Service Welding


- Factors controlling cracking
The older generation of steel pipes has compositions (high carbon equivalent) which are susceptible to
hydrogen assisted cracking. Much experimental work on in-service welding was therefore concerned
with determining which of the process variables controlled post weld hardness, since this is commonly
the property used to assess crack susceptibility. Figure 3.1 broadly summarises the relationships
between post weld hardness and the weld process variables.

Bailey’s work (13), reported by Graville & Read (14), identified that with manual metal arc (MMA)
welding using low hydrogen or rutile electrodes, cracking did not occur below a critical HAZ hardness
of 350 HVN. Graville & Read considered that in Bailey’s work even the ‘low-hydrogen weld’ exhibited
significant hydrogen implying some conservatism in the critical hardness specified by Bailey.

For in-service welding, as in any fusion weld, the cooling rate and the chemical composition of the
steel are the main factors influencing post weld hardness. The weld heat input clearly influences the
cooling rate, but for in-service welding the complicating factor is the high heat loss generated by the
gas flowing in the pipe. This accelerated cooling, which in-turn depends on the pressure, the flow rate
of the gas and the pipe geometry, has a dominating influence on the process.

Experiments on operational pipelines and under simulated conditions have clearly shown that gas flow
strongly increases the cooling rate of the weld, particularly in thin walled pipes (1,3), see Figure 3.2 for
example. This observation is of particular relevance to Australian conditions, where there is a
significant trend towards the use of thin-walled pipes. Phelps et al(7) reported that the HAZ of in-
service welds could become brittle, giving rise to cracking immediately after welding, or at a later stage
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 15

through hydrogen embrittlement. They quoted the case of a 200mm diameter 8mm thick, steel pipe
with a carbon equivalent (CE) of 0.48. Welds on this pipe gave a HAZ hardness of 415 VHN for a gas
flow of 0.518 scm/h and 285 VHN with no gas flow. Whilst their main concern was with the pipe’s HAZ
hardness they also indicated that the composition and cooling rate within the fitting should not be
ignored.

Cola et al(1) reported on the cracking propensity of welds made with basic and cellulosic electrodes.
They considered that with cellulosic electrodes a limiting t8/5 (time to cool from 800°C to 500 °C) could
be established independent of pipe composition. This conclusion was based on data generated from
simulated in-service welds using an E6010 electrode at 1kJ/mm heat input. For a range of steels with
a CE of 0.3-0.5 cracking was only found for t8/5 cooling times below 5 seconds. When using basic
electrodes they recognised that the pipe composition was an important factor, and determined that
cracking was only a concern for an HAZ hardness >400 VHN. They also reported that, provided basic
electrodes and a sound low-hydrogen welding practice were used, the risk of cracking was not
significant for steel having a CE < 0.5. The significance of this relates to the age of the pipe since
modern steel compositions generally give a CE < 0.4-0.45. The use of basic electrodes is not just
related to their cracking propensity. Whilst they clearly reduce the risk of hydrogen embrittlement, as
Phelps et al(7) showed, they also generate significantly lower penetration for a given heat input and
hence reduce the risk of burn-through (see Section 9). Welding vertically-down or welding with the
electrode DCEN were also reported (7) to give a lower penetration than welding vertically-up, or using
electrode DCEP.

Boran(15) found that post-weld hardness was influenced by the electrode polarity of the MMA welding
process. The polarity influenced the apportionment of heat between the electrode and the workpiece,.
DCEN gives the greater fraction of heat in the weld region, which reduces the weld cooling rate and
gives the least post weld hardness for a given heat input.

Preheating the joint before welding is an obvious, traditional, way of controlling and reducing the
cooling rate of during welding. For example, DeHertogh & Illeghems (16) preheated a 323.5 mm
diameter 4.4 mm thick, X60 pipe with a gas flow of 25,000 m3/hr at 4.8 MPa. They found that the
hardness of a weld decreased considerably from 367 to 317 when using an 80°C preheat. Using
inductive heating they examined preheat levels from 50-200°C and reported that minimum hardness
was achieved at a 100°C preheat, although no explanation of the minimum value was proposed.

Cassie et al (17) listed the recommended features of a satisfactory in-service welding procedure as;
use of basic low hydrogen electrodes, a preheat of 100°C for material with a CE < 0.4 and 150°C for
material with CE>0.4, and the use of a stringer bead technique.

Preheating has an economic penalty of course, but for in-service welding there are other significant
difficulties. These relate to the extremely high heat loss generated by the flowing gas. Under such
conditions achieving a consistent preheat is difficult. A number of methods were tried, including direct
gas flame heating, electrical resistance and inductive heating. Phelps et al (7) reported that none of
these was entirely satisfactory. The method they recommended was direct flame heating using a
hand-held propane torch. This was used to heat the region ahead of the weld to a maximum
temperature of 250°C. This preheating step was followed by welding for a short time, until the
temperature of the region fell below the desired preheat level. This cyclic process was then repeated.
Cooling rates were extremely high so it followed that welding-runs would be short. Since consistency
of heat input relies on manual skill of the welder, such an intermittent process can only contribute to
the variability in welding speed and in heat input.

In-service welds are multi-pass, so the sequence of welding can be controlled to minimise HAZ
hardness. Using a stringer bead technique, Figure 3.3(c), in which the current weld tempers the
hardening created by the previous one, is recommended by Cassie (17) and Bruce & Threadgill (2).
Bruce (18) describes the sequence recommended by British Gas which uses a buttering layer and a
temperbead sequence, also shown in Figure 3.3(b). Rietjens (19) made reference to a desirable weld
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 16

preparation for split circumferential sleeves, as shown in Figure 3.3(a) and identified that buttering and
temperbead sequences were useful in minimising crack susceptibility. He also advocated using weld
metal with low yield strength, in order to relieve residual stress.

Cassie et al (17) examined the application of post-weld heating as a means of reducing the hardness
at the weld toe. They found that using a gas tungsten arc was effective but they did not consider it a
viable field technique, other methods were not reported in detail but were considered ineffective.

Variability in the manual welding process was identified as a concern by a number of researchers.
Cola et al (1) referred to the inherent variability in the manual process. Cassie (17) identified variations
in hardness due to different welders using different welding speeds and hence a varied heat input.
Bruce et al (3) recommended the use of controlled deposition rates in order to minimise such
variations. There are some reports of a systematic variation in weld properties around circumferential
welds. DeHertogh et al(16) indicated an increase in the depth of HAZ as the weld progressed from the
top to the bottom-dead-centre of a circumferential MMA weld. They considered that this was a
consequence of the general heating of the pipe as welding progressed. For 4.8 mm thick pipe and
vertical-down MMAW Phelps et al (7) also found that penetration was significantly greater at the 6
o’clock position (1.3-1.4 mm) than at the 3 o’clock position (1.0 mm).

3.3 Experimental Studies of In-service Welding


- Factors Controlling on Burn-through
Considering the safety implications, there have been few attempts to determine the conditions
necessary to avoid pipe-wall failure during in-service welding. Experimental work has generally used
a small number of test welds under widely varied experimental conditions, so conclusions tend to be
general directions rather than quantified limits. The relevant factors are shown in Figure 3.4. Clearly,
the risk of burn-through is related to the loss of pipe wall strength in the weld zone, and its inability to
resist local stress. The reduction in wall strength depends on the elevated temperature around the
weld, and on the depth of weld penetration relative to the original wall thickness. Observations of
burn-through generally show significant local plastic distortion of the pipe wall, and a fracture along the
weld pool axis (11,12).

3.3.1 The influence of pipe wall thickness on burn-through


In-service welds on thin pipe walls have a high risk of burn-through. Weld penetration is largely
influenced by the welding heat input, so the same heat input on a thinner walled pipe causes a greater
relative reduction in the wall strength. Research efforts have generally sought to determine the lower
limit of pipe wall thickness that could be safely welded. For longitudinal welds on X60 using 3.2 mm
and 4 mm diameter electrodes, Wade(8) found that successful welds could be made on 6 mm thick
pipe at normal operational pressures with a heat input of up to 1.8 kJ/mm. For 5 mm thick pipe he
considered that welding could still take place but recommended a “considerable restriction” of
pressure (below 3 MPa) and heat input (less than 1.4 kJ/mm). He established that burn-through was
probable for pipe with 3 mm wall thickness, even at low pressure. Specifying a minimum pipe-wall
thickness below which in-service welding should not take place has been a convenient way of
designating a safe procedure. It is generally accepted that burn-through risk is minimal for pipes
which have walls thicker than 6.35 mm(18). Many operational standards have restricted in-service
welding by specifying a minimum wall thickness, with restrictions in the range 4-5 mm being
common(20).

In 1983 Hicks(21) listed a recommended approach for avoiding burn-through. This indicated that at
the ‘maximum allowable operating pressure’ (MAOP), {equivalent to the pressure giving a hoop stress
of 72% of the yield strength} burn-through was possible with a 4 mm wall thickness. His
recommendations included limiting in-service welding for longitudinal welds to pipes greater than
4.8mm thick, and for circumferential welds to greater than 4mm wall thickness. He also recommended
that pressure should be restricted to that specified by the ASME Gas Piping Standard,
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 17

1.44.σ y (t − c )
p=
D
Where p is pressure, σy is minimum yield strength, D is pipe diameter, t is pipe wall thickness, and,
c is an assumed thickness reduction, to account for reduced wall strength equivalent to 2.38 mm.
3.3.2 Influence of heat input on burn-through
As the welding heat input is increased the weld penetration, and size of the heated region increases,
with a consequent increase in the possibility of burn-through. Cassie(17) investigated the process
parameters controlling weld penetration using various out-of-position welds on 6, 9 and 12 mm thick
plate. From these results, he determined that welds in a vertical-down position using basic electrodes
provided the lowest penetration, and therefore would be the most appropriate for in-service welding.
Cassie(17) used simulated in-service welds on pressurised cylinders of 450 mm diameter X52 steel
with wall thickness of 3.2, 4.8 and 6.4 mm. He defined a safe limit by specifying the maximum arc
current allowable for a given pipe thickness, see Figure 3.5. Safe welding currents could be
determined for all pipes with wall thickness above 3.2 mm. However, the final recommendation was
that welding should not take place on pipes of less than 4 mm thickness at an internal pressure
greater than 7 MPa.

Bruce et al(11) also referred to a restriction on welding current for safe in-service welding on pipe walls
≤ 3 mm thick. They identified that at the same heat input, welding with a smaller diameter electrode
(equivalent to a reduced current) reduced the burn-through risk. The limits suggested are shown
diagrammatically in Figure 3.6. This shows an interesting difference in behaviour between 3.2 mm
and 4 mm thick pipes. For a 3.2 mm pipe wall the division between safe and unsafe welding was
strongly dependent on electrode diameter (arc current), with 2 mm diameter (50 A) electrodes safe,
and 2.4 mm diameter (80 A) electrode borderline. With a 4 mm thick pipe-wall the conditions are more
dependent on heat input. This apparent difference is unexplained. Bruce et al(11, 12), determined
recommendations for safe in-service weld repair as:
• A maximum internal pressure of 6.7 MPa during welding on a minimum remaining wall thickness
of 3.2 mm,
• Electrode type to be a hydrogen controlled, E7018, with electrode size to be restricted to 2.4 mm
diameter,
• A heat input of 0.51 kJ/mm for the first weld runs.

3.3.3 Influence of welding technique on burn-through


Manual metal arc welding (MMAW) is a process requiring considerable skill and hence the control of
penetration and heat input is welder dependent. Wade(8) recommended that electrode types should
be specified for their capacity to run smoothly and produce uniform penetration at all points in the
weld. He also recommended that welds should have equal leg length and that weaving should be
minimised and a planned sequence of short welds should be used to minimise heat build up ahead of
the welding arc. Phelps et al(7) observed that penetration was greater at the 6 o-clock position, during
vertically-down welding.

3.3.4 Influence of preheat on burn-through


Cassie(17) considered that preheat did not influence burn-through limits. Wade(9) found a small
effect, noting an increased tendency for local bulging with increased preheat. However, for typical
preheat levels of around 100°C the effect was not large.

3.3.5 The influence of pipe diameter on burn-through


Wade(10) considered that the diameter was not important. He argued that internal pressure acting on
the weakened pipe wall was the main driving factor, not the hoop stress. Bout & Gretskii(22) however
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 18

considered burn-through limits in terms of hoop stress, implying that the pipe diameter has an
influence.

3.3.6 The influence of welding direction on burn-through


Longitudinal welds are more prone to burn-through than circumferential ones. Bout & Gretski(22)
quote limiting heat inputs for welds in both directions indicating that welds in a circumferential direction
can tolerate higher heat input before burn-through, see Table 3.1. This observation appears to point
towards the importance of the applied hoop stress on burn-through, and it is difficult to rationalise this
with Wade’s(10) conclusion that the pipe diameter is not significant.

3.3.7 Influence of internal pressure on burn-through


Internal pressure is recognised as a factor influencing burn-through, although Bruce et al(11)
considered it secondary to heat input. Wade(8) carried out longitudinal fillet welds on 250mm
diameter cylinders pressurised with nitrogen. For safety, a mechanised welding system was used in
the down-hand position. He observed that significant plastic deformation occurred within the weld
zone prior to bursting. Therefore he measured the local pipe wall deformation at the weld site and
determined the onset of burn-through to be a critical local bulge height of 1 mm. He produced graphs
relating pressure, bulge height and heat input, and determined a critical pressure/heat input line for
burn-through (see Figure 3.7). His work lead to a diagram, as shown in Figure 3.8, which specified
acceptable working zones (pressure, versus heat input) for in-service welding on pipes of different wall
thicknesses.

3.3.8 The influence of pipe grade on burn-through


Steel undergoes a dramatic reduction in strength as its temperature is increased, such that at
temperatures over 800°C its yield strength is 4-10% of its room temperature value. The strength at
temperature of a higher grade steel such as X70 is not significantly higher than that of a low strength
grade. So although a high room temperature strength allows pipes to be thinner, this increased
strength is not present in the weld pool region during welding. This non-proportional reduction in
strength is an additional factor that increases the burn-through risk with thin X70 or X80 pipe materials.

Because many of the above observations have been made on statically pressurised pipes, and the
gas flow within a normal operational pipe is an aggressive coolant, it is thought that many of these
burn-through conditions are conservative. It is interesting that Hicks(21) recommended that a
minimum gas flow of 0.4 m/sec be maintained during welding since this acts to reduce pipe wall
temperatures. It is recognised that the unsafe region for MMA welding is on pressurised pipes with
wall thickness around 4 mm, however no clearly defined, quantitative limits exist.

3.3.9 A Burn-through Avoidance Strategy


Pipe failures at defects such as corrosion cavities or notches are referred to as pressure controlled
failure. Bursting strength is dependent on defect dimensions and yield strength. Bout & Gretskii(22)
were the first to use this approach to estimate the load carrying capacity of the pipe during in-service
welding. They considered that above 700°C the metal’s strength was effectively zero, therefore, the
700°C isotherm around the weld pool could represent a surface defect, of a given depth and length.

They determined that the following formulae could be used to estimate the maximum hoop stress that
could be sustained during welding.

⎛t ⎞ ⎛t 0.85 ⎞
σ θ = σ 0.2 ⎜ − 0.85 ⎟ / ⎜ − ⎟
⎝r ⎠ ⎝r M ⎠

Where,
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 19

0.4 L
M = 1+
Rt
t is the pipe wall thickness
L is the axial length of the weld pool perpendicular to the hoop stress
r is the maximum depth of the defect
R is outside radius of the pipe

Using these formulae Bout & Gretskii(22) then estimated the critical length of heated zone (Lcritical)
before the pipe would not be able to operate at its maximum load carrying capacity.
⎛ ⎛ t ⎞⎞
Lcritical = 1.12 2 Rt. ⎜⎜ 0.85 / ⎜ 0.1 − 9.935 ⎟ ⎟⎟ − 1
⎝ ⎝ r ⎠⎠

This was a potentially valuable approach to assessing the impact of a weld on the load carrying
capacity of the pipe but Bout & Gretskii(22) did not continue with this or integrate this method with a
numerical thermal model. Instead, they determined that the permissible sizes of heated zone were
very restrictive and suggested a way of overcoming this.

They considered that by increasing the support of the pipe wall during welding the freedom for the
pipe wall to plastically deform would be reduced, and burn-through would be prevented. They
achieved this by surrounding the weld with reinforcing rings or bands as shown in Figure 3.9. The
gap between the rings could be adjusted to give varied support to pipes of different wall thickness or
pressure. The relationship between allowable pressure, effective thermal penetration and gap width
(ao) between the reinforcing rings was given as follows.

p = 4σ 0at.2700C (t − r ) / a o
2 2

For a 3 mm wall thickness, 320 mm diameter pipe with an internal pressure of 4 MPa, burn-through
occurred at a heat input of 0.475 kJ/mm with a normal sleeve. With a constraining band the critical
heat input was 0.915 kJ/mm.

This represents an interesting novel approach. The only concerning feature is that the practical field
application quoted are for 9mm thick pipes. Also no consideration was given to the increased cooling
rate that would be generated within this joint configuration.

3.4 The Determination of Safe In-Service Welding Procedures


The major thrust of research work on in-service welding has been to generate sound welding practices
and to determine processes whereby weld procedures can be easily and efficiently established for a
given hot-tap. Because the gas flow, gas pressure and pipe geometry strongly influence welding
outcomes this latter aspect is particularly relevant, since a weld procedure must be established for
each individual hot-tap. As Bruce et al(3) indicates, many codes require that a qualified welding
procedure must be determined for a given in-service weld. AS 2885(23) requires that, ‘welds shall be
made following a qualified welding procedure which takes into account pressure and cooling effects
from the flow of fluid within the pipe and simulates site condition’.

Although such techniques seem essential, not all pipeline companies use them. It has been
reported(3) that of eleven member companies of the American Pipeline Research Committee, only
three used specific procedure development. The others either weld on decommissioned pipelines, or
weld using reduced pressures, flows and preheat to overcome high cooling rates.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 20

An important feature of procedure development concerns selecting the minimum suitable heat input.
This should be selected to achieve a weld with HAZ hardness below the level likely to cause cracking,
but not be so high that the pipe wall may burn-through during welding. The most comprehensive
approach is to carry out a conventional weld procedure development using test welds on a special flow
loop or pipe by-pass. This provides a controllable segment of pipe under identical conditions to those
in-service. However such facilities are not common, and the normal approach is to physically simulate
the thermal characteristics of the operational pipeline by using water, air or oil flow through a section of
test pipe.

Following on from early work by British Gas, EWI developed an experimental method of determining
the cooling capacity of a working pipeline(2). This is achieved by a simple procedure suitable for use
in the field. An oxy-propane torch is used to heat a 50mm diameter area on the exposed pipe wall.
Heating is stopped when the temperature reaches approximately 325°C. Using a stopwatch and a
thermometer, the temperature drop is monitored as the pipe wall cools. The time taken to cool
between 250 and 100°C is taken as the cooling capacity of the pipeline. This value is often referred to
as the EWI cooling time, Tewi. The experiment is repeated on other spots upstream from the first, and
a final cooling time is arrived at by averaging six readings.

The philosophy behind the test is to measure a parameter which reflects the cooling capacity of the
working pipeline, and to do this in a safe, experimentally simple and practical way. An experimental
test bed can then be set up by duplicating the measured Tewi for this test pipe. To use this method for
establishing a physical simulation it is not necessary to adhere to the EWI procedure. Simply
duplicating the experimental procedure, heat source, spot size, temperature range etc, in the field and
in the workshop should be sufficient. The general philosophy is satisfied provided the heat transfer
behaviour of the gas or the test fluid is not significantly altered by the wall temperature achieved during
welding as compared with that generated during EWI testing. Tests have been reported in which
water, water-mist sprays, compressed air or oils are used to achieve the desired cooling capacity.

EWI’s work also generated much experimental data relating the weld heat input and cooling rate for in-
service welds. The data were measured from flow loop tests and from test welds carried out under
simulated conditions. The t8/5 weld cooling times were measured for heat inputs between 1 – 2 kJ/mm
using pipe materials ranging in thickness from 4.8-8.0 mm and E6010, E7018 and E8018 electrodes.
The cooling capacity was measured using the EWI process and t8/5 values were determined from
temperatures measured by thermocouples harpooned into the weld pool.

For a given pipe thickness and gas flow an approximately linear relationship between t8/5 and heat
input were often found. In addition, EWI established that there was a proportional relationship linking
the cooling capacity Tewi, and the t8/5 value obtained for a given heat input. Experimental data were
used to form graphs such as Figure 3.10 for a 4.8 mm thick pipe. To use this graph to derive a
suitable heat input requires a measurement of the cooling capacity of the operational pipeline. Take
the value of 20 seconds for example. This represents a particular line between t8/5 and weld heat input
on Figure 3.10. For a given CE, the IIW relationship between hardness and t8/5 can be used to
estimate the minimum t8/5 cooling time required to give a hardness level of 350 HVN, (6 seconds in
this example). The point on the relevant cooling capacity line at this t8/5 value gives the required
minimum heat input (1.1 kJ/mm). The empirical data used to establish these lines are also shown in
Figure 3.10, so it can be seen that the data points are relatively sparse and the extrapolations have
generally been chosen conservatively. With this approach, the required heat input to achieve a
desired hardness can be estimated directly from the measured heat capacity.

To make use of the existing EWI data linking t8/5, and EWI cooling time it is necessary to have more
concern about using the same procedure as EWI for determining the cooling capacity. The EWI test is
not rigorously defined, and there is some concern about possible variability due to the changes in the
size of the heated region and the heating rate used (see Appendix 1). EWI carried out some numerical
analysis of the test process and concluded that the cooling capacity was not overly sensitive to the
chosen spot size or heating rate(1). There are other difficulties here however, since the relationships
between heat input and t8/5 are empirical and may vary with electrode type, welding position, joint type
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 21

etc. Similarly, the EWI approach makes no reference to fluid type, joint position, weld preparation,
electrode type or preheat. It simply relies on the measured heat capacity and the developed empirical
data set. It was pointed out by Cola & Threadgill(1) that the correlation between heat sink capacity and
the cooling rate of the weld had only been established over a limited range of field conditions.
Certainly the data are sparse and requires considerable extrapolation in some instances. Data are not
available for pipes less than 4.0 mm thick.

EWI was mainly concerned with determining a weld heat input that would produce a suitable HAZ
hardness. The approach is conservative. Empirical relationships are chosen such that t8/5 values are
likely to be under estimated, and heat inputs are likely to be over estimated, thus being sure to obtain
hardness below the chosen limit. The critical level of hardness is chosen at 350 HVN, which is also
considered as a conservative limit for low hydrogen welds(14).

Both physical simulation and the EWI test have some problems. There is the additional cost of
determining the cooling capacity of the pipeline, for although it is a simple process it may require
excavation and incur site costs. To use the EWI methodology requires considerable empirical data to
build up a database relating EWI cooling time, t8/5 cooling time and heat-input. The EWI methodology
does not make any distinction between the saddle or sleeve configuration, and does not consider the
effect of multi-pass welds or tempering effects. It also provides no direct information about burn-
through risk.

Simulating the field cooling capacity directly and developing the welding procedure on a sample of the
same pipe clearly overcomes the reliance on a database, but in some cases achieving the same
cooling capacity can be difficult. Figure 3.11 illustrates the difficulties of selecting a fluid that is
capable of duplicating the thermal characteristics of flowing high-pressure gas. In this figure the
convective heat transfer coefficients for a selection of test fluids has been estimated using the Dittus-
Bohler equation (Holman(24)). The heat transfer coefficient for methane gas flow under typical
operational conditions is also included. It seems reasonable to assume that a fluid that gives a similar
heat transfer coefficient to that of methane would be suitable as a experimental substitute. Air clearly
gives much lower values than methane at the same flow rate. Engine oil likewise gives lower
convective coefficients. A low viscosity cooling oil does give heat transfer coefficient similar to that of
methane, but it also requires a similar flow rate (1-10m/sec). Such high flow rates are not practical
under laboratory testing. Water generally gives higher values and is clearly a more aggressive coolant
than a gas flow. Values equivalent to gas flow could be achieved by using a low flow of water, <0.1
m/sec. However, this analysis does not include the boiling or vaporisation of water, which drastically
increases heat extraction.

The difficulty of generating cooling rates similar to gas flow is also illustrated in Figures 3.12 & 3.13.
In these figures, data from EWI are replotted to illustrate how test pipes artificially cooled with water
generally have more aggressive cooling than pipe cooled by gas flows. In these graphs, this effect is
manifested as a reduced t8/5 cooling time for a given heat input when using water. Oil and air both
give less severe cooling characteristics. A comparison between Figure 3.12 for 6.35 mm thick pipe,
and Figure 3.13 for 4.8 mm thick pipe demonstrates that the differences between these fluids
increases for thin walled pipe.

Bruce et al(3) attempted to set up guidelines for the appropriate selection of model fluid, water, oil or
air to be used in qualification procedures. Often the rationale is to accept the more aggressive
coolant, normally water, since then the experimentally determined critical heat input value is
conservative. That is on the operational pipeline it will produce longer cooling times and less HAZ
hardness. It should be borne in mind however, that an over conservative heat input with respect to
hardness could be non-conservative with respect to burn-through. Unless care is taken to match
cooling capacities, Figures 3.12 & 3.13 indicate that this conservatism would increase as the pipe wall
thickness is decreased, because there is a greater differential between the cooling effects of water and
gas flow. This trend combined with the increasing sensitivity of thin walls to burn-through should be of
concern.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 22

The simulated flow is usually unpressurised, so these tests do not rigorously establish burn-through
conditions. If burn-through limits are required, these are commonly determined with a static
pressurised test using compressed inert gas. A test vessel is fabricated using a sample length of pipe.
This should include an internal cylinder to reduce the volume of compressed gas and increase safety.
The static gas has a lower cooling capacity than the gas flow in an operational pipeline so burn-
through limits established on such test are also conservative. That is, for a given heat input wall
temperatures during the test are likely to be higher than those achieved on the working pipeline.
Hence, under test conditions burn-through is likely to take place at a lower heat input than it would do
in practice.

3.5 Summary Of Section 3 and Required Research


With care and systematic planning it is clear that systems and technology are in place to allow in-
service welding to be carried out on pipes with wall thickness greater than 6 mm. However as pipe
wall thickness is reduced it is clear that experience and information about in-service welding becomes
less and the process becomes inherently more difficult.

3.5.1 Increased use of thin walled pipes – burn-through limits


In Australia, there is a significant trend towards the use of high yield-strength steels for pipeline
construction. Future pipelines using X70 and X80 steels could have wall thickness as low as 3 mm.
The reduced wall thickness is more sensitive to strength loss and increases the risk of burn-through
during welding. Thin walls are more easily cooled by the flowing gas and high strength steels can be
susceptible to the generation of excessive hardness for a given cooling rate. If the economic
advantages of in-service welding are to be maintained then technology to support the safe and
effective welding of thin-walled high-strength pipelines must be established.

This inevitably raises a question about the wall thickness at which in-service welding becomes unsafe.
The consensus is that the current lower limit is approximately 4-5 mm, although welds have been
made on pipes of 3.2 mm thickness. Current burn-through limits largely rely on experimental data that
have been determined under conservative conditions, or use a restricted pipe wall temperature. The
degree of conservatism in the established limits is unknown. Clear information on the heat input and
pressure limits to avoid burn-through particularly on thin walled pipes is desirable.

3.5.2 Experimental weld procedure development


Weld procedure development is particularly difficult for in-service welding. Clearly, for safety and
practicality, experimental test welds can not be carried out on ‘live’ pipelines. Hence other means of
establishing weld procedures have to be used. Simulation of accelerated cooling generally use
systems that give greater cooling rates during the test than will be experienced on the live pipe. This
generates a weld procedure with a conservative heat input with respect to the generation of excess
hardness. However this may be non-conservative with respect to burn-through. In such cases, say
with pipe < 6.4 mm thick, then additional experiments on pressurised segments of pipe are generally
required. Again because such tests rarely use flowing gas the heat input limits will be conservative.
Such conservative outcomes may unnecessarily restrict in-service welding to thicker pipes.

Approximations in such systems may lead to inaccuracies and excess conservatism in the choice of
weld parameters. Such difficulties will be enhanced with thin walled pipes. On thin pipes the heat
input to avoid burn-through is very restricted, and it is necessary to recognise the danger of the
potential loss of control over heat input in a manual process.

3.5.3 Heat Input Variation


Manual processes are inherently variable in welding speed, and hence in heat input. This is a clear
problem for a welding procedure that depends on meeting a critical working range of heat input.
Interestingly the possibility of addressing this through the introduction of automated systems was
raised by Cassie(17) some 20 years ago. He suggested the possibility of using an automated GMA
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 23

welding system to accurately control heat input. In addition, he speculated on the possibility or
incorporating pre-heat or post heating within such a system.

3.5.4 Hardness limits to assess potential for cracking


With appropriate validation computer models can reliably predict weld cooling behaviour, however
hardness is conventionally used to assess the potential for cracking. This raises two possible gaps in
our knowledge. Firstly there is some inaccuracy possible in the determination of hardness from the
relationships between composition, and cooling rate. There are many such relationships, which may
have varied relevance for certain ranges of steel. The current solution to this is to adopt the
conservative approach, but here again for thin-walled pipe this may lead to the selection of higher heat
inputs than necessary, and run into conflict with burn-through conditions. Secondly, HAZ hardness
may not be the sole criteria for the assessment of hydrogen assisted cold cracking and t8/5 may not be
the most relevant factor to assess weld cooling.

3.6 Technical Challenges


Techniques for in-service welding on pipes with wall thickness greater than 6 mm are well established.
Provided a careful and systematic approach is taken, such welds can be produced with safety. To
apply similar techniques to pipes with wall thickness below ≈ 5 mm, or for weld repair where remaining
wall thickness is low is a technical challenge. For it is below this thickness, that our conservative
approach, and lack of knowledge about suitable HAZ hardness levels, burn-through limits, and the
impact of a poorly controlled manual process, becomes prohibitive.
The technical challenge is to address in-service welding on the thin walled high strength materials that
will be used in future pipeline construction. The major aspects will be:

• the determination of quantifiable, validated burn-through limits for thin walled pipes,
• improve confidence in the hardness limits that are used for modern pipe compositions,
• quantifying the role of temperbead techniques, and
• addressing improved control of heat input.

3.7 References
1. Cola M. J. & Threadgill P. L., Final Report on Criteria for Hot Tap Welding, American Gas
Association, Edison Welding Institute Project J7038, March 1988.

2. Bruce W. A. & Threadgill P. L., Welding Onto In-Service Pipelines,


Welding Design & Fabrication Feb 1991, pp19-24.

3. Bruce W. A. & Threadgill P. L., Effect of Procedure Qualification Variables for Welding Onto In-
service Pipelines, American Gas Association Report J7141, July 1994.

4. Kiefner J.F & Fischer R. D., Models Aid Pipeline Repair Welding Procedure,
Oil & Gas Journal March 1988, pp41-47.

5. Fischer R. D., Kiefner J. F. & Whitacre G. R., User Manual for Model1 & Model 2 Computer
Programs for the Predicting Critical Cooling Rates and Temperatures During Repair and Hot Tap
Welding on Pressurised Pipelines, Battelle Memorial Institute Report, June 1981.

6. Bruce W. A., Bubenik T. A., Fischer R. D. & Kiefner J.F., Development of Simplified Weld Cooling
Rate Models For In-Service Gas Pipelines, Line Pipe Research Proceedings, 8th Symposium,
September 1993, Paper 31, pub Arlington VA 22209, 1993, pp31.1-31.22.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 24

7. Phelps B., Cassie B. A., & Evans N. H., Welding Onto Live Natural Gas Pipelines, Metal
Construction, August 1976, pp350-354.

8. Wade J. B., Hot Tapping of Pipelines, Australian Welding Research Association Research Report,
Snowy Mountains Corporation 1978.

9. Wade J.B., Effect of Preheat on Hot Tapping Procedures, Australian Welding Research
Association Research Report, Snowy Mountains Corporation, September 1978.

10. Wade J.B., Description of Experimental Results on the Effects of Pipeline Damage on
Performance and Hot Tapping Techniques, paper presented at Australian Welding Research
Association’s conference Pipeline Welding in 80’s, Melbourne March 1981, Paper 4a.

11. Bruce W. A., Holdren R.L., Mohr W. C., Kiefner J.F. & Swatzel J.F., Repair of Pipelines by Weld
Metal Deposition, Paper presented at PRCI 9th Symposium on Pipeline Research, Houston Texas,
September 1996.

12. Bruce W. A., Holdren R.L. & Mohr W. C., Repair of Pipelines by Direct Deposition of Weld Metal –
Further Studies, Final report Edison Welding Institute, EWI Project J7283, November 1996.

13. Bailey N., Welding Procedures for Low Alloy Steels, The Welding Institute Cambridge England
1970.

14. Graville B. A. & Read J. A., Optimization of Fillet Weld Sizes, Welding Journal Research
Supplement pp161s-167s.

15. Boran J
The Hot-Tapping of Sub Sea Pipelines
Welding Review, vol6, no 4 Nov 1987, pp283-284

16. DeHertogh J. & Illeghems H., Welding Natural Gas Filled Pipelines, Metal Construction & British
Welding Journal, March 1972, pp224-227.

17. Cassie B. A., The Welding of Hot Tap Connections to High Pressure Gas Pipelines, paper
presented Pipeline Industries Guild J. W. Jones Memorial Lecture, October 1974.

18. Bruce W. A., Welding Onto In-Service Pipelines: A Review, paper presented at Pipeline Welding
’98, International Symposium on Pipeline Welding, May 1998.

19. Rietjens I. P., Safely Weld and Repair In-Service Pipelines, Pipeline Industry, December 1986,
pp26-29.

20. Considerations of Welding Methods Adopted on Pipelines During Operation, IIW Document XI-E-
477-87, 1987.

21. Hicks D. J. Guideline for Welding on Pressurised pipe, Pipeline & Gas Journal,
March 1983, pp17-19

22. Bout V.S. & Gretskii Yu.Ya., Arc Welding Application on Active Pipelines, Pipeline Technology,
Volume 1, R. Denys Ed. R. Denys, pub Elsevier Science BV. pp550-558.

23. Australian Standard – Pipelines – Gas & Liquid Petroleum AS 2885 1987 Clause 7.13.11.

24. Holman J.P., 1976, Heat Transfer, 7th edition pub. McGraw-Hill.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 25

gas flow hydrogen

gas pressure
electrode type
gas thermal properties

Pipe geometry weld HAZ Cracking


cooling
rate
hardness > 350

susceptible microstructure

weld heat input material composition

Figure 3.1
Factors influencing the possibility of hydrogen assisted cracking during in-service welding.

8
4.8 mm
7 6.4 mm
9.3 mm
6
T85C Cooling Time (secs)

15.1 mm

0
0 0.5 1 1.5 2 2.5 3 3.5 4
Gas Flow (mmscmd)

Figure 3.2
For a constant heat input of 0.9 kJ/mm this graph shows the reduction of t8/5 cooling time due to increased
gas flow in pipes of different wall thickness(3).
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 26

10
8 9 4
5 6 7
45 deg . 1 2 3
2t
t
(b) Buttering layers & temperbead

6
3 5
1 2 4
(a) Less than 1 mm

(c) Stringer bead arrangement

Figure 3.3
Recommended weld bead deposition sequences in order to make most benefit of tempering,
(a) general view of recommended joint configuration
(b) buttering layers and a temperbead at weld toe
(c) stringer bead arrangement.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 27

Arc current

Arc voltage

Heat Input Welding speed

Electrode polarity
Weld Penetration Electrode type

Pipe Wall Temperature Electrode diameter

Welder technique/direction
Local Pipe Wall Strength

Pipe wall thickness

BURN THROUGH Preheat temperature


Pipe Wall
RISK
Cooling Gas flow rate

Gas temperature

Pipe diameter
Local Gas pressure
Applied Stress
Yield strength at temperature

Weld orientation

Local wall support

Figure 3.4
Causal factors involved in burn-through during in-service welding
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 28

Not allowed without significant


pressure restriction

220
Burn-through Limits
200

Maximum Allowable Welding


4 mm diameter electrode
180
Current (Amps)
160
3.2 mm diameter electrode

140

burn-through
120
safe < 0.87 kJ/mm
100
3 4 5 6 7
Pipe Wall Thickness (mm)

Figure 3.5
Burn through limits expressed as limits to the allowable welding current, for both 3.2 mm and 4 mm
diameter electrodes from Cassie(17). Additional points from Bruce et al (11).

1.1

1
Maximum Heat Input (kJ/mm)

4 mm wall thickness
0.9

0.8

0.7
3.2 mm wall thickness
0.6

0.5

0.4

0.3

0.2
50 60 70 80 90 100 110
2.0 mm 2.4 mm 3.2 mm

Welding Current (Amps) / Electrode Diameter

Figure 3.6
Recommended limits for avoiding burn through during repair welding from Bruce et al(11)
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 29

Table 3.1 Conditions for burn-through from Bout & Gretskii(22)

Pipe Diameter Wall Thickness Internal Pressure Welding Direction Critical Heat Input for
(mm) (mm) (MPa) Burn-through
(kJ/mm)
320 3.0 4.0 Longitudinal 0.37
320 3.0 4.0 Circumferential 0.48
320 3.0 3.0 Longitudinal 0.48
320 3.0 3.0 Circumferential 0.51

14
Wade results, bulge height < 1.0 mm
Wade results, bulge height > 1.0 mm
12 Wade results, burst
Pipe Internal Pressure (MPa)

10

0
0.5 0.7 0.9 1.1 1.3 1.5 1.7 1.9
Welding Heat Input (kJ/mm)

Figure 3.7
Burn-through limits established by Wade (8) for welding onto a 300 mm diameter, X60 steel pipe of
5 mm wall thickness.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 30

1.8

1.6

1.4

Arc Energy (kJ/mm)


1.2 3mm thickness 6mm thickness
5mm thickness limit
limit
limit
1

0.8 4mm thickness


limit
0.6

0.4
welding not permitted
0.2

0
0 1 2 3 4 5 6
Gas Pressure (MPa)

Figure 3.8
Burn-through limits recommended by Wade(8) based on tests carried out on 300 mm diameter, X60 steel
pipe, using test pipes pressurised with non-flowing nitrogen.

Gap ao

Band to support weld zone

fitting

Figure 3.9
Schematic illustration of the joint configuration proposed by Bout & Gretskii(22) to eliminate burn-through
by supporting the pipe wall during in-service welding.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 31

4.8 mm thick pipe


2
1.8 Tewi = 10 sec
Tewi = 25.5 secs
1.6 Tewi = 20 sec
Heat Input (kJ/mm) Tewi = 14.4 secs
1.4 Tewi = 23.9 secs
1.2 Tewi = 30 sec
1
0.8
0.6 Tewi = 40 sec

0.4
0.2
0
0 2 4 6 8 10
T85C Cooling Time (secs)

Figure 3.10
A typical diagram derived by EWI(6) to enable the determination of required heat input
from a measurement of the pipes heat capacity.

100000
Effective heat transfer coefficient (W/m 2.K)

methane 'real' pipe


air flow low pressure
w ater flow 10000
Mobiltherm 603
Engine Oil
1000

100

10

1
0.01 0.1 1 10
Flow velocity (m/sec)

Figure 3.11
The estimated effective heat transfer coefficient at the inside of the pipe wall for different fluids and flow
conditions.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 32

15 T85 for 6.35 mm thick pipe


Water
Motor Oil
12.5
Air Flow
Data from Natural Gas Pipes

T85 cooling times (secs)


10

7.5

2.5

0
0 0.5 1 1.5 2 2.5
Heat Input (kJ/mm)

Figure 3.12
The relationships between t8/5 cooling time and the weld heat input on 6.35 mm pipe containing different
fluids, compared with points for gas flow. Data replotted from (3).

25 T85 data for 4.8 mm thick pipe


Water
Motor Oil
Air flow
20
Data from natural gas pipes
T85 cooling times(secs)

15

10

0
0 0.5 1 1.5 2 2.5
Heat Input (kJ/mm)

Figure 3.13
The relationships between t8/5 cooling time and the weld heat input on 4.8 mm pipe containing different
fluids compared with points for gas flow. Data replotted from (3).
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 33

4. Literature Review:
Numerical Simulation of Fusion Welding
4.1 Introduction
Using a computer simulation of a fusion welding process to calculate thermal effects is a well-
established technology. Simulation can be used to estimate weld penetration, the heat affected
zone size and the cooling rates, based primarily on the welding process parameters of arc current,
voltage and speed. In this way simulations can be used to determine an acceptable weld
procedure, and the normal ‘experimental’ trial and error approach is replaced with trial and error
computer simulation.

The core techniques which are used to create a weld model are available in commercial software;
these are, heat transfer, computational fluid dynamics and stress analysis. However, using these
tools to create an accurate weld simulation, is still a significant technical challenge.

Weld simulations are essentially solutions of heat-transfer equations. To fully model all of the
physical phenomena occurring within a fusion welding process would be too complex and too time
consuming to generate a industrially attractive ‘useful’ outcome. It is essential to reduce this
complexity by making some approximations.

Models have historically been simplified by:


• modelling the workpiece only, and representing the welding arc and the transferred molten
droplets of metal as a distributed heat flux,
• not modelling the complex metal flow within the molten weld pool, so the solution is achieved
by only considering conductive heat transfer,
• solving the model under idealised conditions which reduce the degrees of freedom, for
example using quasi-steady state solutions rather than transient ones, or 2D solutions in
place of 3D ones,
• using material properties which do not vary with temperature, giving a linear problem.

Calculations of this type have been applied to welding since Rosenthal(1) first applied analytic
solutions of the heat transfer equation to welds in 1930, and gradually the constraining
approximations have been removed. Contemporary models use direct numerical methods of
solving the heat conduction equation. Non-linear, temperature dependent conductivity and thermal
diffusivity are used and the latent heat at phase changes allowed for. Distributed heat sources are
used to represent the welding arc. Radiative and convective heat losses can be included.

Early numerical solutions by Westerby(2), Pavelic et al(3) and Kou & Le(4), used the finite difference
approach, while Hibbett & Marcel(5), Krutz & Segerlind(6), Friedman(7) and Paley & Hibbett(8)
pioneered the use of the finite element method (FEM). In general, most work now uses the finite
element approach since this method is more flexible in its treatment of variable part geometry. In
such methods the governing equation of heat transfer, becomes:

∂2T ∂2T ∂2T dT


k 2 + k 2 + k 2 + Q = ρC
∂x ∂y ∂z dt

with boundary conditions represented as,

∂T
k + h(T - T o ) + σε ( T 4 - T 4o ) = 0
∂n
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 34

Finite element discretisation leads to an equation of the form;

∂T
KP(T) + PM( )=Q
∂t
This set of first order differential equations is then solved. The simplest methods of solution discretise
the time differential by finite differences and integrate, through time, using a fixed time step.

In the period 1975-1985, solutions were limited to 2D approximations by lack of computer


capabilities. The models developed by Friedman et al(9,10), and applied to bead-on-plate gas
tungsten arc (GTA) welding were typical. The 2D approximation is based on the assumption that the
welding speed is so high that effectively no heat is conducted ahead of the welding arc. Accordingly,
the heat flow through a planar 2D slice is considered to be negligible in comparison to the heat flow
within the plane. The calculation then determines the temperature at each point within the plane as a
time dependent heat source representing the welding arc acts on it, (Figure 4.1). A quasi-stationary
condition is assumed so the temperature at any plane in the plate (say, xn,y,z,t) is simply determined
from a displacement in the time scales. i.e.

xn
T( x n , y, z,t) = T(0, y, z,t - )
v

Rectangular 4-noded elements were used (9) and these were decreased in size towards the centre
of the welding arc. Heat losses by radiation and convection were considered and temperature
dependent physical properties were used. Following the earlier work of Pavelic(11), the welding arc
was represented as a gaussian distributed surface heat flux. The calculated results were compared
with experimentally measured temperature cycles determined from thermocouples embedded within
the workpiece. Calculated values of weld pool size were found to be sensitive to the arc efficiency,
the arc radius, and to a lesser extent, on the coefficient of convective heat loss. The arc radius was
linearly increased in proportion to the arc current, consistent with observed GTA arcs (9).

The major concern in applying these numerical methods was the uncertainty in knowing which heat
source, heat distribution factor or transfer efficiency to use. Applying the arc energy as a surface
distribution tended to produce predicted weld pools with a half cylindrical section, but in practice a
wide range of pool shapes are obtained depending on welding conditions and process parameters.
Westby(2) was the first to adopt the idea that heat applied to the weld pool was not only distributed
spatially on the surface but also within the pool volume by convection. It was recognised that the
force from the arc plasma and the convection of heat by fluid flow within the weld pool could generate
an 'arc-digging' effect which could distribute the arc energy within the volume of the material,
Mills(12) and Essers & Walter(13). The 'disc' surface heat source representation could be
appropriate for some heat sources which do not cause melting, for example a pre-heat torch, but are
too restrictive to simulate the welding arc.

In their numerical modelling work, Battelle(14) adopted a similar 2D strategy to that of Friedman(9). In
their case the heat input in the weld zone was uniformly distributed in the weld elements or nodes.
Its variation with time as the heat source ‘passed’ through the 2D section was chosen in order to get
good agreement between predicted and experimentally measured t8/5 cooling times.

The concept of using a heat source description that correctly reflects the character of the heat
distribution in a particular welding process has become the cornerstone of good thermal models. In a
series of publications Goldak and his co-workers (15,16,17,18), proposed a more general description
of an arc heat source which they termed the 'double ellipsoidal' source. In the initial case the heat flux
was distributed in a gaussian manner throughout a semi-ellipsoidal volume. Their experience with
this source showed that the predicted temperature gradients in front of the arc were less steep than
experimental observed ones, and gradients behind the arc were steeper than those measured. To
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 35

overcome this difficulty, two ellipsoids were combined to give the final 'double ellipsoidal' source as
illustrated in Figure 4.2. This source was described by six parameters, η the arc efficiency, such that
Q=η.VI, ellipsoidal axis a, b, cf , cb, the depth, width and length in front and behind the arc centre. In
addition an apportionment of heat to the front or rear section of the source is made by factors, rf and
rb where, rf + rb= 2. The heat flux at a point x,y,z within the ellipsoid, in front of the arc centre, is
then given by the following equation,

6 3 r f ηVI - 3 x2 3 y2 3 z 2
q(x, y, z) = exp [ - 2 - 2 ]
ab c f π π cf
2
a b

and a similar equation with the substitution of cb, and rb applies to points within the rear section of the
ellipsoid.

Goldak et al(15) argued that this formulation provided a volume heat source which reflected the
concept that the weld pool motion would decrease towards the boundary of the weld pool.
Additionally the gradual reduction in heat flux avoided a sharp step change in heat load and avoided
the need to use exceptionally small finite element mesh. Goldak et al(16) implied an equivalence
between the source dimensions and the weld pool, and suggested that appropriate values for a,b,cf
& cb could be arrived at by direct measurement of a weld. It was suggested that reasonable values
were, a, the width of weld, b the depth, and cf=a, cb=2a. If values were not available then
estimations from Christensens(19) non-dimensional method could be used. The credibility of the
source was demonstrated by using it in a 2D FEM analysis of a submerged arc weld and an electron
beam weld.

Mahin et al(20) applied FEM to both static, axisymmetric and 3D moving GTA welds on 304 stainless
steel. A gaussian heat source was used with an arc radius of 2 mm, in line with measured heat flux
distributions. Arc efficiency was measured by calorimetry to be 0.75. In order to simulate the more
rapid transfer of heat due to fluid convection the thermal conductivity of the liquid (T > Tm) was
increased in a linear manner such that at 3000K it was 10 times greater than actual. They also
introduced the idea of using an anisotropic thermal conductivity to control the calculated weld pool
shape. With isotropic thermal conductivity calculated weld pools had depth to width ratios ≤ 0.5, by
enhancing the conductivity in the thickness direction relative to that in the plate the pool depth could
be preferentially increased. Weckman, Mallory & Kerr(21) used essentially the same approach, and
obtained good agreement with experimental pool shapes in GTAW.

Tekriwal & Mazumber (22,23,24), used a 3D commercial code (ABAQUS) to model single(22) and
multiple(24) pass gas metal arc (GMA) butt-welds. Temperature cycles were found to be close to
experimental with ones determined from embedded thermocouples and predicted HAZ and fusion
zone widths were within 5-15% of the experimental ones. In their models, metal deposition was
simulated by the periodic addition of mesh and a radially symmetric gaussian heat source was used.

Pardo & Weckman(25) developed a 3D model for bead on plate GMA welds with some novel
features. A gaussian heat source was used but the thermal conductivity of the liquid material was
artificially increased to twice the real value to reflect the increased heat transfer engendered by
convective heat flow. The area of the weld reinforcement was determined from the known wire feed
rate, wire diameter and welding speed. A parabolic weld profile was assumed. An initial calculation
was carried out using an estimated bead width and then compared with the predicted width of the
melting point isotherm. The initial bead width was then adjusted and further calculations carried out
iteratively until the bead width and pool width converged. Agreement with experiment was good,
although it was pointed out with some welding conditions a characteristic 'finger' penetration was
obtained and the model could not account for this situation. In discussing similar problems
Easterling(26), points out that models using gaussian distributions predict smooth cylindrical
penetrations which are only found with a limited number of weld conditions.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 36

It is clear that the major failing of early heat transfer models is that they inherently consider the heat
input as a total energy. They did not directly recognise that the individual components of the arc
energy can effect process outcomes, or that other process parameters, such as shielding gas type,
can have an influence. Such effects have been extensively studied, Chakravarti et al(27), Dillenbeck
& Castagno(28) and Chandel(29). A clear result of such effects is that different cooling times, fusion
zone and HAZ geometries can be obtained by changing process parameters such as voltage, current
and speed, even though the net heat input is kept constant. Such effects are accounted for in
models by the manipulation of variable factors, often those defining the heat source distribution. In
much of the published work, good agreement between experiment and model was reported but this
was often a 'forced' agreement brought about by the appropriate choice of 'tuning'. In a review,
Goldak(18) pointed out this basic difficulty stating that "a cynic may argue that these models require
that the solution is known before the solution can be computed".

The following factors appear in most heat transfer models and for an accurate simulation they must
be chosen with careful regard to the specific welding process:

(a) Arc efficiency


Arc efficiency is generally taken to be the ratio of heat energy used within a weldment to the heat
equivalent of the total power output. For manual metal arc (MMA) welding values are quoted in the
range 0.70-0.80(30). However, there have been different interpretations of this parameter, which
confuse these values. There have been two approaches:

1. Indirect methods where the efficiency values are the correction factors necessary to obtain a
‘best fit’ between thermal histories from models and experimental measurement.
2. Direct measurement by calorimetry.

Giedt et al(31) considered that this resulted in two classes of 'arc efficiency'. In most FEM models
the arc efficiency used is of the 'best-fit' kind.

(b) Heat source distribution


The fundamental problem in applying conduction models to welding is the arbitrary choice and
formulation of the heat source through factors that are not known apriori. Despite this difficulty
considerable effort has been made to improve such models. Gaussian distributions are favoured,
usually justified by the evidence that current distributions in arcs have been measured to have similar
distribution. This may not be completely acceptable however, since Choo et al(32) in modelling GTA
weldpools with distorted surfaces has predicted bimodal current and heat flux distributions. Surface
gaussian distributions are most favoured for modelling GTA welding and in those cases the arc
radius is often chosen to proportionally increase with arc current. Double ellipsoidal sources have
been generally related to the actual size of the weldpool, as originally proposed by Goldak et al(16)
but in subsequent work Goldak et al(33) used a completely variable distribution. Kamala &
Goldak(34) introduced the term 'power density distribution function', (pddf) which is a variable power
distribution equivalent to the welding arc.

Smailes(35) has presented an excellent review of the various heat source definitions that can be
used and the reasons for their derivation. The majority of work has concentrated on GTA and GMA
welding processes. Weld pool flow and weld pool surface distortion are particularly significant in GMA
welding, since they are driven by the incoming stream of impacting droplets. This effect has been
addressed by splitting the heat input into two parts, Allum & Quinto(36), Painter et al(35), Tsao &
Wu(38) and Kumar & Bhaduri(39).

c) Liquid thermal conductivity


To simulate the effect of convective heat transfer in a conductive model the thermal conductivity of
the molten material is artificially increased. There is no set, prescribed method, however and the
level of increase has been varied. Pseudo-conductivity values can be fixed at values of between 2-6
times the actual liquid conductivity, or they can be increased in proportion to temperature above the
melting point, up to maximum values of 10 times the actual conductivity at the metal’s boiling point.
The introduction of directional, anisotropic values of conductivity has also been used to simulate an
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 37

assumed directional convective flow.

(d) Joint configuration and metal deposition


Finite element models can be freely applied to any geometry. However most work to date has
focussed on relatively simple parts, bead-on-plate or bead-in-groove style, single pass welds. Metal
deposition, where it has been included, has generally been tackled by dynamic meshing, i.e. creating
additional mesh during the solution, or by a birth-and-death scheme, activating pre-meshed elements
during the solution.

4. 3 Current status of weld Simulation


Even though there are some problems, numerical solutions based on conductive heat transfer are in
practical use. Various strategies to minimise total mesh size are generally necessary to increase the
computational efficiency, particularly in 3D solutions. This is desirable because to maintain solution
accuracy and numerical stability a fine finite element mesh must be used near the welding arc.
However, using a fine mesh along the full length of a weld path has computational penalties.

An adaptive meshing and remeshing scheme can be used (Kannatey-Asibu et al(40)) or a fast
analytic solution can be used to determine the boundary conditions on a restricted domain
surrounding the welding arc. These boundary conditions are then applied to a numerical solution
within the restricted zone, Kumar et al(41).

Accepting that there are some computational strategies to be developed to improve efficiency, the
major commercial FE codes, ANSYS, NASTRAN, ABAQUS etc, can currently address non-linear
heat transfer problems and can therefore be applied to the calculation of temperatures during
welding. There are also codes such as SYSWELD specifically produced to address weld simulation.
Current examples concern relatively simple structures, butt-joints, plates and pipes. Leung &
Pick(42), and Leung, Pick & Mok(43) have applied ABAQUS code to both single and multi-pass
semi-automatic submerged arc welding. Brown & Song(44,45) have also applied the ABAQUS code
to the determination of temperatures and residual stresses during pipe welding. The determination of
residual stress, although again computationally demanding is relatively straight forward if the
mechanical properties of the material can be considered to be temperature dependent but
independent of microstructure. Then the thermal calculation can be de-coupled from the stress
calculation and the calculation becomes a two-part procedure. Firstly a calculation of the thermal
history during welding and cooling is made. This is followed by a stress calculation using the thermal
data as a loading. The uncertainty in choice of heat source model is also of less concern here, since
the heat source distribution tends to influence the near-weld temperatures to a greater extent than
those in the bulk of the material. However, significant residual stresses are only generated in steels at
temperatures below 600°C, and are therefore more dependent on temperature changes away from
the immediate weld site.

Temperatures some distance from the weld can be accurately calculated from such models.
However, for the calculation of near-weld temperatures the status of conduction models is less
satisfactory. In this case the choice of heat source distribution is more problematic and the resulting
temperatures are more sensitive to that choice. Goldak et al(46) applied such numerical approaches
to in-service welds and pointed out that the detailed definition of the welding heat source, its
orientation and positioning relative to the weld bead were important in such models, and could lead to
inaccuracies in the calculation of near-weld temperatures. This observation has important
ramifications for our approach to the simulation of in-service welding, for in this case the accuracy of
the near weld thermal cycle is important. It has a bearing on the near weld thermal cycle and the
estimation of heat affected zone hardness, and on the estimation of local pipe wall strength and the
assessment of burn-through risk.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 38

References
1. Rosenthal, D., Mathematical Theory of Heat Distribution During Welding and Cutting,
Welding Journal Research Supplement v 20, n5, 1941, pp220s-234s.

2. Westby 0., Temperature Distribution in the Workpiece by Welding,


Department of Metallurgy and Metal Working, Technical University Trondheim, Norway 1968.

3. Pavelic V., Weld Puddle Shape and Size Correlation in Metal Plates Welded by GTA Process,
Proceedings: Arc Physics and Weld Pool Behaviour Conference, Welding Institute, London, May
1971, pp.251-258.

4. Kou S. & Le Y., 3D Heat Flow and Solidification During the Autogenous GTA Welding of
Aluminium Plates, Metallurgical Trans A. v14A, Nov. 1983, pp2245-2253.

5. Hibbet H. & Marcel P., A Numerical Thermomechanical Model for the Welding and Subsequent
Loading of a Fabricated Structure, Computers and Structures v3, 1973, pp1145-1174.

6. Krutz G.W. & Segerlind L.J., Finite Element Analysis of Welded Structures,
Welding Journal v57, n7, 1978, pp211s-216s.

7. Friedman E., Thermomechanical Analysis of the Welding Process Using the Finite Element,
Method, Trans. ASME J Pressure Vessel Technology v97, n3. Aug. 1975 pp206-213.

8. Paley Z. & Hibbet P., Computation of Temperature in Actual Weld Designs,


Welding Journal, v54, n10, Nov. 1975, pp385s-392s

9. Friedman E., On the Calculation of Temperature Due to Arc Welding,


Computer Simulation for Materials Application: Nuclear Metallurgy v20 n2 April 1976
pp1160-1170.

10. Friedman E. & Glickstein S. S., Effect of Weld Pool Configuration on Heat Affected Zone Shape
Welding Journal, v60 n6 June 1981, pp.110s-112s.

11. Pavelic R., Tanbakuchi R., Ujehara O. & Myers P., Experimental and Computed Thermal
Histories in GTA welding of Thin Plates, Welding Journal, v48, n7, July 1969, pp295s-305s.

12. Mills G.S., Fundamental Mechanisms of Penetration in GTA Welding,


Welding Journal, v58, n1 Jan 1979 pp.21s-24s.

13. Essers W.G. & Walter R., Heat Transfer and Penetration Mechanisms with GMA & Plasma-GMA
Welding, Welding Journal, v60, n2, Feb. 1981, pp37s-42s.

14. Kiefner J.F & Fischer R.D., Models Aid Repair Welding Procedure,
Oil Gas Journal, March 1998, pp41-47.

15. Goldak J. A., Chakravarti A. & Bibby M.J., A Double Ellipsoid Finite Element Model for Welding
Heat Sources, IIW Doc 212-603-85, Jan. 1985.

16. Goldak J.A., Chakravarti A. & Bibby M.J., A New Finite Element Model for Welding Heat
Sources, Metallurgical Trans B, v15B, June 1984 pp299-305.

17. Goldak J. A., McDill M., Oddy A., House R., Chi X. & Bibby M. J., Computational Heat Transfer
for Weld Mechanics, Proceedings: Advances in Welding Science & Technology, TWR'86, ed.
S.A.David, pub. ASM International , May 1986, pp15-21.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 39

18. Goldak J., Keynote Address: Modelling Thermal Stresses and Distortions in Welds,
Proceedings: Recent Trends in Welding Science and Technology TWR'89, ed. S. A. David & J.
M. Vitek, May 1989, pp71-83.

19. Christensen N., Davies V de L. & Gjermudsen K., The Distribution of Temperature in Arc
Welding, British Welding Journal v12, n2, Feb. 1965, pp54-75.

20. Mahin K.W., Shapiro A.B. & Hallquist J., Assessment of Boundary Condition Limitations on the
Development of a General Computer Model for Fusion Welding, Proceedings: Advances in
Welding Science and Technology. ed S.A.David, pub. ASM International, pp.215-223.

21. Weckman D.C., Mallory L.C. & Kerr H.W., A Technique for the Prediction of Average weld Pool
Width and Depth for GTA Welds On Stainless Steels, Zeitschrift fur Metallkunde v80, n7, July
1989, pp459-468.

22. Tekriwal P. & Mazumder J., Effect of Torch Angle and Shielding Gas Flow on TIG Welding - A
Mathematical Model, Met. Construction June 1988 v20, n6, pp.275R-279R.

23. Tekriwal P. & Mazumder J., Finite Element Modelling of Arc Welding Processes,
Proceedings: Advances in Welding Science and Technology. ed S.A.David, pub. ASM
International, pp.71-80.

24. Tekriwal P. & Mazumder J., 3D Finite Element Analysis of Multi-pass GMAW,
Proceedings; Modelling and Control of Castings and Welding Processes IV, ed. A.F.Giamei &
G.J.Abbaschian, pub. Minerals, Metals & Materials Soc. May 1988, pp167-177.

25. Pardo E. & Weckman D., A Numerical Model of the GMA Welding Process
Proceedings; Modelling and Control of Castings and Welding Processes IV, ed. A.F.Giamei &
G.J.Abbaschian, pub. Minerals, Metals & Materials Soc. May 1988, pp187-197

26. Easterling K. E., Keynote address: Predicting Heat Affected Zone Microstructures and Properties
in Fusion Welds, Proceedings: Recent Trends in Welding Science & Technology, May 1989, ed.
S.A.David & J.M.Vitek, pub. ASM International 1990, pp177-189.

27. Chakravarti A. P., Thibau R. & Bala S. R., Cooling Characteristics of Bead-on-plate Welds
Metal Construction, v17, n3, 1985 pp178R-183R.

28. Dillenbeck V. R. & Castagno L., The Effects of Various Shielding Gases and Associated
Mixtures, in GMA Welding of Mild Steels, Welding Journal v66, n9, 1987 pp45-49.

29. Chandel R.S., Oddy A.S. & Goldak J.A., Computer Prediction of Weld Bead Shapes,
Paper presented at Computer Technology in Welding, Conference June 1988

30. Lancaster J.F.


The Physics of Welding,
Pergamon Press 1986.

31. Giedt W.H., Tallerico L.N. & Fuerschbach P.W., GTA Welding Efficiency: Calorimetric &
Temperature Field Measurement, Welding Journal v68, n1, 1989, pp28s-32s.

32. Choo R.T.C, Szekely J. & Westhoff R.C, Modelling of High Current Arcs with Emphasis on Free
Surface Phenomena in the Weld Pool, Welding Journal Research Supplement, v69, n9, 1990,
pp346s-361s.

33. Goldak J.,A, Bibby M.J., Moore J., House R. & Patel B., Computer Modelling of Heat Flow in
Welds, Trans AIME vol17B, September 1986, pp587-600.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 40

34. Kamala V. & Goldak J.,A., Error due to Two Dimensional Approximation in Heat Transfer
Analysis of Welds, Welding Journal, v72, n9, September 1993, pp440s-446s.

35. Smailes A., Thermal Modelling of Gas Metal Arc Welding Using Finite Element Analysis, MSc
Thesis, The University of Adelaide, South Australia, January 1999.

36. Allum C.J. & Quinto L., Control of Fusion Characteristics in Pulsed MIG Welding: Part 1,
Metal Construction, v17, April 1985, pp242R-245R.

37. Wahab M. A, Painter M. J. & Davies M. H.,


The Prediction of Temperature Distribution and Weld Pool Geometry in the Gas Metal Arc
Welding Process, Journal of Materials Processing Technology, vol. 77, no 1-3, pp233-239.

38. Tsao K.,C. & Wu C.,S., Fluid Flow and Heat Transfer in GMA Weld Pools, Welding Journal, v67,
n3, March 1988, pp70s-75s.

39. Kumar S. & Bhaduri S.C., Three Dimensional Finite Element Modelling of Gas Metal Arc
Welding, Metallurgical and Materials Transactions, B, v25B, June 1994, pp435-441.

40. Kannatey-Asibu E., Kikuchi N., Jallad A.R., Experimental Finite Element Analysis of Temperature
Distribution During Arc Welding, Trans. ASME. Journal of Engineering Materials Technology,
v111, n1, Jan. 1989 pp.9-18.

41. Kumar B.V., Mohanty O.N. & Biswas A., Welding of Thin Steel Plates: A New model for Thermal
Analysis, Journal of Materials Science 27, 1992, pp203-209.

42. Leung C.K., Pick R.J. & Mok D.H.B., Finite Element Modelling of a Single Pass Weld,
WRC Bulletin 356, August 1990.

43. Leung C.K. & Pick R.J., Finite Element Analysis of Multi-Pass Welds,
WRC Bulletin 356, August 1990.

44. Brown S. & Song H., Implications of Three Dimensional Numerical Simulations of Welding of
Large Structures, Welding Journal Research Supplement, v71, n2, 1992, pp55s-62s.

45. Brown S. & Song H., Finite Element Simulation of Welding of Large Structures,
Journal of Engineering for Industry, v114, November 1992 pp441-451.

46. Goldak J.A., Oddy A.S., & Dorling D.V., Finite Element Analysis of Welding On Fluid Filled
Pressurised Pipelines, Published in International Trends in Welding Science and Technology,
Proceedings of International Conference, Trends in Welding Research, Gatlinburgh,
ed. S. A. David & J. M. Vitek, 1993.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 41

Surface heat source to represent z


welding arc
y

welding velocity Vx Graded finite element mesh on


x the 2D analysed section

plate

symmetry

Figure 4.1
An illustration of the system for 2D finite element modelling of a welding process as developed by
Friedman (9).

Gaussian distribution in x

y
Gaussian distribution in y

x z

a
cb

cf b
Gaussian distribution in z

Figure 4.2
Schematic representation of the ‘double ellipsoidal’ heat source developed by Goldak et al. (15).
Heat flux is distributed in a Gaussian manner in the x, y and z directions within an ellipsoidal volume
defines by lengths, a, b, cf and cb.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 42

5. Appraisal of the Battelle Computer Program


The Battelle Computer program(1,2) is well known within the pipeline industry, but it does not seem to
have supplanted the use of the EWI procedure(3). Experimental evaluation of welding conditions still
appears to be the main method of establishing in-service welding procedures (see report on the
Industry Survey-Section 6).

The Battelle program uses a 2D approximation to the main types of in-service weld. This simplification
can be invalidated by changing joint geometry or by slow welding speeds, which infringe the basis of
the 2D approximation. It is reasonable for circumferential fillet welds, but may be less satisfactory for
direct-branch welds, where welding angles and geometry vary around the weld. This is unlikely to be
a major problem however, since direct branch welds are less common than circumferential fillet welds.
In addition, model sections taken at the top dead centre, and the side of a branch connection (2)
effectively ‘bracket’ the variation in conditions around the weld.

To account for the cooling effect of the flowing gas the Battelle program utilises the Sieder & Tate non-
dimensional relationship.
0.14
hd ⎛μ ⎞
= 0.023.(Re) − 0.2 .(Pr) − 2 / 3 ⎜⎜ b ⎟⎟
C p .V .ρ ⎝ μw ⎠
Where:
hd is the effective convective heat transfer coefficient at the pipe wall,
Cp is the specific heat of methane,
V is the velocity,
Re the Reynolds Number,
Pr the Prandtl Number,
ρ the density,
μb , μ w are the viscosity at the wall temperature and in the bulk of the gas.
This equation relates the effective convective heat transfer coefficient at the pipe wall to the physical
properties of the gas and the current flow, pressure and pipe diameter. It is generally accepted that
this approach is a reliable method of estimating heat losses, with an accuracy of ± 25% (Holman(4)).

An extensive library of fluid data is included within the Battelle program.

The Battelle calculation is a conductive heat transfer calculation, and in keeping with such an
approach, it uses a distributed heat flux to represent the welding arc. The accuracy of near-weld
temperatures is largely dependent on the way in which the welding energy is distributed. Details of
how the heat is represented in the Battelle model are not available. Kiefner & Fischer(1) describe it as
a time dependent heat input, in which the time dependence was chosen to suit measured cooling data
from bead-on-plate tests. This may mean that the heat within the weld bead was assumed to be
uniform. Such decisions can have an influence on the predicted weld penetration, and for thin pipe
walls, on the inside wall temperature.

It is normal practice to vary the heat source distribution to reflect the different modes of heat
distribution achieved with different welding consumables. Kiefner & Fischer(1) acknowledge this,
indicating that although their models do not vary the heat source for different consumables, they are
appropriate for low-hydrogen electrodes.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 43

The predicted cooling rates from the Battelle program were extensively validated using different pipe
geometry and wall thickness. As shown diagrammatically in Figure 5.1 the test range of pipe wall
thickness was focussed on 6.35 mm, and did not go below 4.8 mm. A typical comparison between
measured and predicted t8/5 values is shown in Figure 5.2.

Bruce(5) has reported that the calculated cooling times for thin walled pipes tended to be higher than
the measured values. This was particularly noted under flow conditions giving cooling times between
4-5 seconds. Because such a variation could lead to choosing a heat input that was lower than
desirable, Bruce(5) noted that this was non-conservative. Since this is also typical of the operating
conditions for many pipelines this observation is a concern. Kiefner & Fischer(1) state that the Battelle
program is suitable for pipe wall thickness between 3.2 – 9.6 mm., and not appropriate for wall
thickness greater than 12.5 mm.

Care must be taken with respect to the representation of carbon equivalent and hardness within the
Battelle program. This uses the IIW formulae, which does not truly represent the hardening capacity of
modern high strength steels. Kiefner & Fischer(1) recommend that the results of the Battelle program
should only be used for pipe grades up-to and including X52 steel. However, this observation has no
bearing on the accuracy of predicted cooling times.

Battelle’s in-service welding model is primarily used for predicting weld cooling rate and heat affected
zone hardness, but it can be used to assess the risk of burn-through. The approach is based on an
experimental observation that burn-through only occurred when the maximum inside, pipe wall
temperature exceeded 980°C. The level of 980°C was considered to be appropriate for low-hydrogen
electrodes(1). Wall temperature is important, but as indicated in section 3, the wall temperature is only
part of the picture, and pressure is another important variable. The burn-through limits for repair welds
established by Bruce et al(6) were broadly in line with the predictions of the Battelle program but they
considered that additional work was required to validate the model predictions.

Goldak et al (7) when using more complex 3D numerical models pointed out that the accuracy of the
inside wall temperature was very dependent on the way the welding arc was represented in the model.

5.3 Summary
The Battelle model has made an important contribution to in-service welding technology. However,
there are some concerns with respect to its potential accuracy on thin walled pipes. In addition the
burn-through criterion, which relies on the calculated inside wall temperature of the pipe, fails to
account for internal pressure and may also be liable to some inaccuracy with electrodes of different
penetration characteristics.

5.4 References
1. Kiefner J. F & Fischer R. D. Models Aid Pipeline Repair Welding Procedure,
Oil & Gas Journal March 1988, pp41-47.

2. Fischer R. D., Kiefner J.F. & Whitacre G.R., User Manual for Model1 & Model 2 Computer
Programs for the Predicting Critical Cooling Rates and Temperatures During Repair and Hot Tap
Welding on Pressurised Pipelines, Battelle Memorial Institute Report, June 1981.

3. Bruce W. A. & Threadgill P. L. Welding Onto In-Service Pipelines,


Welding Design & Fabrication Feb 1991, pp19-24.

4. Holman J. P., 1976, Heat Transfer, 7th edition pub. McGraw-Hill.

5. Bruce W. A. & Threadgill P. L., Effect of Procedure Qualification Variables for Welding Onto In-
service Pipelines, American Gas Association Report J7141, July 1994.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 44

6. Bruce W. A., Holdren R.L., Mohr W. C., Kiefner J.F. & Swatzel J.F., Repair of Pipelines by Weld
Metal Deposition, Paper presented at PRCI 9th Symposium on Pipeline Research, Houston Texas,
September 1996.

7. Goldak J. A., Oddy, A.S. & Dorling D.V., Finite Element Analysis of Welding On Fluid Filled
Pressurised Pipelines, published in International Trends in Welding Science And Technology,
Proceedings of International conference, Trends in Welding Research, Gatlinburgh, ed S.A.David
& J.M.Vitek, 1993.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 45

80

70
N umber of validation w elds carried out

60

50

40

30

20

10

0
4.88 5.58 6.35 7.9 9.2 10.9 12
Pipe Wall Thickness (mm)

Figure 5.1

An indication of the number of test welds and the pipe wall thicknesses used to validate the Battelle
Program(5).
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 46

12
4.8 mm wall thickness

Predicted t 8/5 Cooling Times


10

(secs)
6

0
0 2 4 6 8 10 12
Measured t 8/5 Cooling Times (secs)

18
6.4 mm wall thickness
16
Predicted t 8/5 Cooling Times

14
12
10
(secs)

8
6
4
2
0
0 2 4 6 8 10 12 14 16 18
Measured t 8/5 Cooling Time (secs)

18
9.3-9.5 mm wall thickness
16
Predicted t 8/5 Cooling Times

14
12
10
(secs)

8
6
4
2
0
0 2 4 6 8 10 12 14 16 18
Measured t 8/5 Cooling Time (secs)

Figure 5.2

Example of the accuracy achieved with the BATTELLE Program (5) on pipes of different wall
thickness.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 47

6. Welding on In-Service Pipelines: Industry Survey

6.1 Introduction
A questionnaire was constructed and circulated to 25 companies to determine current practice
within the Australian pipeline industry with respect to in-service welding.

The survey sought to:

• determine the current range of pipe grades and pipe geometry within the Australian pipe
network,
• determine the typical operating condition for gas flow and pressure,
• determine the preferred hot-tap fittings used,
• determine typical welding consumables and joint types,
• assess the current preferences with respect to weld procedure development,
• evaluate the awareness and usage of the Battelle in-service welding software(1), and the
EWI test procedure(2).

In particular a proposed standard for the EWI test (see Appendix 6.1) was circulated to get
comments on possible problems with this procedure and the need for a standardised approach.

The response to the survey was poor and only 5 (20%) were returned. However, the information
was comprehensive and gave a good picture of the status of in-service welding techniques. The
results of the survey are not covered in detail here, but some selected extracts from the returned
questionnaires are presented in Table 6.1- 6.6.

6.2 Observations
Table 6.1 gives the range of typical pipe materials and operating conditions for Australian gas
pipelines. Steel grades range from X42-X65, with a significant spread of carbon equivalents,
from 0.28-0.5. Pipe wall thickness ranges from 3 – 17 mm, but common values are from 4.2-
6.35. The maximum allowable operating pressure (MAOP) was 15 MPa, but a typical value is 7
MPa, with working pressures ranging from 2-7 MPa. Maximum flow rates are up to 20 m/s, but
typical values are less than 10 m/sec.

As Table 6.2 shows that MMA welding is the dominant process, with low hydrogen electrodes
being used in both the vertical-up and vertical-down welding positions. Pre-heat is always used,
typically in the 70-100°C range. This is achieved by LPG heating.

As shown in Table 6.3, the full encirclement sleeve is the preferred hot-tap fitting although some
direct-welded outlets are used. The least popular joint form is a branch and compensating
saddle.

Only one of the five respondents did not use the Battelle program to analysis in-service welding,
as indicated in Table 6.4, although there was one indication that it was only used occasionally.
There were some concerns that the software was not user-friendly, and may have some
limitations. The EWI test was used by three of the five respondents, see Table 6.5. There was
general agreement that some form of standardised approach was desirable. There was some
concern that the method of heating (oxy-LPG torch) could produce a variable maximum
temperature which could lead to inconsistency in results.

Table 6.6 indicates the current limitations that are placed on in-service welding. Limitations on
pipe wall thickness, which restrict welding to pipes thicker than 4.8 mm are typical of international
standards.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 48

References
1. Fischer R. D., Kiefner J.F. & Whitacre G.R., User Manual for Model1 & Model 2 Computer
Programs for the Predicting Critical Cooling Rates and Temperatures During Repair and Hot
Tap Welding on Pressurised Pipelines, Battelle Memorial Institute Report, June 1981.

2. Bruce W. A. & Threadgill P. L. Welding Onto In-Service Pipelines,


Welding Design & Fabrication Feb 1991, pp19-24.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 49

Table 6.1
Typical pipe materials and operating conditions for the existing Australian Pipe network.

Steel Grade Audit

Maximum Minimum Maximum


Steel Grade CEQ Diameter Thickness gas working working Flow range
pressure pressure pressure
(mm) (mm) (MPa) (MPa) (MPa) (m/sec)
X42 0.35 323 5.2 4.6 4.3
X42 100 3 10 4.3 7 0.2-2.0
X42 150-200 4.2-4.8 7.3 4.3 7.3 3.6-7.0
X42 0.36 762 9.5 2.8 1.8 2.8 19.7
X42 0.28-.50 50-864 3.2-17 MAOP 0.7*MAOP 3.0-5.0
X42 0.41 219 5.6 7.4 4 6.5 0.3-10.6

X46 0.41 273 4.8-6.35 7.1 7.1

X52 350 4.7-6.0 7.45 2.5 6.3


X52 550 7.9-9.5 7.3 2.2 7.3 2.8-8.5
X52 0.28-0.50 50-864 3.2-17 MAOP 0.7*MAOP 3.0-5.0

X60 0.36 355 6.4 7.4 2 5.5 2.0-10.5


X60 0.31 406 6.4-7.7 9.6 7.1
X60 500 8.9-9.5 7.3 2.2 6.2 1.4-4.0
X60 0.45 762 10.3 7 3 6.7 1.5-19.6
X60 0.28-.50 50-864 3.2-17 MAOP 0.7*MAOP 3.0-5.0

X65 400 9.5 15 10


X65 0.28-.50 50-864 3.2-17 MAOP 0.7*MAOP 3.0-5.0

ASTM106 50-400 4.0-12.0 10 0.15 10


CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 50

Table 6.2
Typical welding processes / consumables / and conditions used for Australian in-service welding
procedures: Circumferential fillet welds

Welding Process / Electrode type

Current
Diameter DC+ OR weld Preheat preheat
Process Electrode range stringer temp control
(mm) DC- direction (C) method
(Amps)

1 MMAW(10) E7018 2.5 65-100 + VU yes 70 LPG crayons

1 MMAW(10) E4110 3.0-4.0 88-140 + VU yes 70 LPG crayons

1 GTAW(1)

crayon/digital
2 MMAW(10) E5518 2.7-3.2 120-180 + VD yes 75 LPG
thermo
LPG/Induction/ crayon/digital
3 MMAW(10) E8018G 2.0-5.5 145-200 + VD yes 100
Heatbeads thermo

3 MMAW(10) E4816 2.6-3.2 90-100 VU yes 45-95 LPG

4 MMAW(10) E5518G 2.5-4.0 70-140 + VD yes 75-100 LPG/induction digital thermo

5 GTAW(1)

5 MMAW(10) E4816 2.5-4.0 75-140 + VU yes 75-100 LPG/induction digital thermo

Table 6.3
The typical types of hot-tap fittings used in Australian and an indication of their relative usage

TYPES OF PIPE FITTING USED

Direct welded Branches with


Full encirclement
outlets. compensating saddles.
sleeves (rating 0-10)
(rating 0-10) (rating 0-10)

1 No Few only Yes


2 Yes (2) No Yes (5)
3 Yes (7) Yes (3) Yes (7)
4 Yes (10) No Yes (10)
5 Yes (5) Yes (5) Yes (10)
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 51

Table 6.4
Awareness and usage of the Battelle In-Service Welding Computer Program

USAGE OF BATTELLE SOFTWARE

Do you use the Battelle program ? Comments

1 No
2 Yes
3 Occasionally
not user friendly, inadequate guidance on usage of
4 Yes
results
limitations regarding thickness, and application to
5 Yes
high strength steels.

Table 6.5
Awareness and usage of the EWI test procedure to establish the cooling capacity of a pipe as
part of weld procedure development.

USAGE OF EWI TEST

Do you use the Would a standard for


Comments
EWI test ? EWI test be beneficial

1 No Yes
2 No
3 Yes Yes
time consuming and difficult to control temps to
4 Yes Yes
300-325C, can achieve 450C
concern about consistency of results and
5 Yes Yes
correlation with t 8/5 and heat-input data
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 52

Table 6.6
Typical restrictions placed on the performance of in-service welds in Australia.

Limits to In-Service Welding

Pipe wall thickness Pressure Limits


Is there a limit to Gas flow limitations to
limit for in-service for in-service
pipe grades which allow in-service welding
welding welding
can be welded (m/sec)
(mm) (MPa)

1 No >4.7
2 less than X46 >4.8 < 2.3 26 kscmh
3 No >4.3 todate < 5.5 depends on heat-sink
use Battelle
4 No >4.8 use EWI test
software

No but only welds No, but only welds


if required, depending on
5 up to X65 on >4.8 mm maybe
EW I test
completed todate completed todate
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 53

7 Experimental Appraisal of Hydrogen Controlled


Electrodes
7.1 Introduction
Numerical models cannot be developed in isolation from experimental data. There are often factors
within models that require empirical control. Although there is some published data for in-service
welds, often this data is not complete, for example, t8/5 cooling times may be quoted, but weld
penetrations and bead shapes are not.

The objective of our experimental work was therefore to generate data relating to in-service welds with
hydrogen controlled electrodes, since these are the recommended type of electrode for in-service
manual metal arc (MMA) welding, and they are typical of the electrode types used in Australia.

The experimental data were required to:

1. Accurately describe important parameters within the numerical models, such as the possible
variation of heat input with position around a circumferential weld, and the variation with welding
angle, and with electrode type.

2. To allow the assessment of crack susceptibility or hardness from predicted cooling rates. For this
an accurate relationship between hardness, composition and cooling rate must be established,
and tested against data from Australian steels.

Hence the specific objectives of these experiments were:

• to establish the typical variations of heat input with welding position around a circumferential weld;
• to measure deposition rate and weld bead geometry for single and multi pass welds, establishing
the variation with position around the pipe;
• to establish empirical relationships between bead volume, shape and process parameters, such
as heat input;
• to test the differences in behaviour between welds in the vertical-up and vertical-down position, on
a range of typical pipe steels;
• to develop and test the relationships between hardness and cooling behaviour;
• to give some indication of an appropriate representation of carbon equivalent for Australian
pipeline steels.

7.2 Experimental In-Service Welding Simulation


The typical approach is to artificially cool a small section of pipe on which to carry out test welds. In
general this simulates the increased cooling rate obtained during in-service welding but does not
simulate a pressurised system. There are a few published examples where, in order to investigate
burn-through, a pressurised, non-flowing gas has been used. Some limited experimentation has been
reported from welds carried out on a flow-loop or a by-pass in which test welds are carried out in
conditions that closely simulate ‘live’ welding. However for obvious safety reasons the majority of
experimental work has used some form of artificial cooling.

The EWI has carried out many of these types of experiments using a range of fluids, in an attempt to
determine the most suitable simulation system (1).
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 54

The basic difficulty in simulating the cooling conditions equivalent to a gas flow is illustrated in Figure
7.1 & 7.2. Here the relationship between heat input and t8/5 cooling time for a ‘live’ pipe is compared
with similar relationships for test pipes cooled with water, motor oil or compressed air. The water-
cooled pipe generates the lowest t8/5, that is, it produces the greatest cooling. Compressed air gives
significantly less cooling and motor oil slightly less cooling, than the pressurised methane in the live
pipe. Interestingly this effect is more significant with thinner walled pipes, as shown by comparing
Figure 7.1 with 7.2.

An alternative way to examine the same problem is to consider a calculation of the heat transfer from
the pipe wall. The approximate convective heat transfer coefficient can be determined from a non-
dimensional relationship such as Dittus-Boeler (Holman (2)), which relates the thermal properties of
the fluid and it’s flow rate to the effective heat transfer coefficient. Figure 7.3 shows the results of
such calculations for; methane under typical pipeline flows and pressures, air at low pressure, water
and two oils. To achieve the same convective heat transfer as methane with air would require a flow
rate approaching 100 m/sec. The light oil considered here, Mobiltherm 603 could generate the same
heat transfer coefficient as methane but it would also require the same flow rate 1-10 m/sec. These
flows would require a large quantity of fluid and huge pumps which are too high for practical
consideration. Water on the other-hand can achieve the same cooling effect as methane at a flow rate
of 0.1-0.5 m/sec. Unfortunately, a second effect also occurs with water which is not accounted for in
this graph, for at low flow rates the water boils, and this change in phase generates a large increase in
cooling effect.

This difficulty has generally been sidestepped in practical weld procedure development. The argument
has been proposed that water is a more aggressive coolant, therefore if a satisfactory t8/5 cooling time
is achieved with water, then the cooling rate will decrease and the weld hardness will be less, and
more satisfactory, on the operational pipeline. This means that weld procedures developed with water
flow will tend to be conservative. Unfortunately welds which are conservative with respect to hardness
may be non-conservative with respect to burn-through. Clearly burn-through is not considered by
such test procedures.

This topic remains unresolved, and for safety, and practical convenience, water is still commonly used
to simulate the accelerated cooling of a gas flow.

7.3 Early Difficulties


Early in the project a series of experiments were carried out using the “AUTORON”, an automated
stick welding machine at The University of Adelaide, see Figure 7.4. In principle this machine offered
an excellent experimental set-up, giving controlled heat input, and an adjustable fixed welding angle
for out-of position welding. A simple cooling chamber was manufactured, and simulated sleeve welds
attempted, using a flat plate held at a chosen out-of position angle, as illustrated in Figure 7.5.

Unfortunately a number of unresolved problems eventually caused this approach to be abandoned


since:

• good consistent quality welds could not be obtained,


• with thin electrodes (2.5 mm diameter) the position of the tip could not be held in a stable position,
• the electrode feeder failed to operate satisfactorily.

Following the failure of these “AUTORON” tests an experiment was established by Paul Grace at AGL
Ltd. This used a pipe segment and a similar water-cooled jacket, and a qualified, experienced pipeline
welder carried out the welding.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 55

7.4 AGL Experiments

7.4.1 AGL Experimental Equipment


The equipment consisted of a half segment of test pipe to which a quenching jacket could be clamped.
The cross sectional area of the jacket was kept small, so that the linear flow rates of water would be
high. A second segment of pipe approximately 100 mm wide was used to simulate the full
encirclement sleeve, see Figure 7.6. In this way all the welding positions used for a full
circumferential sleeve weld could be simulated. Thermocouples and a data logger gave temperature
versus time at regions near to the weld. The pipe segment together with thermocouples and water
jacket is shown in Figure 7.7.

A welding monitor was used to record arc current and voltage at 1 second intervals. In order to
determine heat input the time for each weld run was recorded and its length measured. From this the
average welding speed was determined for each weld run and used to calculate heat input.

Initially k-type thermocouples were spot-welded to the surface of the pipe and “MIM” ≈1 mm diameter,
sheathed thermocouples were inserted in holes drilled into the test pipe. Cooling of the weld bead
was also measured by using a ‘harpooned’ k-type thermocouple manually placed in the weld pool just
behind the arc, see Figure 7.8. This technique was sometimes prone to failure, usually caused by
melting of the thermocouple, but successful insertion generally gave a consistent record.

It proved to be extremely difficult to position the inserted “MIM” thermocouples into the extremely
narrow heat affected zone, which was often less than 1 mm thick. Eventually the difficulties with this
approach caused it to be abandoned and temperature cycles were measured by harpooned k-type
thermocouples only.

7.4.2 Experimental Plan


Two electrode types were chosen, namely E5548G(E8018G), used in the vertical-down position, and
E7016 used in the vertical-up position. Electrode diameters were 2.5, 3.2 and 4 mm diameter. These
are the popular electrodes used for in-service welding in Australia. When necessary priority was given
to E8018G in keeping with its relative popularity. Test circumferential single pass welds were carried
out using a simulated sleeve on non-cooled pipe to determine suitable welding conditions for each
electrode. These arc current and voltage settings are given in Table 7.1.

Samples from a range of pipes where obtained, as shown in Table 7.2. These were chosen to reflect
the typical range of pipe sizes, thickness and compositions used in Australia. An X80 pipe was
included.

A standard test pattern of welds was established as shown schematically in Figure 7.9 and in Table
7.3. For each weld the welding conditions were monitored, and the heat input determined. The t8/5
cooling times were measured from harpooned thermocouples. Each weld was radially sectioned and
macroscopic photographs taken. The HAZ hardness was established following the recommended
AGL procedure given in Appendix 7.1.

7.5 Experimental Results


A typical set of experimental results collected from one pipe test is given in Appendix 7.2. This
consisted of data for four multi-pass welds and 2 single pass root welds as identified in Figure 7.9 and
Table 7.3. This data set consisted of:
• measured t8/5 cooling time at identified positions from the pipe’s top-dead-centre for each weld run,
• measured heat input at each corresponding point,
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 56

• etched metallographic sections and macroscopic photographs taken at 5 positions from the top-
dead-centre for each weld,
• a metallurgical assessment of possible weld defects at each section,
• measured hardness profiles taken in the pipe’s HAZ for a section at the ‘3 oclock’ position in all
welds

Feng & Drmota (2) compiled the complete experimental results into a report, and a copy can be
obtained from CSIRO Manufacturing Science & Technology, if required.

7.6 Analysis of Experimental results


7.6.1 Average heat input under test conditions
As expected Figure 7.11 shows that the cooling conditions experienced during these tests are
generally greater then those obtained for an in-service weld on a pipe containing methane. Typical t8/5
cooling times for water-cooled tests lie in the range 2-6 seconds. In addition the variation of t8/5 with
welding heat input shows significant scatter, and less sensitivity to the heat input level than results
from EWI flow loop tests (1) would suggest, which is again possibly due to the increased cooling rate
in these tests.

7.6.2 The variation of heat input with welding position


It was established that during a given weld run there was a significant variation in heat input. This is
illustrated in Figure 7.12 where it is manifested as a stepped variation in heat input at varied position
around the circumferential sleeve weld. The stepped nature of the plot shown in Figure 7.12 is a
result of determining the average welding speed for a given weld run. The arc current and voltage did
not deviate, so the overall changes in heat input are a consequence of varied welding speed. It is
understandable that for a manual out-of-position weld there could be variations in welding speed.
However, the level of this variation is significant, for it is conservatively estimated to be ±20% of the
nominal level.

Because of this potential deviation in heat input between given weld runs it is not possible to make
rigorous statements with respect to the variation in heat input with angular position during
circumferential welding. However as Figure 7.12, 7.13, 7.14, 7.15 show there is a tendency for the
heat input at the 5 o’clock position to be higher than that at the 1 o’clock position. It is not possible to
differential between the two electrode types in this behaviour, both show a tendency to increase in
heat input at the 5 o’clock position by between 0% to +30% of the heat input at the 1 o’clock position.
However there are anomalies, which show the completely opposite behaviour so these observations
can only be taken as a guide.

Clearly one conclusion that can be made from this work is that if a specific heat input is desired some
method of assisting the welder to control welding speed is desirable.

7.6.3 The variation of weld penetration with welding position


The variation in weld penetration with angular position was also considered and this is shown plotted
in Figure 7.16 for E8018G and in Figure 7.17 for E7016. There is a significant scatter in measured
penetrations, although in agreement with previous reports (3) the E7016 electrodes used in the
vertical-up position do tend to give greater penetration. With the E8018G electrode there is a small
tendency for the penetration to be greater at the 5 o’clock position, whilst there is no discernable
relationship with the E7016 electrode.

7.6.4 The variation of weld penetration with heat input


From simple heat transfer considerations we expect that the weld penetration should increase with
increased heat input, but these cooled welds only show this tendency in an approximate way, as
shown in Figures 18 & 19. There is no clear relationship between heat input and penetration. It may
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 57

be that this is due to the nature of the electrodes which promote low-penetration, or due to the manual
influence of the welder.

7.6.5 Empirical relationships between bead volume, bead shape and process parameters
To form a numerical model of a welding process it is necessary to make some approximate predictions
of the weld-bead cross-section. This is usually calculated from the chosen heat input. This technique
allows the inclusion of a suitable bead shape within the model and often provides a datum for the heat
source definition (see Section 8).

In this work the deposition rate and characteristic dimensions of test welds were determined using the
following procedure. Firstly radial sections of each single pass weld were taken. These were
mounted, polished and photographed. A typical section is illustrated in Figure 7.20. From these
photographs the geometry of each weld was measured following the scheme also illustrated in Figure
7.20. The profile of the section was also digitized and the cross-sectional area determined. Simple
linear regression relationships were then established between the characteristic values and the
welding heat input, for both electrode types. Whilst the relationships indicated in Figure 7.21 & 7.22
are not precise, in that there is limited data and some scatter, the accuracy is satisfactory for the
intended use.

7.6.6 The relationships between hardness and cooling behaviour


The hardness data was measured at the 3 o’clock position as indicated in Appendix 7.1, and the
maximum value within the pipe’s HAZ was identified. These values are tabulated in Table 7.4.

The hardness of the heat affected zone (HAZ) is generally taken as a indicator of the potential for
hydrogen cracking. This is a simplification in that other factors such as microstructure, or applied
stresses also have an influence on cracking, but the approach is well established. Therefore the
usefulness of the numerical model lies not just in the ability to predict the weld’s cooling rate but in its
ability to predict hardness. In order to achieve this step it is necessary to establish the appropriate
relationship between hardness, material composition and cooling rate.

This has been carried out using the simple approach of comparing the measured hardness values for
single-pass welds with those predicted using various published regression equations.

t8/5 cooling time was used as the cooling parameter. Because there was some scatter in the values
determined from the harpooned thermocouples, the average value was used from each weld rather
than attempt to select a value corresponding to the 3 o’clock position.

A rather restricted range of values resulted. This was characteristic of these water-cooled tests, which
create greater cooling than that found in methane pipelines (see section 7.3.2). Using a Excel
spreadsheet produced by Bowie at BHP research (4), the predicted hardness using 12 different
empirical equations was determined for each pipe composition and average t8/5.

The relative error between the predicted hardness value and that measured was determined and
plotted in Figure 7.23. Feng, Drmota & Barbaro (5) have produced a technical report that suggests a
modification to allow for pipe compositions with Manganese contents less than 1%.

The lowest mean error of < 5% was achieved using the Yurioka –1 equation. The least spread of error
was obtained with the Dueren equation. Generally the relationships underestimated the HAZ
hardness.

7.6.6 Comparison of defects for welds in the vertical-up and vertical-down position
An attempt to qualitatively assess the relative integrity of welds achieved with both the E8018G and
E7016 electrodes was made. This involved visually assessing and categorising defects observed in
metallographic sections. This assessment was therefore not quantitative and relied on judgement and
past experience to give valid results.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 58

Figure 7.24 compares the spread of defects and indicates that both electrodes show similar types of
defect, with approximately the same level of defect-free welds. Slightly higher incidence of weld HAZ
cracking was observed with E8018G electrodes.

Figure 7.25 shows that similar defects, and defect levels were obtained with both high heat input and
low heat input welds.

Figure 7.26 compares the defect types and levels found for the different pipe materials welded with
both electrodes. It is difficult to be conclusive but there is some indication that a greater number of
defective welds were observed when welding X80, with a greater incidence of HAZ cracking when
using the E8018G electrode

7.6.7 Influence of multi-pass welds on tempering.


Table 7.4 and Figure 7.27 show the influence of multi-pass welds on the hardness achieved in the
completed weld. It is apparent that a significant reduction of HAZ hardness was not achieved in all
materials. In this case only AGL4 (350 to 300 HVN) and AGL8 (380 to 300 HVN) exhibit a significant
reduction in hardness after a 3-pass weld by comparison to the hardness achieved after the first root
pass. This difference in behaviour is unexplained. Welds on AGL15 (X80) exhibited a HAZ hardness
below 350 HVN for all single and multi-pass welds.

7.7 Conclusions
The analysis of welding conditions and welds produced using pipe cooled with a water jacket to
generate a rapid quench, gave the following results;

1. It was possible to generate a weld on 7.8 mm thick X80 pipe with a maximum hardness of 325
HVN with a weld t8/5 cooling time of 3.8 seconds.

2. For the range of pipe grades examined (X42-X80), the Yurioka-1 carbon equivalent (CE)
relationship provided the best agreement between composition, hardness and t8/5 cooling time.
This relationship gave an absolute error of 5.7% for E8018G electrodes, and a 3.4% error with
E7016. Based on this relationship the X80 grade of pipe used in this work would only require a t8/5
of 0.8 seconds in order to achieve a hardness >350 HVN.

3. A small beneficial tempering effect was measured from multi-pass [3 passes] welds. This was
most noticeable for the X70 grade steel, which with a CE of 0.288 gave a hardness of 381 HVN
after a single root pass. After three passes, the maximum hardness was reduced to an
acceptable 321 HVN.

4. In these tests only the two X60 steels with CEQ of 0.38 and 0.41 gave multi-pass hardness >350
HVN.

5. Some minor variation in the incidence of weld defects was observed between the two electrode
types tested. There was a greater incidence of HAZ cracking with E8018G in keeping with its
reduced heat input. Some HAZ cracks were detected with X80 grade although the maximum
hardness was 299-310 HVN.

6. There was no clear systematic variation in heat input as the weld progressed around the pipe.
However, whilst total weld energy remained reasonably constant the natural variation in welding
speed, since this is a manually skilled process, caused variation in the heat input. This was
significant and amounted to approximately a ±20% variation on a nominal value.

7. Penetration into the run pipe was generally greater when using the E7016 electrode in the vertical-
up position rather than with E8018G in the vertical-down.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 59

8. Penetration into the run-pipe slightly increased with increasing heat input although this effect was
largely swamped by a significant variability at a given heat input. At a nominal heat input of about
1 kJ/mm this variability was approximately, 0.2-0.8 mm with E8018G and 0.5-1.0 mm with E7016.

9. The deposited weld bead volume, and weld bead leg lengths on the pipe and sleeve were found to
be approximately proportional to the welding heat input, and hence simple empirical equations can
be used to establish weld bead geometry.

References
1. Bruce W. A. & Threadgill P. L., Effect of Procedure Qualification Variables for Welding Onto In-
service Pipelines, American Gas Association Report J7141, July 1994.

2. Holman J.P., 1976, Heat Transfer, 7th edition pub. McGraw-Hill.

3. Phelps B., Cassie B. A., & Evans N. H., Welding Onto Live Natural Gas Pipelines, Metal
Construction, August 1976, pp350-354.

4. Bowie G. BHP Research, Software for Calculation of HAZ Hardness. Private Communication.

5. Feng B., Drmota R. & Barbaro F., Mathematical Model To Calculate and Predict the HAZ
Maximum Hardness On the Basis of Steel Chemistry and Cooling Rate t8/5 After Welding,
Technical Report BHP-Integrated Steel Division, Port Kembla, NSW 2500. October 1998
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 60

15 T85 for 6.35 mm thick pipe


Water
Motor Oil
12.5
Air Flow

T85 cooling times (secs)


Data from Natural Gas Pipes
10

7.5

2.5

0
0 0.5 1 1.5 2 2.5
Heat Input (kJ/mm)

Figure 7.1
The comparison between the weld cooling times achieved on a 6.25 mm thick pipe containing
flowing methane and on the same pipe using laboratory simulations with different fluids (EWI
data(1)).

25 T85 data for 4.8 mm thick pipe


Water
Motor Oil
20 Air flow
Data from natural gas pipes
T85 cooling times

15

10

0
0 0.5 1 1.5 2 2. 5
Heat Input (kJ/mm)

Figure 7.2
The comparison between the weld cooling times achieved on a 4.8 mm thick pipe containing
flowing methane and on the same pipe using laboratory simulations with different fluids
(EWI data (1)).
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 61

100000
Effective heat transfer coefficient (W/m 2.K)

methane 'real' pipe


air flow low pressure
w ater flow 10000
Mobiltherm 603
Engine Oil
1000

100

10

1
0.01 0.1 1 10
Flow velocity (m/sec)

Figure 7.3
Predicted convective heat transfer coefficients for different flowing fluids compared with the
values calculated for methane gas under typical pipeline operational conditions.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 62

a ng ula r m ovem ent

test hea t c o nvec tive


c oeffic ient
5000 W/ m 2K

tilt, w eld a ng le
a d justm ent

w eld
m ovem ent

c oola nt
fl

Figure 7.4
Schematic of an in-service welding simulator used in early experiments on the AutoRon
automatic MMA welding machine at The University of Adelaide.

Figure 7.5
Autoron automatic welding machine used in simulation of in-service welding.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 63

WELD THERMAL
MONITOR DATA LOGGER
Embedded thermocouples

Arc current, voltage &


time Harpooned thermocouples

Velocity=L/t

Water inlet & outlet temperature

Figure 7.6
Schematic representation of the AGL in-service welding simulation
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 64

Figure 7.7
Pipe segment and the water jacket used for simulated in-service welds at AGL Sydney.

Figure 7.8
Weld pool harpooning to measure t8/5 cooling times.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 65

Table 7.1
Welding Parameter Sets established for E7016 & E8018G Electrodes

Test Electrode Dia. Weld Volts Amps Time Weld Heat


No Direction Length input
(mm) (min) (mm) KJ/mm
1 E7016 2.5 Vert-up 22.0 66.0 1.43 205.00 0.61
2 E7016 2.5 Vert-up 24.5 87.0 1.58 205.00 0.99

3 E7016 3.2 Vert-up 22.0 92.0 1.18 200.0 0.72


4 E7016 3.2 Vert-up 27.0 120.0 0.95 200.0 0.92

5 E7016 4.00 Vert-up 22.0 118.0 0.90 190.0 0.74


6 E7016 4.00 Vert-up 26.0 132.0 1.12 195.0 1.18

7 E8018G 2.5 Vert-down 21.0 80.0 1.00 205.0 0.49


8 E8018G 2.5 Vert-down 22.0 106.0 1.10 195.0 0.79

9 E8018G 3.2 Vert-down 19.0 92.0 0.97 170.0 0.60


10 E8018G 3.2 Vert-down 23.0 158.0 0.70 170.0 0.90

11 E8018G 4.0 Vert-down 20.0 150.0 0.95 210.0 0.81


12 E8018G 4.0 Vert-down 22.0 194.0 0.77 210.0 0.93
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 66

Table 7.2
Pipe Materials used in the AGL trials

Test Identification Outside Diameter Wall Thickness Material grade


Number (mm) (mm)
AGL 01 508.0 8.6 API 5LX60 AGL
AGL 02 406.0 8.6 API 5LX70(Nippon)AGLPWA
AGL 03 406.0 7.8 API 5L X60 AGLPQ
AGL 04 323.9 5.2 API 5L X42 AGLPQ
AGL 08 323.9 5.2 API 5L X52
AGL 09 168.0 3.6 API 5L X42
AGL 11 558.0 9.5 API 5L X42 AGL
AGL 12 508.0 6.4 API 5L X60 AGLC
AGL 15 406.0 7.8 X80
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 67

Table 7.3 Compositions of Line Pipe used in the AGL tests.

C P Mn Si S Ni Cr Mo Cu Al Sn Nb Ti V B Ca N IIW CEQ
AGL.01 0.16 0.009 1.28 0.24 0.004 0.01 0.026 0.002 0.009 0.021 <0.002 0.055 0.003 <0.003<0.0003 <0.0005 0.0064 0.38
AGL.02 0.07 0.014 1.61 0.25 0.003 0.019 0.022 0.003 0.009 0.012 <0.002 0.05 <0.003 0.064 <0.0003 <0.0005 0.0024 0.358
AGL.03 0.065 0.018 1.35 0.12 0.003 0.02 0.024 0.004 0.021 0.025 0.002 0.037 0.02 0.004 <0.0003 0.0028 0.0057 0.299
AGL.04 0.17 0.009 0.68 0.005 0.014 0.014 0.008 0.002 0.025 0.005 0.002 <0.003<0.003<0.003<0.0003 <0.0005 0.0024 0.288
AGL.08 0.09 0.014 1.13 0.14 0.006 0.018 0.018 0.003 0.013 0.004 0.002 0.021 <0.003<0.003<0.0003 <0.0005 0.0054 0.285
AGL.09 0.115 0.021 0.68 0.12 0.006 0.021 0.021 0.003 0.012 0.022 <0.002<0.003<0.003<0.003<0.0003 0.0033 0.0042 0.236
AGL.11 0.06 0.015 1.06 0.22 0.003 0.02 0.011 0.003 0.009 0.017 0.002 <0.003 0.013 <0.003<0.0003 0.0005 0.0041 0.242
AGL.12 0.19 0.015 1.29 0.24 0.004 0.013 0.022 0.003 0.01 0.021 <0.002<0.003 0.003 0.033 <0.0003 <0.0005 0.005 0.418
AGL15 0.065 0.018 1.63 0.34 0.003 0.029 0.027 0.21 0.009 0.03 0.012 0.051 0.013 0.004 <0.0003 0.0007 0.006 0.387

Data provided by BHP Research.


IIWCEQ = C + Mn/6 + (Cr + Mo +V)/5 +(Ni + Cu)/15
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 68

AGL XX.3 AGL XX.4

AGL XX.2 AGL XX.5

AGL XX.1 AGL XX.6

Figure 7.9
Schematic plan for each AGL Welding trial

Table 7.4

Weld Welding Welding Target Number of Cooling


Code Electrode Direction Heat input weld runs Water
AGL.00.WX Vertical Down high (HHI) Flow rate
(VD), or or low (LHI) Litres/min
Vertical Up (VU)
1 3.2 mm diameter, VD LHI Multi-pass 3
E8018G 2-3
2 3.2 mm diameter, VD HHI Multi-pass 3
E8018G 2-3
3 3.2 mm diameter, VU LHI Multi-pass 3
E7016 2-3
4 3.2 mm diameter, VU HHI Multi-pass 3
E7016 2-3
5 3.2 mm diameter, VD LHI 1 root 3
E8018G
6 3.2 mm diameter, VU LHI 1 root 17
E7016
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 69

14

EWI Flow Loop, 4.8 mm


12 EWI Flow Loop, 6.35 mm
AGL Trials, Water flow , 8.6 mm
Cooling Times T85C (seconds)

AGL Trials, Water flow , 5.2 mm


10

0
0 0.5 1 1.5 2 2.5
Heat Input (kJ/mm)

Figure 7.10
Comparison between the range of cooling times achieved within the AGL experiments and those
obtained from flow loop tests using methane flow published by EWI
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 70

1.4

1.2

1
Heat Input (kJ/mm)

0.8

0.6

Low Heat Input - Root Pass


0.4 High Heat Input - Root Pass

0.2

0
0 30 60 90 120 150 180
Degrees After Top Dead Centre

Figure 7.11
Typical variation in heat input for two different circumferential fillet welds carried out during the
AGL tests.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 71

2
2 E8018G E8018G
root pass 0.75 kJ/mm target 1.8 root pass, 1.1 kJ/mm target
1.8
2nd pass, 0.75 kJ/mm target 2nd pass, 1.1 kJ/mm target
1.6 1.6
3rd pass, 0.75 kJ/mm target 3rd pass, 1.1 kJ/mm target
1.4
Heat Input (kJ/mm)

1.4 single pass

Heat Input (kJ/mm)


1.2 1.2

1 1

0.8 0.8

0.6 0.6

0.4 0.4

0.2 0.2

0 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6

Circumferential position from TDC Circumferential position from TDC

2 2
E7016 E7016
1.8 1.8

1.6 1.6

1.4 1.4
Heat Input (kJ/mm)
Heat Input (kJ/mm)

1.2 1.2

1 1

0.8 0.8

0.6 0.6
Root pass, 1.4 kJ/mm target Root pass, 0.8 kJ/mm target
0.4 0.4
2nd pass
2nd pass
3rd pass
0.2 0.2
3rd pass Root pass

0 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6

Circumferential position from TDC


Circumferential position from TDC

Figure 7.12
The variation of heat input with circumferential angular position from top-dead-centre.
These results are for pipe AGL 01.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 72

1.2
1.2
E8018G E8018G
1
1
Heat Input (kJ/mm)

0.8

Heat Input (kJ/mm)


0.8

0.6
0.6

0.4 0.4

root pass, 0.75 kJ/mm target root pass, 1.4 kJ/mm target
0.2 2nd pass 2nd pass
0.2
3rd pass
3rd pass
root pass
0 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6
Circumferential position from TDC Circumferential position from TDC
1.6
1.6 E7016
E7016
1.4
1.4

1.2
1.2
Heat Input (kJ/mm)
Heat Input (kJ/mm)

1
1

0.8
0.8

0.6 0.6

0.4 0.4
Root pass, 0.8 kJ/mm target Root pass, 1.4 kJ/mm
2nd pass target
0.2 3rd pass 0.2 2nd pass
Root pass
0 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6
Circumferential position from TDC Circumferential position from TDC

Figure 7.13
Variation of heat input with circumferential angular position from top-dead-centre.
This data is for pipe AGL 12
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 73

1.4 1.4
E8018G E8018G

1.2 1.2

1 1

Heat Input (kJ/mm)


Heat Input (kJ/mm)

0.8 0.8

0.6 0.6

0.4 0.4

root pass, 0.75 kJ/mm target root pass, 1.1 kJ/mm target
0.2 2nd pass 0.2 2nd pass
3rd pass
3rd pass
root pass
0 0
0 1 2 3 4 5 0 1 2 3 4 5
Circumferential position from TDC Circumferential position from TDC

1.6 1.6
E7016 E7016
1.4 1.4

1.2 1.2
Heat Input (kJ/mm)

Heat Input (kJ/mm)

1 1

0.8 0.8

0.6 0.6

0.4 0.4
Root pass, 0.8 kJ/mm target
2nd pass Root pass, 1.4 kJ/mm
0.2 3rd pass 0.2 target
Root pass 2nd pass

0 0
0 1 2 3 4 5 0 1 2 3 4 5
Circumferential position from TDC
Circumferential position from TDC

Figure 7.14
Variation of heat input with circumferential ‘o-clock’ position from top-dead-centre.
This data is for pipe AGL 02
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 74

1.4 1.4
E8018G E8018G
1.2 1.2

1 1

Heat Input (kJ /mm)


Heat Input (kJ/mm)

0.8 0.8

0.6 0.6

0.4 0.4
root pass, 0.75 kJ/mm target
2nd pass root pass, 1.1 kJ/mm target
3rd pass 2nd pass
0.2 0.2
root pass
3rd pass

0 0
0 1 2 3 4 5 0 1 2 3 4 5
Circumferential position from TDC Circumferential position from TDC
2
E7016
1.6
1.8
E7016
1.4 1.6

1.2 1.4
Heat Input (kJ/mm)
Heat Input (kJ/mm)

1.2
1

1
0.8
0.8
0.6
0.6
0.4
Root pass, 0.8 kJ/mm target
0.4 Root pass, 1.4 kJ/mm target
2nd pass 2nd pass
0.2
3rd pass 0.2 3rd pass
Root pass
0 0
0 1 2 3 4 5 0 1 2 3 4 5
Circumferential position from TDC Circumferential position from TDC

Figure 7.15
Variation in heat input circumferential ‘o-clock’ position from top-dead-centre.
This data is for pipe AGL 15
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 75

2 agl 1 agl 2
E8018G Electrode 3.2 mm diameter
agl 4 agl 8
1.8
agl 12 agl 15
Weld penetration into Run pipe (mm)
1.6

1.4

1.2

0.8

0.6

0.4

0.2

0
0 1 2 3 4 5 6
o'clock position from TDC

Figure 7.16
Changed weld penetration into the run pipe with different positions from top-dead-centre during a
circumferential sleeve weld. For single weld passes using E8018G electrodes of 3.2 mm
diameter.
2 E7016 Electrode 3.2 mm diameter agl 1 agl 2

agl 4 agl 8
1.8
Weld Penetration in Run Pipe (mm)

agl 12 agl 15
1.6

1.4

1.2

0.8

0.6

0.4

0.2

0
0 1 2 3 4 5 6
o'clock position from TDC

Figure 7.17
Changed weld penetration into the run pipe with different positions from top-dead-centre during a
circumferential sleeve weld. Results are for single weld passes using E7016, vertical-up,
electrodes of 3.2 mm diameter.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 76

agl 1, X60, 8.6 mm


agl 2, x70, 8.6 mm
2
E8018G Electrode, 3.2 mm diameter agl 4, X42, 5.2 mm
Weld Penetration in Run Pipe (mm)

1.8 agl 8, X52, 5.2 mm

1.6 agl 12, X60, 6.4 mm


agl 15, X80, 7.8
1.4

1.2

0.8

0.6

0.4

0.2

0
0.6 0.8 1 1.2 1.4
Heat Input (kJ/mm)

Figure 7.18
Relationship between weld penetration into the run pipe with heat-input for circumferential sleeve
welds. These results are for single weld passes using E8018G electrodes of 3.2 mm diameter.

1.4 E7016 Electrode 3.2 mm diameter

1.2
Weld Penetration in Run Pipe (mm)

0.8

0.6
agl 2 X70, 8.6 mm
agl 4, X42, 5.2 mm
0.4
agl 8, X52, 5.2 mm
agl 12, X60, 6.4 mm
0.2
agl 15, X80, 7.8 mm
agl 1, X60, 8.6 mm
0
0.6 0.7 0.8 0.9 1 1.1 1.2 1.3 1.4

Heat Input (kJ/mm)

Figure 7.19
Changed weld penetration into the run pipe with different positions from top-dead-centre during a
circumferential sleeve weld. These results are for single weld passes using E7016, vertically-up,
electrodes of 3.2 mm diameter.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 77

Pipe leg length

Sleeve leg length

45° diagonal length

Figure 7.20

Metallographic section of weld bead and values taken to indicate shape and volume.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 78

Weld Bead Area E8018G Electrode


20
18
Weld Bead Sectional Area

16
14
12
(mm 2 )

10
8 8
6 E8018G Pipe leg
4 y = 24.05x - 4.2445 7
2

Pipe Leg Length (mm)


6
0
0 0.2 0.4 0.6 0.8 1 1.2
5
Heat Input (kJ/mm)
4

2
6
E8018G bead diagonal y = 4.2566x + 1.4035
1 2
R = 0.5237
5
0
Diagonal length (mm)

4 0 0.5 1 1.5
Heat Input (kJ/mm)

2
8 E8018G sleeve leg
y = 3.3669x + 1.3676
1 2 7
R = 0.7289
Sleeve leg length (mm)

6
0
0 0.5 1 1.5 5
Heat Input (kJ/mm)
4

2 y = 3.7421x + 2.0404
2
1 R = 0.3517

0
0 0.5 1 1.5
Heat Input (kJ/mm)

Figure 7.21
Simple linear relationships between weld bead geometry and welding heat input for 3.2 mm
diameter E8018G electrodes used in the vertical-down position.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 79

Weld Bead Area E7016 Electrode


20
18
Weld Bead Sectional Area

16 y = 14.176x - 0.4863
14
12 8
(mm 2 )

10
7 E7016 Pipe leg
8
6 6

Pipe Leg Length (mm)


4
2 5

0
4
0 0.5 1 1.5
Heat Input (kJ/mm) 3

y = 1.9803x + 3.1389
6 1
R2 = 0.3664
E7016 bead diagonal
0
5 0 0.5 1 1.5 2
Heat Input (kJ/mm)
Diagonal length (mm)

3 7

E7016 sleeve leg


2 6

1 y = 1.2178x + 2.7903 5
Sleeve leg length (mm)

R2 = 0.2421
4
0
0 0.5 1 1.5 2
3
Heat Input (kJ/mm)

1
y = 1.0495x + 3.403
R2 = 0.1836

0
0 0.5 1 1.5 2
Heat Input (kJ/mm)

Figure 7.22
Simple linear relationships between weld bead geometry and welding heat input for 3.2 mm
diameter E7016 electrodes used in the vertical-up position.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 80

E8018G, 3.2 mm diameter, nominal heat input 0.75 kJ/mm

agl 1, X60
Percentage error in Predicted

80 agl 2, X70
70 agl 4, X42
60 agl 8, X52
Hardnesss

50 agl 12, X60


40 agl 15, X80
30 agl 11, X42

20
10
0
1 2 3 4 5 6 7 8 9 10 11 12
Regression relationship

E7016 3.2 mm diameter, nominal heat input 0.8 kJ/mm

80 agl 1,X60
Percentage Error in Predicted

agl2, X70
70
agl 4, X42
60 agl 8, X52
50 agl 12, X60
Hardness

agl 15, X80


40
agl 11, X42
30

20

10

0
1 2 3 4 5 6 7 8 9 10 11 12
Regression Equation

Figure 7.23
An appraisal of the accuracy with which various regression relationships can predict HAZ
hardness based on the measured t8/5 cooling times.
1 Suzuki BL70, 2 Suzuki BL70S (Pcm), 3 Suzuki BL70S after Ito, 4 Suzuki BL70SM, 5 Suzuki
BL70S Cem, 6 Suzuki BL70S CE, 7 Beckert, 8 Yurioka –1, 9 Yuroka –2, 10 Yuroka –3, 11
Dueren, 12 Terasaki JOM-2.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 81

0.6

0.5

Proportion of samples
Electrode type E8018G (0.75-1.1 k J/mm)
0.4 Electrode type E7016 (0.8-1.4 k J/mm)

0.3

0.2

0.1

0
ok wms p lof uc hazc rootc wmc misc
Microstructural defects identified

Figure 7.24
A diagrammatic representation of the spread and distribution of weld defects observed from
metallographic sections of AGL experimental welds
Ok satisfactory weld,
wms weld metal segregation
p porosity
lof lack of fusion
uc undercut
hazc heat affected zone cracking
rootc cracks indicated in region of weld root
wmc cracks observed in the weld metal
misc other defects not otherwise identified above
0.6

0.5

Low Heat Input (0.75-0.8 k J/mm)


Proportion of samples

0.4 High heat Input (1.1-1.4 k J/mm)

0.3

0.2

0.1

0
ok wms p lof uc hazc rootc wmc misc
Microstructural defects identified

Figure 7.25
The variation of defect types and population for welds at high or low heat input.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 82

AGL test pipes welded with E8018G


40
AGL01,X60,8.6 mm

35 AGL12, X60, 6.4 mm


AGL4, X42, 5.2 mm
30
Percentage of samples AGL8, X52, 5.2 mm

25 AGL2, X70, 8.6 mm

AGL15,X80, 7.8 mm
20

15

10

0
ok wms p lof uc hazc rootc wmc misc
Microstructural defects

50
AGL test pipes welded with E7016
45
AGL01, X60, 8.6 mm
40
AGL12, X60, 6.4 mm
35
Percentage of samples

AGL4, X42, 5.2 mm

30 AGL8, X52, 5.2 mm


AGL2, X70, 8.6 mm
25
AGL15, X80, 7.8 mm
20

15

10

0
ok wms p lof uc hazc rootc wmc misc
Microstructural defects

Figure 7.26
The influence of electrode type on the population of weld defects observed during the AGL
experiments.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 83

Table 7.4 Experimental data for single and multi-pass welds carried out during the AGL
experiments.

DATA FROM SINGLE ROOT PASS WELDS


Pipe
Sample Heat Input Maximum Heat Input Maximum
Steel Grade IIWCEQ Thickness T85C (sec) T85C (sec)
Number (kJ/mm) HVN5 (kJ/mm) HVN5
(mm)
E8018G,vertical down E7016, vertical up
agl 11 X42 0.242 9.5 0.84 3.3 212 0.84 4 223
agl 8 X52 0.285 5.2 0.75 3 317 0.8 2.3 329
agl 1 X60 0.38 8.6 0.75 4.3 391 0.8 4.3 391
agl 12 X60 0.418 6.4 0.75 4.1 391 0.8 3.7 412
agl 2 X60 0.358 8.6 0.75 3.9 321 0.8 4.2 317
agl 4 X70 0.288 5.2 0.75 2.1 381 0.8 2.2 381
agl 15 X80 0.387 7.8 0.75 3.85 325 0.8 4.9 317

AVERAGED DATA FROM MULTI-PASS WELDMENTS


Pipe
Sample Heat Input Maximum Heat Input Maximum
Steel Grade IIWCEQ Thickness T85C (sec) T85C (sec)
Number (kJ/mm) HVN5 (kJ/mm) HVN5
(mm)
E8018G low heat E8018G high heat
agl 8 X52 0.285 5.2 0.75 4.4 283 1.1 4.2 289
agl 1 X60 0.38 8.6 0.75 3.4 391 1.1 3.7 386
agl 12 X60 0.418 6.4 0.75 2.9 376 1.1 3.4 401
agl 2 X60 0.358 8.6 0.75 3.5 321 1.1 3.7 313
agl 4 X70 0.288 5.2 0.75 3.6 274 1.1 3.3 321
agl 15 X80 0.387 7.8 0.75 3.4 313 1.1 3 329
E7016 low heat E7016 high heat
agl 8 X52 0.285 5.2 0.8 2.9 283 1.4 3.1 283
agl 1 X60 0.38 8.6 0.8 4.3 386 1.4 4.1 362
agl 12 X60 0.418 6.4 0.8 3.8 401 1.4 3.2 407
agl 2 X60 0.358 8.6 0.8 3.9 303 1.4 4.2 345
agl 4 X70 0.288 5.2 0.8 3.3 296 1.4 3.1 303
agl 15 X80 0.387 7.8 0.8 4.5 299 1.4 4.8 310
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 84

E8018G target heat input 0.75 kJ/mm


420
Single root pass
400
After 3rd multi-pass
Maximum HAZ hardness HVN5
380
360
340
320
300
280
260
240
220
200
agl 8 agl 1 agl 12 agl 2 agl 4 agl 15
Test pipe (see Table 7.4 )

420
E7016 target heat input 0.8 kJ/mm

400 Single pass


After the 3rd multi-pass
380
Maximum HAZ hardness HVN5

360

340

320

300

280

260

240

220

200
agl 8 agl 1 agl 12 agl 2 agl 4 agl 15
Test pipe (see Table 7.4)

Figure 7.27
Representation of the beneficial reduction in hardness caused by multi-pass welds.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 85

8 Numerical Modelling of In-Service Welding


8.1. Introduction
Using a computer simulation in order to calculate thermal effects of a fusion welding process is a
well-established technology. Simulation can be used to estimate weld penetration, the heat
affected zone (HAZ) size and the weld cooling rates, based primarily on the welding process
parameters of arc current, voltage and speed. In this way, simulations can be used to determine
an acceptable weld procedure, and the normal ‘experimental’ trial and error approach is replaced
with trial and error computer simulation.

Computer simulation has some drawbacks however, for it takes some time to carry out
calculations and accuracy is not guaranteed. The numerical methods have to make some
approximations in order to represent the complex physics of a welding process and achieve a
minimum calculation time. This often means the inclusion of empirical ‘fitting’ parameters to
achieve an agreement between simulation and practice. With present simulation costs and
computer speeds the numerical approaches find it difficult to compete with conventional
experimentally based weld-procedure development for simple weld types. However, for complex
experimental tests, the use of computer simulation is a cost effective alternative.

Since there is no simple comprehensive way of experimentally developing an in-service weld


procedure, the competition between an empirical based approach and the use of simulation is
tipped in favour of simulation. For example, experimental welds on pipes cooled with flowing water
can be used in procedure development but this approach fails to simulate the pressurised gas
flow, and water is a more aggressive coolant than gas, so the resulting weld procedures are
generally conservative. That is they produce heat-inputs which when applied to a pipeline
containing gas will generate a longer cooling time and less hardness than achieved in the
experiment.

The useful role of numerical simulation of in-service welding has already been demonstrated.
During the 1970’s, researchers at Battelle Memorial Institute (Keifner et. al. (1,2)) developed a 2D
finite difference approach to simulate sleeve and direct-branch in-service welds. These models
calculated an effective heat transfer coefficient at the inner pipe-wall, based on the pressure and
flow conditions within the pipe. For a given pipe geometry, and a set of welding parameters, the
weld cooling time, e.g. t8/5 (time to cool from 800° to 500°C) and the maximum inside-wall
temperature were calculated. The HAZ hardness was then estimated from the cooling rate and
the steel’s carbon equivalent. Hardness below 350 HVN was considered to have a low cracking
potential. In the work at Battelle, burn-through limits were not directly calculated but were based
on applying a limit for the maximum inner-wall temperature of 980°C. Above that temperature, the
pipe-wall was considered to have almost no strength. These models were extensively tested
against experimental data, derived largely from tests with relatively thick (>6.35 mm), low strength
pipes. Data analysis concentrated on comparisons between measured and calculated thermal
data.

In 1992, Goldak et al (3) applied a more general 3D finite element method to calculate the thermal
fields for circumferential fillet and direct-branch welds. They also calculated the residual stress
field after completion of a circumferential weld.

8.2 Current Numerical Models of an In-Service Weld


The following sections describe the modelling strategies developed in our current work including
some typical results. The broad thrust of this work is to develop advanced models and use these
to compare predicted and measured values of in-service welds, cooling times e.g. t8/5, or weld
penetrations. Having established suitable simulations, these will be used to establish guidelines for
safe welding procedures.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 86

2D models were developed at Battelle(1,2), but as indicated by Goldak et al(3) the joint types in in-
service welding do not ideally conform to 2D approximations. Stress analysis of local region within
the wall of a pressurised pipe cannot be treated in a 2D manner, so in our work we are
concentrating on full 3D models for both circumferential fillet joints and direct branch pipe welds.
(Of course 2D models of the type generated by Battelle can also be produced using the
commercial software used in this project. Such models for sleeve and saddle welds were initially
produced, but since they contribute little to the advancement of the numerical simulation of in-
service welding they are not reported here).

8.3 A 3D Quasi Steady State (3DQSS) Model of a Circumferential Sleeve


Weld

Naturally, during welding, the welding-arc moves around the fillet joint. However when stop and
start conditions are ignored and the welding conditions, such as heat input and electrode angles
are constant, then a quasi-steady state condition can be expected for which the temperature field
around the welding arc is unchanging with time. This approach leads to a simplified 3D model.
The solution does not have to consider transient effects, or stepping through time, so solution time
is greatly reduced. Since the arc is effectively fixed, the finite element mesh can be easily graded
in element size in the arc region as shown in Figure 8.1, giving improved accuracy. There are
generally less elements required for a given pipe geometry leading to reduced calculation times.
A moving arc requires a refined mesh along the whole welding path leading to more unknowns
and slower calculations.

These models are based on a conductive heat transfer with the welding arc represented as a
distributed heat flux. For smoothness and stability of solution it is generally expected that around
10 elements are required within the heat source. This density is related to the presence of high
thermal gradients in this region. Similarly the radius or width of the heat source is typically of the
same order as the weld pool. For the manual metal arc (MMA) weld pool this may be about 5-10
mm. This gives an approximate element size within the weld of 1-2mm. In a transient calculation,
the length of the weld may need to be divided into elements of this size. Thus even for relatively
short welds, this clearly requires large numbers of elements. Being able to model an effectively
static weld in a quasi steady state has substantial benefits in such cases.

By making some compromises on boundary conditions only a portion of the pipe needs to be
considered. A gap of 0.25 mm is allowed between the sleeve and the pipe. Using a material of
reduced thermal conductivity in this region simulates the intermittent contact between sleeve and
run pipe. Thermally dependent material properties are used and latent heat during phase changes
is accounted for. The convective heat coefficient is calculated from the pipe flow conditions and
applied to the pipe-wall. Natural convection and radiation losses are applied on other surfaces.
The values used are typical of those used for thermal models. For example:

• convective heat transfer coefficient 0.00012 W/mm2.C


• emmissivity 0.9
• Stephan-Boltzmann constant
• ambient temperature 300°K

Various heat source models were tested, based around the ‘double ellipsoidal’ source (4). The
heat source is positioned at an angle to simulate a variable electrode position as shown in Figure
8.2. A typical distributed of heat flux is plotted in Figure 8.3; the details of heat source formulation
will be described in the following section.

The calculation methodology has been implemented within the NISA Fluid Finite Element Analysis
software. In order to easily create finite element meshes of various welds the following approach
has been developed:
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 87

• Initially a 2D finite element mesh is developed for the pipe and sleeve segment using linear 4
noded quadrilateral elements (Figure 8.4).
• The 3D mesh of 8 noded ‘brick’ elements is then determined using the ‘cylindrical’ copy
function within the NISA-Display3 (5) software (Figure 8.5).
• The final mesh is then generated by grading the mesh spacing as shown in Figure 8.6 to
achieve the desired high density of mesh in the weld region.

External programs have been created which automatically expand the mesh and calculate and
apply both the heat source boundary conditions and the convective and radiative losses at the
inside and outside of the pipe surface.

The weld bead is modelled, based on an empirical estimation of the deposition area. A dummy
weld region (one having very low thermal conductivity) is placed in front of the weld source, to
simulate the progressively deposited weld, as shown in Figure 8.7. A non-linear quasi-steady-
state solution is achieved using temperature dependent viscosity, thermal conductivity, specific
heat and density.

A calculated thermal field for a 95 Amp, 22 Volt, circumferential fillet weld made at 2.3 mm/second
is shown in Figure 8.8. The welded pipe is 273 mm in diameter and 4.88 mm wall thickness, and
contains methane at 4.14 MPa pressure and a flow velocity of 6.3 m/second.

Cooling cycles at any point can be determined from the thermal field and plotted on a section
through the weld model as shown in Figure 8.9. Maximum temperatures can be calculated and
are shown in Figure 8.10. An estimation of weld pool penetration (2.3 mm here) and the HAZ
depth can be made from this plot by using the melting point, 1500°C, and the 720°C isotherms.
This is the model equivalent of a metallographic section.

Making use of an empirical relationship between the cooling rate, carbon equivalent of the steel,
and resulting hardness, a calculation of hardness can also be made. In this case the t8/5 cooling
time and the Yurioka -1 equation as described in [Section 7] was used, to give the estimated
hardness through out the weld zone of this X70 steel as shown in Figure 8.11.

8.4 Transient, Moving Arc, Thermal Models of Branch Fillet Welds.


Direct branch welds cannot be simulated using the quasi-steady-state approximation because the
weld path is not of constant geometry. In this joint the welding angle changes relative to the run
pipe and the weld geometry varies as the joint progresses. This must be simulated by a transient
model in which the heat distribution simulating the welding arc is stepped around the welded joint.
Figure 8.12 shows a typical model result.

Whilst it is feasible to generate a model of the full weld, as illustrated in Figure 8.12, it is not
necessary to do this. An in-service weld has a very localised temperature field so it is acceptable
to restrict the model to a segment of the full joint. This reduces the problem size and helps reduce
calculation time.

Because of the changing joint angle the heat source must be manipulated as the weld proceeds
around the joint, see Figure 8.2(a), and Figure 8.13. The coordinate transformations and
numerical procedures necessary to achieve this have been developed and incorporated into a C-
program, which calculates the necessary transient heat flux variationon the nodes or elements in
the weld path. These heat flux values are also formed into the syntax used by the NISA software
to describe a time varying heat source.

Transient models for direct branch fillet welds or circular sleeve flanges have been developed, as
shown in Figure 8.14. The solution strategy is implemented using NISA-Heat2 software. It uses
non-linear thermal properties and an iterative procedure. The convective heat transfer losses from
inside and outside of the pipe are calculated in the same manner as discussed in the 3DQSS
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 88

models (see section 8.3).

After the creation of the finite element mesh within NISA-Display3, pre-processing programs have
been developed to carry out the calculation of boundary conditions.

Transient models of circumferential fillet welds can also be developed and t8/5 cooling times, and
weld and HAZ geometry and hardness can be derived. Figure 8.15 shows the typical results
indicating the overall temperature field and the size and shape of the weld pool. Since this is a
transient model it is also capable of estimating the non-steady state cooling effects at the end of
the weld run, as illustrated by the thermal cycles in Figure 8.16, at different points along the weld
bead.
An assessment of accuracy has been made for calculated weld cooling times using some t8/5 data
for various circumferential fillet test welds carried out by Edison Welding Institute (6). As shown in
Figure 8.17, the calculated t8/5 values are lower than measured ones, with the transient model
being generally more accurate than the 3DQSS one. (A more comprehensive evaluation of the
model’s accuracy will be reported later in Section 10)

8.5 Development of Numerical Models of In-Service Welding


Four significant research aspects have to be considered in the development of numerical models
for in-service welding. These are:

• an appropriate formulation of the heat source for the common in-service welding processes,
i.e. vertical-up and vertical-down MMA welding with hydrogen controlled electrodes,

• selection of the best, most acceptable method of representing the heat losses created by the
flowing gas (having regard to computer time and practicality),

• the procedure for calculating the weld HAZ hardness from the calculated cooling rate and steel
composition,

• the definition and adoption of a safe and relevant criteria to signify the onset of burn-through.

8.5.1 Forming a Heat Source to Represent the In-service Welding Arc


In thermal models, heat transfer by convective flow of molten material within the weld pool is
complex and it is usually modelled as though it were an equivalent conductive heat transfer
process. For such thermal models, the key to accuracy lies in choosing an appropriate heat source
to represent the welding process. Often the ‘double-ellipsoidal’ heat source shown in Figure 8.18,
suggested by Goldak et al (4) forms the mathematical basis of such sources.

The heat of the welding arc transferred to the weldment (Q = I.V.η, where I is arc current, V the arc
voltage an η is the arc efficiency) is distributed in a Gaussian manner throughout an ellipsoidal
volume. The efficiency factor (η) accounts for energy loss, and it is the first of a number of
essentially arbitrary parameters used to define the heat source. These parameters are related to
the expected non-uniform distribution of energy within the arc, and to the simulation of pool flow
which distributes heat within the molten weld pool. For example in the double-ellipsoidal source,
convective effects are partly simulated by the elongation of the source (as described by cf & cb the
source axial lengths, in Figure 8.18); and partly by dividing the heat content between the front and
rear of the source, using factors rf & rb; thus giving heat inputs Q.rf & Q.rb, into the front and rear of
the heat source.

For out-of-position MMA welding the definition of the heat source must consider other aspects of
the welding process. These include simulating the effect of welding angle and welding position,
and the need to consider manipulation of the electrode during this skilled manual process. Cola et
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 89

al (7) considered that the effect of welding angle was important and it is included in the Battelle
model. It was stated (1) however, that a wide variation in welding angle was not used, since the
welder must penetrate both the sleeve and the run-pipe.

The Battelle models (1) use a uniformly distributed heat source and an idealised fillet weld, with a
flat profile. Goldak et al (3) found that the assumed volume and shape of the fillet weld had a
significant influence on the penetration and temperature profile around the weld pool. The Battelle
models do not adjust the heat source for different electrode types or vertical-up, or vertical-down
welding positions. It was recognised that low-hydrogen electrodes have low penetration
characteristics, however in their models, penetration characteristics were allowed for by altering
the burn-through limits set for inside-wall temperature.

In thermal conduction models empirical factors must be incorporated, whether these are welding
efficiencies or source geometry The ‘correct’ choice can only be determined by comparison
between predictions and experimental data. Experimental validation is therefore an essential part
of any weld simulation.

The heat source adopted in this work has the following features:

1. It produces a relatively shallow penetration in keeping with the characteristic of low-hydrogen


MMA welding. This is achieved by using a shallow source ‘disc’ like source.
2. Has a size that is related to the electrode diameter and the deposition rate of the process,
which is in turn dependent on the heat input.
3. Is orientated at a weld angle determined by the shape of the weld bead. This in turn is
determined from an empirical relationship between, heat input and weld leg lengths for a given
electrode type.
4. Has a defined orientation and position which is related to the weld bead shape and size, with
the bead shape established from an empirical relationship with heat input (see Section 7).

The distribution is based on a modification to the ‘double ellipsoidal’ source. However since the
MMA welding process almost always involves some small degree of weaving, the heat distribution
is modified from the Gaussian form.

The mathematical basis of the energy distribution followed that used by Goldak et al(4) in the
derivation of the ‘double ellipsodal’ heat source. The energy distribution in x,y,z is described by
the following equation:
q( x, y, z ) = Qmax . exp( A.x m + By n + Cz p )

This describes the variation of heat flux over a volume defined by:
m n p
⎛ x⎞ ⎛ y⎞ ⎛z⎞
⎜ ⎟ +⎜ ⎟ +⎜ ⎟ =1
⎝a⎠ ⎝b⎠ ⎝c⎠

Assuming that the value of heat flux q(x,y,z) decays to 0.05 of its value at x=y=z=0,
i.e. 0.05 Qmax then at x=0,y=0 and z=c
q(0,0, c) = Qmax . exp(C.c p ) = 0.05.Qmax
Therefore, approximately:

ln(0.05) 3
C= ≈−
p
c cp
A similar approach yields A & B, hence:

x y z
q( x, y, z ) = Qmax . exp(−3( ) m − 3( ) n − 3( ) p )
a b c
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 90

The value of Qmax must be determined by integrating the total energy over the volume and equating
this to the weld input energy. But unlike the ‘double ellipsoidal’ heat source which takes a
Gaussian distribution; m=n=p=2, and allows the q(x,y,z) to be integrated directly, the value of
Qmax in this case must be determined by numerical integration, as follows:

η.V .I x =0 y =0 z =c
x y z
= Qmax . ∑ ∑ ∑ exp(−3.( ) m − 3.( ) n − 3( ) p )
2 x =− a y =−bz =− c a b c

The resulting distribution is illustrated in Figure 8.19. It is sufficiently accurate in a numerical


model to consider the distribution of heat flux over a cubic volume given by: 2a x b x 2c.
Increasing the value of the coefficients m,n or p, increases the uniformity of the heat flux in the
respective direction. However in keeping with the Gaussian distribution the heat flux gradually
reduces at the edge source, which is beneficial in ensuring stability in numerical calculations.

Like the ‘double ellipsoidal’ heat source, this definition can allow for different values of c in the front
or rear segment, and incorporates a differential fraction of heat in different regions. Finally, if
required, the heat quantity directed towards the sleeve or the pipe can be adjusted. However in
the current models this has not been necessary to achieve acceptable accuracy.

The constants that define the heat source have to be arrived at by ‘trial and error’. However the
characteristics of the welding process give some guidance to the appropriate distribution. In
addition the upper bound to the pool temperature is a useful guide to the overall size of the
distribution, since it should be approximately 2400-2800°K. Comparisons with measured t8/5
cooling times also provides a useful information but perhaps the most critical test is the prediction
of fusion zone and HAZ depths since these are most sensitive to changes in the source
distribution.

After a number of trial models the shape of the heat source for E8018G vertical-down electrodes
was determined to be that shown in Figure 8.20. The flux distributions of a typical ‘double
ellipsoidal’ source and the shallow MMA welding source used here is also shown in Figure 8.21.

In a circumferential fillet weld the distribution is angled to account for deposited metal and the 90°
angle of the fillet. The width of the distribution is equal to half of the weld bead width, i.e.
a = ( weld .bead .width) / 2

and the distribution in that direction is made relatively uniform by using an exponent of m=10. It is
believed that this relates to the ‘weave’ action of the welder distributing heat evenly on the sleeve
and the pipe, and to the characteristic low penetration characteristics of the electrode type. For
E8018G electrodes, the best value for the depth of the source b, was found to be a constant 0.5
mm. This is again the result of a low penetration process with relatively large droplet transfer. For
E7016 electrodes used in the vertical-up welding position the parameter b in the heat source
definition was increased to reflect the slightly increased penetration characteristics of that
electrode type.

Another characteristic of the process is its relatively slow welding speed so there is little
justification for artificially elongating the rear of the heat source. The pool size has some
dependence on the electrode diameter, both in terms of the size of the arc and in the increased
current level used with larger electrodes. To represent this the semi-axis along the weld, c, was
chosen to equal the electrode diameter.

This heat source of this form was found to give good agreement between predicted and measured
t8/5 values, and better agreement between predicted and experimental penetration than the
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 91

conventional ‘double ellipsoidal’ heat source.

One important aspect of this heat source definition is its position relative to the joint. Goldak et al
(3) identified the importance of this in his work on in-service welding. In this work the datum for
the heat source has been taken to be the centre line of the weld bead. The angle of the source is
related to the weld bead shape as defined by the weld leg lengths on the sleeve and on the pipe.
These parameters were determined from an empirical relationship established between weld bead
geometry and the weld heat input, as described in Section 7. However since in-service welding is
a manual process it is possible that such conditions can vary. Models using varied weld bead
shape and leg lengths showed that although there was some effect on cooling times and on
penetration the magnitude was not great.

Provided the weld bead is deposited to achieve approximately an equal leg length on the sleeve
and pipe (approximately 45°) variation in model prediction will not be excessive. The measured
weld bead shapes taken from 3 experienced in-service welders supports the above position and is
in agreement with the observations made by Keifner et al (1).

Figure 8.22 shows that there is a good agreement between the predicted and experimental t8/5
values for a range of in-service welds, while Figure 8.23 shows the accuracy of prediction of weld
fusion zone penetration and HAZ depth.

8.5.2 Consideration of the Heat Losses at the Inner Pipe-wall


Both Battelle(2) and Goldak et al (3) used a non-dimensional relationship between the effective
heat transfer coefficient at the inside pipe-wall and the gas flow and pressure within the pipe. This
technique is based on the Sieder and Tate non-dimensional correlation for fully developed
turbulent gas flow (8). In this analysis the heat transfer coefficient (h) is related to a product of the
Reynolds and Prandtl numbers representing the gas flow, as follows:
0.14
h ⎛μ ⎞
= 0.023. Re 0.8 Pr 0.333 ⎜⎜ bulk ⎟⎟
C p .v.ρ ⎝ μ wall ⎠

where:
Re is Reynolds Number = ρ .v.d ,
μ
C p .μ
Pr , is Prandtl Number = k
,
k is thermal conductivity, (W/m.K),
Cp is the specific heat (J/kg.K),
ρ is the density (kg/m3),
μ bulk, μwall is the viscosity, at wall temperature or in the bulk (kg/m.s),
v is the gas velocity (m/s),
d is a characteristic length, in this case the pipe diameter (m).

Allowing for heat loss to the gas in this way simply requires specifying the appropriate convective
boundary condition in the numerical model. Accuracy of the non-dimensional approach can be as
low as ± 25% (8), however. An alternative is to expand the finite element calculation to include a
calculation of the flow field within the gas. In such calculations heat transfer from weld, through the
pipe-wall to the gas, and the convective dissipation of heat by the gas flow is intrinsically included
in the simulation. These calculations are considerably more unstable and time consuming by
comparison to the non-dimensional approach, and dealing with turbulent conditions at the pipe-
wall is a complex issue. In both methods, it is the sensitivity of the resulting temperature field to
the variation in gas flow, and its agreement with practical conditions that is important.

A number of models were used to examine the advantages of using a computational fluid
dynamics approach in order to avoid the use of non-dimensional relationships. However, it was
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 92

established that to obtain reasonable results it was necessary to artificially increase the thermal
conductivity of the gas in the region of the wall to allow for the enhanced heat exchange during
turbulent flow. This somewhat arbitrary scaling was no better than the application of the Seider &
Tate non-dimensional relationships. In addition it was established that the sensitivity of the model
outcomes to variations in gas flow and pressure was not high hence a tolerance to some
uncertainty in the heat transfer coefficient was considered acceptable.

8.6 Validation of Thermal Models by Comparison with EWI & Battelle data.

Figures 8.24, 8.25 and 8.26 show a comparison between the results predicted using a 3D quasi-
steady state model and those generated by using the Battelle model. There is some scatter in the
results but in general the agreement is reasonably good particularly in terms of the t8/5 cooling
times. This is to be expected since the t8/5 value is less dependent on the detailed description of
the heat source. There is some weak indication that the current 3D QSS models tend to predict a
lower inside wall temperature than Battelle. However the scatter in results could not be explained.

For circumferential fillet joints there is no significant difference in the accuracy with which both
models predict the t8/5 values ( Figure 8.26).

Further validation is presented in Section 9 & 10.

8.7 Extension of the 3DQSS Model to Include Multi-pass Welds and Preheat
Preheating is relatively easily dealt with provided some assumptions are made in relation to the
extent of the preheat zone. For example it was considered unacceptable to increase the overall
temperature of the pipe to a preheat temperature since this is not achieved in practice. Commonly
preheating is carried out by locally heating the weld zone with LPG torches. To approximate this
in the models the initial temperature of a band of pipe ahead of the welding arc was increased to a
preheat level. This band is limited in width and thickness and set to a fixed input temperature.
Figure 8.27 shows the temperature field that would result from such input conditions without any
welding taking place. The predicted increase in t8/5 cooling times and maximum inside wall
temperatures for various preheat temperatures is shown in Figure 8.28

Modelling of multi-pass welds is also relatively easily dealt with because it is acceptable to assume
that prior passes do not effect the starting temperature of a weld pass. Therefore, the only
difference between a model of a multi-pass weld and one of a single-pass is a change in the initial
joint configuration. As Figure 8.29 shows this is easily accounted for. The thermal history of a
fixed point in the material can be accumulated and the full thermal history can be predicted, as
shown in Figure 8.30.

8.8 Conclusions
1. Using commercial finite element software a wide range of both 2D and 3D models of in-service
welds have been developed. Stable and accurate models of all in-service weld joints has been
generated.

2. To consider the heat loss due to the flowing gas these numerical models follow the Battelle
Program and used the Seider and Tate non-dimensional approach. This determines an
effective heat transfer coefficient at the inside surface of the pipe, based on the pipe’s
diameter, gas pressure and gas flow speed. By incorporating empirical data describing weld
bead volume and joint form the representation of the welding heat input has been tailored for
both vertical-up and vertical-down welding with low-hydrogen electrodes. For circumferential
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 93

fillet welds, predictions of both t8/5 cooling times and weld penetrations have been made with
acceptable accuracy.

3. Model development has concentrated on a 3D quasi-steady state analysis of circumferential


fillet welds. In practice this is the most popular joint form, which provides a numerical analysis
that is efficient in computer calculation time.

4. Pre-processing software has been developed to efficiently construct and mesh models of
varied geometry. Post-processing software has been established to calculate the distribution
of t8/5 cooling times through out the weld zone. This can be further used to calculate a
distribution of hardness based on an empirical relationship with composition and t8/5.
Isotherms are also calculated to give an indication of the weld penetration and the extent of
the HAZ.

5. Models have mainly considered single root pass welds but some examples of multi-pass
welds have been produced.

6. The ‘double ellipsoidal’ heat source has been modified to simulate the characteristics of the
MMA welding process. Modifications have been developed to account for low hydrogen
electrodes of various diameters used in the vertical-up or vertical down welding position.

References
1. Kiefner J. F & Fischer R. D. Models Aid Pipeline Repair Welding Procedure,
Oil & Gas Journal March 1988, pp41-47.

2. Fischer R. D., Kiefner J. F. & Whitacre G. R., User Manual for Model1 & Model 2 Computer
Programs for the Predicting Critical Cooling Rates and Temperatures During Repair and Hot
Tap Welding on Pressurised Pipelines, Battelle Memorial Institute Report, June 1981.

3. Goldak J. A., Oddy, A. S. & Dorling D. V., 1993, Finite Element Analysis of Welding On Fluid
Filled Pressurised Pipelines, published in International Trends in Welding Science And
Technology, Proceedings of International conference, Trends in Welding Research,
Gatlinburgh, ed S. A. David & J. M. Vitek.

4. Goldak J. A., Chakravarti A. & Bibby M., 1984, A New Finite Element Model for Welding Heat
Sources, Metallurgical Transactions B, 15B, June, pp299-305

5. EMRC NISA Finite Element Code, Engineering Mechanics Corporation, Michigan, USA
http:\\www.emrc.com.

6. Bruce W. A. & Threadgill P. L., Effect of Procedure Qualification Variables for Welding Onto In-
service Pipelines, American Gas Association Report J7141, July 1994.

7. Cola M. J. & Threadgill P. L., Final Report on Criteria for Hot Tap Welding, American Gas
Association, Edison Welding Institute Project J7038, March 1988.

8. Holman J. P., 1976, Heat Transfer, 7th edition pub. McGraw-Hill.


CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 94

Figure 8.1

Finite element mesh for a 3D quasi-steady state model of fusion welding. Image shows finely
spaced mesh in the region of the arc, at the end of the weld bead. Material effectively flows
from right to left to simulate welding speed.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 95

(a)

(b)

Figure 8.2
Diagrammatic representation of the orientation (welding angle and trailing angle)
used to position the heat source through co-ordinate transformations at all points
throughout the welding run in the manual metal arc process.
(a) direct branch connection
(b) full encirclement or circumferential sleeve
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 96

Figure 8.3

Section through the centre of the heat source - a distributed elemental heat flux (W/mm3),
which represents the arc heat distribution during welding.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 97

Figure 8.4

A 2D section through the circumferential fillet weld. Inset shows detail of meshing
arrangement used in the weld zone.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 98

Figure 8.5
Expanded mesh in cylindrical coordinates

Figure 8.6
Further expansion and grading of element sizes
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 99

dummy weld bead

gap material

Figure 8.7

View of meshed region at the start of the weld bead showing the ‘dummy’ material introduced
in front of the welding arc and in the gap between the pipe and sleeve.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 100

Figure 8.8

A typical thermal field, for a 95 Amp, 22 Volt, circumferential fillet weld made at 2.3
mm/second. The 273 mm diameter pipe of 4.88 mm wall thickness contains methane at 4.14
MPa pressure and a flow velocity of 6.3 m/second.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 101

Figure 8.9

The t8/5 cooling times extracted from the thermal field shown in Figure 8.8 and plotted as a
contour map on a section through the mesh
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 102

Figure 8.10

The maximum temperatures in the circumferential direction extracted from the thermal field
shown in Figure 8.7. Here these temperatures are restricted to the range 720°C to 1500°C
and plotted as a contour map on a section through the mesh.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 103

Figure 8.11

Using the t8/5 cooling time, the composition of the material and the Yurioka-1 carbon
equivalent relationship the hardness can also be calculated and again shown on a section
through the mesh.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 104

Figure 8.12

The thermal field calculated for a direct branch weld


CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 105

z
a welding

b
x
c

y
y

c a
z
b
x
Figure 8.13

A schematic illustration of the necessary coordinate transformation to appropriately position


the heat source as the weld progresses around the joint.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 106

Figure 8.14

Transient models that utilise small sectors of a welded joint can reduce calculation time and
complexity in meshing.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 107

Figure 8.15

Thermal field calculated from a 3D transient heat transfer model for a single pass
circumferential fillet weld. Temperature contours set to the melting point (1500°C) and heat
affected zone (HAZ) limit (720°C) to show the extent of melting and the size of the HAZ.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 108

Figure 8.16

A 3D contour map of the t8/5 cooling times (secs) obtained from a transient model and the
segment of the modelled region which has passed through 800 to 500°C. This indicates the
variation of both start-up and stop at the end of a 150 mm weld.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 109

10

Predicted t 8/5 cooling times (secs)


3D Quasi steady state model
9
3D Transient model
8

0
0 1 2 3 4 5 6 7 8 9 10

Measured t 8/5 Cooling time (secs)

Figure 8.17

A comparison of the experimental weld cooling times measured by EWI (6) on flow loop tests
using 4.88 mm thick pipe against the predicted weld cooling times from both transient and 3D
quasi-steady state models.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 110

Gaussian distribution in x

Gaussian distribution in y
y

x z

a
cb

cf b
Gaussian distribution in z

Figure 8.18

The ‘double ellipsoidal’ heat source definition developed by Goldak et al (4).


CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 111

0.8-1
1
0.6-0.8
0.4-0.6 0.8
0.2-0.4
0-0.2 0.6

0.4

-1.225 0.2

-0.925
0 -1.25
-0.625
-0.85
-0.325
-0.45
-0.025
0.275 -0.05
y 0.575 0.35

0.875 0.75
(a) n1 = n2 = 2, i.e. a x
1.175 1.15
Gaussian distribution

0.8-1
1
0.6-0.8
0.4-0.6 0.8
0.2-0.4
0-0.2 0.6

0.4

-1.225 0.2
-0.925
0 -1.25
-0.625
-0.85
-0.325
-0.45
-0.025
0.275 -0.05
y 0.575 0.35

0.875 0.75
(b) n1= 2, n2 = 10 x
1.175 1.15

Figure 8.19

The distribution of heat flux within the heat source at z=0, (a) indicates the distribution
according to the double ellipsoidal model, (b) indicates the distribution developed in this work
to account for the weaving action of MMA welding.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 112

Figure 8.20

An illustration of the heat source distribution positioned at the front of the weld bead.

Figure 8.21

Sections through the heat distributions representing a conventional ‘double ellipsoidal’ source
(on the left) and the shallow modified Gaussian distribution used in this work (on the right), for
the same total heat input.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 113

10

9
Predicted t 8/5 Cooling Times (secs)

7
6

1 shallow non-gaussian source


shallow double ellipsoidal
0
0 2 4 6 8 10
Measured t 8/5 Cooling Times (secs)

Figure 8.22

A comparison of predicted and experimental values of t8/5 cooling times measured by EWI (6)
on 4.88 mm thick pipe.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 114

2.5

Predicted Fusion Zone Depth (mm)


modified Gaussian distribution
shallow double ellipsoid
2

1.5

0.5

0
0 0.5 1 1.5 2 2.5
Measured Fusion Zone Depth (mm)

(a)
7

6
Predicted HAZ Depth (mm)

modified Gaussian distribution


1
shallow double ellipsoid

0
0 1 2 3 4 5 6 7
Measured HAZ Depth (mm)

(b)

Figure 8.23

An indication of the accuracy in prediction of fusion zone depth and HAZ depth that can be
achieved with appropriately designed heat sources.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 115

10
9

3DQSS Predicted t 8/5 (sec)


8
7
6
5
4
3
2
1
0
0 5 10
Battelle Program Predicted t 8/5 (sec)

Figure 8.24

Comparison of the t8/5 cooling times predicted by the Battelle Program and those obtained from a
3D QSS finite element model, for a 4.88 mm thick pipe under different flow conditions.

1000
3DQSS FEA Predicted Maximum
Inner Wall Temperature(deg C)

900
800
700
600
500
400
300
200
100
0
0 200 400 600 800 1000
Battelle Predicted Maximum Inner Wall
Temperture(deg C)

Figure 8.25

Comparison of the maximum inner wall temperatures predicted by the Battelle Program and
those obtained from a 3D QSS finite element model, for a 4.88 mm thick pipe under different flow
conditions.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 116

10
Battelle Program
Predicted t 8/5 Cooling Time (sec)

9
3D QSS Results
8
7
6
5
4
3
2
1
0
0 2 4 6 8 10
Measured t 8/5 Cooling Time (sec)

Figure 8.26

The accuracy of the predicted t8/5 cooling times from the Battelle program, and from the current
3D QSS finite element models in comparison to those measured for
a 4.88 mm thick pipe by EWI (6).
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 117

Figure 8.27

The thermal field calculated for a preheat temperature of 150°C (red) imposed on the front region
of the model.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 118

14

12

Cooling time t 8/5 (secs)


10

0
0 50 100 150 200 250
Preheat Temperature (C)

1200
Maximum wall Temperature (K)

1000

800

600

400

200

0
0 50 100 150 200 250
Preheat Temperature (C)

Figure 8.28

The relationships between the preheat temperature, the t8/5 cooling times and the maximum inner
wall temperature, for a 4 mm thick pipe and a heat input of 1kJ/mm.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 119

Figure 8.29

Maximum temperatures determined on a section through the first pass (upper image) and through
the last pass (bottom image) of a multi-pass weld.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 120

1400
Initial temperature cycle on first weld
pass
1200 Second temperature cycle on second
weld pass

1000
Temperature (K)

800

600

400

200

0
0 20 40 60 80 100
Time (seconds)

Figure 8.30

An illustration of the multi-stage thermal cycle for a given point within a multi-pass weld.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 121

9. Predicting the Onset of Burn-through


9.1 Introduction
‘Burn-through’ is the term used to designate a pipe-wall failure or the bursting of the pipe-wall
during in-service welding. Considering the safety implications, there have been relatively few
attempts to systematically determine the conditions necessary to avoid pipe-wall failure during in-
service welding. Experimental work has generally used relatively small numbers of test welds
under widely varied experimental conditions, so conclusions tend to be general directions rather
than quantified limits. However, by sensible deduction, combined with some experimental
observations the qualitative causes of burn-through are known. The relevant factors are shown in
Figure 9.1.

The extensive investigation of in-service welding carried out by the Edison Welding Institute(1) and
Battelle Memorial Institute (2), have focussed on weld cooling rates, and did not experimentally
determine burn-through limits directly. Only Cassie (3) ,Phelps & Cassie (4), Wade (5,6,7), Bruce
(8,9) and Bout & Gretskii (10), have investigated the conditions which control the onset of burn-
through during in-service welding. Some of these burn-through investigations have also been
constrained by concerns for safety. For example, Wade (5) studied mechanised welds on X60
pipes statically pressurised with nitrogen and therefore did not use methane or a flowing gas.
Using a non-flowing gas would have an influence on the results, since cooling conditions at the
pipe-wall would be different to those in a natural gas pipeline. Cassie (3) and Bruce (8) also used
similar experimental systems, although some cases of manual welding on pipes containing low
pressure nitrogen were carried out by Cassie (3), who also attempted to consider a flowing gas.

Observations of burn-through generally show significant local plastic distortion of the pipe wall, and
a fracture along the weld pool axis (5,9,10). Figure 9.2 reproduces a section across a failure
presented by Bruce (9) during a weld repair and this shows the typical features.

Clearly, the risk of burn-through is related to the loss of pipe wall strength in the weld zone, and to
its inability to resist local stress, or to retain the internal gas pressure of the pipe during welding.
The reduction in wall strength depends on the elevated temperature around the weld, and on the
depth of weld penetration relative to the original wall thickness.

The consensus from industry practice and research work is that the dominant factors influencing
burn-through are,

1. Pipe wall thickness. Reduced wall thickness increases the risk of burn-through other factors
being constant. Pipe wall thickness is often used as a cut-off to safe in-service welding, with
welding not recommended below a given thickness, often in the 4-5 mm range.

2. Heat input. Increased heat input tend to increases burn-through risk. Most recommended
procedures consider specifying some upper limit at a given pipe wall thickness.

3. Internal pressure. High internal gas pressure increases the risk of burn-through and often a
pressure limit has been set for welding on thin wall pipe.

Typical recommended practices are:


• a specified lower wall thickness of about 5-6 mm below which welds should not be carried out
without significant pressure reductions,
• restricted low heat input welds (typically 0.5-1.0 kJ/mm), using controlled arc current and
welding speeds,
• the use of small diameter, low arc current, low hydrogen electrodes in the vertical-down
welding position.

A review of past work related to the investigation of burn-through has been given in Section 3 (part
3.3, 3.3.1. to 3.3.9)
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 122

9.2 Application of Current Numerical Models to Predict


Burn-through

9.2.1 Elastic-Plastic Stress Analysis


Finite element models can reliably predict the thermal field during in-service welding and this can
be combined with a thermo-elastic-plastic stress analysis to determine the deformation due to the
internal pipe pressure. However, numerical models do not directly predict the onset of burn-
through, and it is necessary to apply some limiting condition to define the onset of, or establish the
risk of, burn-through. To develop a strategy for such a model and to examine possible failure
conditions a simple model of a weld directly onto the pipe wall was considered as an extreme case
with high intrinsic risk of burn-through.

Firstly the thermal field, as shown in Figure 9.3, is calculated from a 3D quasi-steady state model
similar to those described in Section 8. This thermal field is then used in an elastic-plastic stress
analysis. The material is assumed to fit a linear stress-strain relationship with a flow stress
dependent on temperature. The analysis does not take into account any stresses developed due
to thermal expansion, since it is considered that these would not significantly influence the results.
The numerical analysis was carried out using NISA (11) finite element code. Parabolic elements
where used and a full Newton-Raphson iterative solution procedure was adopted. The temperature
field was assumed to be steady and the internal pressure in the pipe was incremented through a
range from 0 to the maximum allowable operating pressure (MAOP). Stresses and deformation
were calculated at each pressure and plotted in cylindrical coordinates. These are shown in
Figure 9.4 and Figures 9.5.

A plot of radial deflection versus internal pressure is shown in Figure 9.6. An inflection occurs in
this graph indicating a rapid increase in radial deflection above a certain pressure. This
corresponds to a rapid increase in strain in the ‘weakened’ region of the pipe as a local bulge
occurs, see Figure 9.7. The pressure limit at this inflection would provide a suitable indication of
burn-through. Alternatively the pressure to generate a fixed bulge height could be applied, which
is similar to the burn-through criteria adopted by Wade (5) (>1 mm bulge). In this case, because of
the form of the pressure versus deflection relationship, both approaches would yield similar burn-
through limits.

The same methodology can be used to examine a circumferential fillet weld with similar results.
However, in this case the model chosen has a less dramatic response since the pipe wall is at a
lower temperature and the sleeve acts to partially support the wall and limit the bulge formation.
Figure 9.8 shows some typical results.

To give some indication of the accuracy and validity of model predictions numerical models were
developed for the experimental test conditions used by Wade (5,6). The convective heat loss from
the pipe wall was estimated to be 300 W/m2.°K. Failure pressures were determined as above and
are compared with the experimental data in Figure 9.9. As can be seen, perfect agreement is not
achieved, but the general trend of the data indicates a reasonable accuracy.

9.2.2 Cavity Model


Applying a full elastic-plastic analysis to assess burn-through has the drawback that such an
analysis takes significant computer time. Whereas, an assessment of burn-through risk based
simply on temperature clearly has the disadvantage of not accounting for the influence of internal
pressure. The following approach attempts to overcome these two difficulties by using the
temperature field, and the known relationship between temperature and material strength, to
assess the effective wall strength within the weld pool region. This is similar to an approach taken
by Bout & Gretskii (10) who suggested using the 700°C isotherm as a effective cavity. Their
approach was based on the assumption that above 700°C the materials yield strength was about
4-10% of its room temperature value, and therefore had effectively zero strength.

However, this assumption is unnecessary if the strength of the wall is determined in the following
way. Firstly, the temperature field is calculated using a finite element analysis. For circumferential
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 123

welds a convenient way of carrying out this calculation is to adopt a 3D quasi-steady state
approximation.

As indicated in Figure 9.10, we firstly consider an element of material dx,dy,dz in the wall of a
pipe. This element is located in the weld region, hence the temperature of the element will vary as
its position translates from the inside (z=0), to the outside of the pipe wall (z=twall).

The force in the x direction, which can be sustained by the wall when the stress level is at yield, is
as follows:
z = twall
Fx = ∑
σ y (T ) .dy.dz
z =0
where: σ y(T ) is the yield strength of the material which depends on the temperature of the wall
and its variation along the z-direction.

If the wall were at an ambient temperature a thinner material could sustain this force since the
yield strength is lower at elevated temperatures. This effective thickness can be calculated as
follows:
z = t effective z = twall
σ y (0).t effective .dy = ∑ σ y(0) .dz.dy = ∑ σ y(T ) .dz.dy
z =0 z =0

Similar expressions would result if the force in the y-direction were considered. Hence at any point
in the pipe wall the effective thickness can be calculated if the current wall temperature distribution
and the variation of yield strength with temperature is known. By considering a number of points at
different x and y coordinates, the variation of effective thickness can be determined and viewed as
an ‘effective cavity’ in the pipe wall at an ambient temperature.

The process is illustrated in Figure 9.11, which shows a view of the thermal field calculated for a
circumferential pipe weld and the derived ‘effective cavity’, an isometric view is given in Figure
9.12, and a contour map is given in Figure 9.13.

The potential for burn-through can now be considered as equivalent to the assessment of the
pipe’s integrity with this ‘effective cavity’. A number of procedures for the assessment of corrosion
cavities have been proposed (12), and any of these could be used. In this work we have used the
modified B31G criteria (12) to determine a safe maximum pressure. This is based on a measure
of the maximum cavity depth and the maximum length of the cavity in the axial direction along the
pipe.
d
1+ 0.85.
69
SMP = MAOP.(1 + ).( t
)
SMYS 1+ 0.85. . 1
d
t M
Where:
M = Folias factor
1
1.255 L2 0.0135 L4 2
M = (1 + . − . )
2 D.t 4 D 2 .t 2
SMP= safe maximum pressure
MAOP = maximum allowable operating pressure
SMYS specified minimum yield strength
D = cavity depth
T = pipe wall thickness
L = maximum cavity width in the axial direction
D = Pipe diameter.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 124

To apply this technique to the appraisal of burn-through risk from a thermal field requires the
following steps:
• calculate the thermal field resulting from the weld,
• determine an effective cavity in the region surrounding the weld, using the known variation in
yield strength with temperature and this thermal field,
• determine the maximum effective cavity depth and its maximum width,
To facilitate the definition of the cavity edge, and hence the width, the cavity edge is defined as
the contour line at a reduction of wall thickness of 1%, as shown in Figure 9.13.

Burn-through risk can then be evaluated by comparing the current operating pressure of the pipe-
line and the calculated safe maximum pressure (SMP).

9.3 Confirmation of the Cavity Model


To confirm this concept three numerical experiments have been made, namely:

1. Calculation of bursting pressure for an idealised cavity in the pipe wall, and a
comparison with the failure values determined from the B31G criterion.

2. Calculation of bursting pressure for the quasi-steady state thermal fields of a small
number of in-service welds directly onto the surface of a thin walled pipe.

3. Comparison between published burn-through limits and calculated values.

9.3.1 Idealised cavities


A number of numerical models were created for a segment of pipe containing an idealised elliptical
cavity, 20 mm wide by 60 mm long, as shown in Figure 9.14. These are typical of the ‘effective
cavities’ calculated from thermal models of circumferential welds. The deformation was
determined from an elastic-plastic stress analysis, assuming symmetry along the cavity centreline.
Figure 9.15 and 9.16 show typical numerical results. Figure 9.17 shows the increased radial
deflection as the internal pressure is increased for a number of models for varied cavity depths.

Alternative assessments of failure pressure can be made from the relationships between pressure
and radial deflection, see Figure 9.18, or between pressure and maximum stress in the base of
the cavity, see Figure 9.19. These are:
• the pressure to obtain a 1 mm radial deflection, Figure 9.18,
• the effective yield pressure defined as shown in Figure 9.18, or,
• the pressure to exceed the yield stress at the base of the cavity Figure 9.19.

Figure 9.20 shows the relationships between the failure pressures determined from these three
methods compared with the calculated failure pressure using the B31G criterion. It is apparent
that agreement between the B31G values and the pressures derived from finite element models is
not exact, it is closest however, when using the effective yield pressure. In addition, it is clear that
the B31G criterion overestimates the failure pressure when the ratio of cavity depth to wall
thickness exceeds 0.8.

The B31G relationship is not recommended for use when the cavity depth is greater than 80% of
the wall thickness. At values greater than this the equation still predicts significant safe maximum
pressure, in fact, this pressure is still greater than zero for a cavity of 100% wall thickness, which is
clearly unacceptable. As a compromise, an adjustment to the equation has been introduced which
forces the safe pressure to approach zero as the cavity depth approaches the wall thickness, and
achieves reasonable agreement with the finite element results, as shown in Figure 9.21. By
suitable choice of function, the ratio of SMP/MAOP is identical to that calculated by B31G between
d/t ratios of 0 to 0.8. Using this approach a ‘conservative’ failure pressure can be estimated for
any ‘effective-cavity’ depth.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 125

9.3.2 Comparison with elastic-plastic thermal models


To further test this concept, the failure pressure determined for a small number of in-service welds
has been compared with the results determined using the above extended-B31G criterion.

As described in section 9.3.1 a 3D quasi-steady state model has been developed for bead-on-
pipe welds and the failure pressures are directly calculated from elastic-plastic models. As Figure
9.22 shows, these directly calculated failure pressures are larger than those determined by the
extended-B31G criterion. However, it was decided to accept the extended-B31G line as shown
because at this stage, only limited results are available, and that approach gives an additional
safety to the method.

9.3.3 Comparison with published data


There is limited experimental data to use for confirmation of this approach, but that available has
been used to validate the proposed model.

9.3.3.1 Comparison with Wade’s data and his suggested burn-through limits
Figure 9.23 shows a comparison between the experimentally determined failure pressures during
in-service welding by Wade (5) and those determined from a model using the ‘effective cavity’
approach. It can be seen that the limiting pressure calculated is considerably below the measured
values giving an indication of the level of conservatism in this approach. However, the trend in the
data is of the correct form, and given additional experimental data the numerical approach could
be fine-tuned to achieve a better agreement. For example, Wade’s failure criteria was for a local
bulge greater than 1 mm in height, which does not correspond directly with the more conservative
yield pressure used in this case.

9.3.3.2 Comparison with limits proposed by Bruce et al (8), and data from Bout &
Gretskii (10).
A comparison between the limiting conditions experimentally determined by Bruce (8) and by Bout
& Gretskii (10) are also tabulated against predicted values in Table 9.1. The results obtained were
of the correct order. However, the calculated values under estimate the measurements of Bruce et
al (8) wereas they overestimated the values suggested by Bout & Gretskii (10).

The limitations on welding heat input and pressure were expressed diagrametically by Wade (5),
and these are replotted in Figure 9.24 together with the calculated limits using the cavity model.
The pattern is in broad agreement although wereas Wade’s diagram indicates a sharp cut-off in
terms of a pressure limit the cavity model predicts welding could take place at a much reduced
pressure.

9.4 Burn-through limits for thin walled pipes


The cavity model can be used to address one of the perennial questions in relation to in-service
welding, namely, what is the safe limit for welding on pipes of less than 4 mm wall thickness.
Figure 9.25 shows the predicted rapid decline in safe pressure as the thickness is reduced. This
graph suggests that; for a pressure of 0.5 MAOP and a heat input of 0.6 kJ/mm, the lowest
thickness for safe welding would be 3.3 mm; for a heat input of 0.8 kJ/mm it would be 4.0 mm, and
for 1 kJ/mm it would be 4.7 mm. These limits are broadly in agreement with those suggested by a
number of workers.

9.5 Conclusions
It has been demonstrated that a combination of a thermal analysis and an elastic-plastic stress
analysis can be used to simulate the potential for burn-through during in-service welding. There is
limited opportunity to compare predicted values with experimental results, but such comparisons
have shown that models give a reasonable prediction of the combination of pressure and heat
input that is likely to cause a burn-through.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 126

Such elastic-plastic analyses do require significant computer time and an alternative approximate
method has been developed.

This new method uses a calculated quasi-steady state thermal field to estimate the residual
strength of the pipe wall during in-service welding. The local reduction in strength is represented
as a reduction in the thickness of the pipe wall at ambient temperature. The weld region becomes
an ‘effective cavity’ in the pipe wall. The safe maximum pressure that can be applied is then
determined from an extended-B31G criterion.

This approach has the merit of accounting for both temperature and gas pressure in determining
burn-through limits.

Where it is possible published experimental data has been compared with those predicted from the
cavity model. It has been shown that generally the model shows the correct trends, but remains
conservative in its estimation of heat input limits or limiting pressures. However, it is anticipated
that should additional experimental data become available the approach could be fine-tuned to
reduce such conservatism.

References
1. Cola M.J. & Threadgill P. L., Final Report on Criteria for Hot Tap Welding, American Gas
Association, Edison Welding Institute Project J7038, March 1988.

2. Kiefner J.F & Fischer R.D. Models Aid Pipeline Repair Welding Procedure,
Oil & Gas Journal March 1988, pp41-47.

3. Cassie B. A., The Welding of Hot tap Connections to High Pressure Gas Pipelines, paper
presented Pipeline Industries Guild J.W.Jones Memorial Lecture, October 1974.

4. Phelps B., Cassie B. A., & Evans N. H., Welding Onto Live Natural Gas Pipelines, Metal
Construction, August 1976, pp350-354.

5. Wade J.B., Hot Tapping of Pipelines, Australian Welding Research Association Research
Report, Snowy Mountains Corporation1978.

6. Wade J.B., Effect of Preheat on Hot Tapping Procedures, Australian Welding Research
Association Research Report, Snowy Mountains Corporation, September 1978

7. Wade J.B., Description of Experimental Results on the Effects of Pipeline Damage on


Performance and Hot Tapping Techniques, paper presented at Australian Welding Research
Association’s conference Pipeline Welding in 80’s, Melbourne March 1981, Paper 4a.

8. Bruce W. A., Holdren R.L., Mohr W. C., Kiefner J.F. & Swatzel J.F., Repair of Pipelines by
Weld Metal Deposition, Paper presented at PRCI 9th Symposium on Pipeline Research,
Houston Texas, September 1996.

9. Bruce W. A., Holdren R.L. & Mohr W. C., Repair of Pipelines by Direct Deposition of Weld
Metal – Further Studies, Final report Edison Welding Institute, EWI Project J7283, November
1996.

10. Bout V.S. & Gretskii Yu.Ya., Arc Welding Application on Active Pipelines, Pipeline Technology,
Volume 1, R.Denys Ed. R.Denys, pub Elsevier Science BV. pp550-558.

11. EMRC NISA Finite Element Code, Engineering Mechanics Corporation, Michigan, USA
http:\\www.emrc.com.

12. Kiefner J. F., & Vieth P.H., Rstrength2 Users Manual, American Gas Association Final Report
PR-218-9205, March 1993.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 127

Arc current

Arc voltage

Heat Input Welding speed

Electrode polarity
Weld Penetration Electrode type

Pipe Wall Temperature Electrode diameter

Welder technique/direction
Local Pipe Wall Strength

Pipe wall thickness

BURN THROUGH Preheat temperature


Pipe Wall
RISK
Cooling Gas flow rate

Gas temperature

Pipe diameter
Local Gas pressure
Applied Stress
Yield strength at temperature

Weld orientation

Local wall support

Figure 9.1
The parameters influencing the possibility of burn-through during in-service welding.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 128

Figure 9.2

A macro-section taken through the centre of a burn-through, after Bruce et al (8)


CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 129

Figure 9.3

The calculated thermal field for a 1 kJ/mm heat input weld on a 5 mm thick X70 pipe
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 130

Figure 9.4

The radial deformation predicted by an elastic-plastic stress analysis: deformation plotted as a


contour map on the distorted mesh. This is for a 1kJ/mm heat input weld on a 5 mm thick X70
pipe, at a pressure of 10 MPa.

Figure 9.5

The circumferential stress field in the region of a in-service weld for a 1 kJ/mm heat input weld on
a 5 mm thick X70 pipe, at a pressure of 10 MPa.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 131

internal pressure (MPa) 2.5

1.5

0.5

0
0 0.2 0.4 0.6 0.8 1 1.2

radial deflection (mm)

Figure 9.6

A typical variation of radial deflection at the centre of the weld zone as the pressure in the pipe is
increased.

3
internal pressure (MPa)

2.5

1.5

0.5

0
0 0.005 0.01 0.015 0.02

plastic strain

Figure 9.7

A typical variation of plastic strain at the centre of the weld zone as the pressure in the pipe is
increased.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 132

(a)

(b)

(c)

Figure 9.8
Example of results from a stress analysis of a circumferential fillet weld.
(a) Radial deflection concentrated in the weld region
(b) Localised plastic strain in the weld zone
(c) Variation of plastic strain with increased pressure 0-10 Mpa on this 1 kJ/mm weld on 5 mm
thick pipe.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 133

1.8

1.6
Weld Heat Input (kJ/mm)

1.4

1.2

0.8

0.6 Wade results, bulge height < 1.0 mm


Wade results, bulge height > 1mm
0.4 Wade results, burst
Predictions for h=300 W/m2.K
0.2 Predictions for still air

0
0 2 4 6 8 10 12
Pipe Internal pressure, Nitrogen gas (MPa)

Figure 9.9

Comparison between the bursting pressures determined by finite element models and the
experimental results determined by Wade (5).
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 134

x
dy

twall
dz
Fx

Figure 9.10

Schematic view of an element within the pipe wall, together with the forces applied.

Figure 9.11

An example of the thermal field used to determine an ‘effective cavity’, by considering the
temperature and the effective material strength along A-A
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 135

Figure 9.12

A perspective view of the effective cavity created in a 3 mm thick pipe welded with a heat input of
0.75 kJ/mm.

Minimum thickness

Figure 9.13

A plan view of the effective cavity created in a 3 mm thick pipe welded with a heat input of 0.75
kJ/mm. The width and depth of the cavity are measured as indicated.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 136

Figure 9.14

An idealised elliptical cavity in the wall of a pressurised pipe.

Figure 9.15

Contour map showing the radial deflection concentrated in the centre of the depression in the pipe
wall.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 137

Figure 9.16

Extreme radial deflection, as the pipe reaches the failure pressure.


1.4

1.2
Internal Pressure / MAOP

1.0

0.8

0.6
d/t = 0.33
d/t = 0.66
0.4 d/t = 0.75
d/t = 0.83
d/t = 0.95
d/t = 0.96
0.2
d/t = 0.97
d/t = 0.98

0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Radial deflection of cavity base (mm)

Figure 9.17

The calculated variation of radial deflection at the base of a pipe wall cavity. For a 300 mm
diameter, 3mm thick, X70 pipe containing an idealised cavity 20 mm wide by 60 mm long
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 138

1.2

1.0
‘Yield’
Internal Pressure / MAOP
Effective ‘yield’
0.8
pressure

0.6

0.4

d/t =0.83
0.2

0.0
0.00 0.10 0.20 0.30 0.40 0.50 0.60 0.70 0.80 0.90 1.00

Radial Displacement of Cavity Base (mm)

Figure 9.18

The derivation of a ‘pressure at radial yield’ from the data in Figure 9.18

1.8
d/t = 0.33
d/t = 0.66
1.6
d/t = 0.75
1.4 d/t = 0.83
Internal Pressure / MAOP

d/t = 0.92
1.2 d/t = 0.95
d/t = 0.97
1.0 d/t = 0.98

0.8

0.6

0.4

0.2

0.0
0.0 100.0 200.0 300.0 400.0 500.0
Equivalent Stress at Cavity Base (MPa)

Figure 9.19

The variation of equivalent stress at the base of an idealised cavity with changes in internal
pressure in a 300 mm diameter pipe of 3 mm thick X70 containing an idealised cavity 20 mm wide
by 60 mm long.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 139

1.6
B31G Criteria
FEA 1 mm deflection
1.4
FEA 'deflection' yield
FEA equivalent yield stress
1.2 Cavity minimum thickness
Internal Pressure / MAOP

0.8

0.6

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1 1.2

Cavity Depth / Pipe Wall Thick ness

Figure 9.20

The predicted failure pressures for a 300 mm diameter, 3 mm wall thickness X70 steel pipe
containing various idealised cavities of varied depth to pipe wall thickness ratios, calculated by
FEA using various failure criteria, and by B31G analytic equation and by an equation based on
remaining wall thickness.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 140

1.4

1.2
Internal Pressure / MAOP

0.8

0.6

B31G + modification
0.4
B31G Criteria
FEA Pressure at deflection yield
0.2
Pressure at equivalent yield stress

0
0 0.2 0.4 0.6 0.8 1 1.2
Cavity depth / Pipe Wall Thickness

Figure 9.21

A comparison between the predicted failure pressures from a modified B31G criterion and the
failure pressures calculated by FEA for a 300 mm diameter, 3 mm wall thickness X70 steel pipe.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 141

1.2

1
Internal Pressure / MAOP

0.8

0.6

0.4 B31G + modification

B31G Crit eria


0.2 FEA E lastic P last ic m odel

0
0 0.2 0.4 0.6 0.8 1 1.2
C avity depth / Pipe Wall Thickness

Figure 9.22

Comparison between the failure pressures predicted by the B31G criteria, the modified B31G line
for d/t > 0.8 and a limited number of values from elastic-plastic stress analysis.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 142

14

12

Bursting Pressure (MPa)


10

2 Predicted from cavity model


Measured by Wade - Burst
Measured by Wade - bulge > 1 mm
0
0 0.5 1 1.5 2
Heat Input (kJ/mm)

Figure 9.23

A comparison between the predicted bursting pressure calculated by the cavity model and the
experimental values determined by Wade (5) for a 5 mm thick pipe

Table 9.1 Comparison of experimentally determined critical levels of heat input to cause burn-
through and values determined from the cavity model.

Source Wall Gas Pressure Critical Predicted


Thickness Heat Input Heat Input
(mm) (MPa) (kJ/mm) (kJ/mm)
Bruce et al (8) 3.2 * 6.2 0.54 0.41

4.0 * 6.2 0.78 0.69

Bout & Gretskii (10) 3.0 4.0 0.48 0.50

3.0 3.0 0.51 0.54

* data determined during weld repair procedure testing therefore this is remaining thickness in
restricted region.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 143

2.5
3 mm w all thickness
5 mm w all thickness
6 mm w all thickness

2 6 mm limit
Heat Input (kJ/mm)

5 mm
1.5

0.5

3 mm

0
0 2 4 6 8 10
Internal Pressure (MPa)

Figure 9.24

A comparison between heat input limits before burn-through, as calculated by the cavity model
and as proposed by Wade (5) from his experimental work. Wade represented a safe zone as a
rectangular region limiting both heat input and pressure.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 144

1.2

1
Bursting Pressure/ MAOP

0.8

0.6

0.4 0.6 kJ/mm


0.8
1
0.2

0
2 3 4 5 6 7
Pipe Wall Thickness (mm)

Figure 9.25

The predicted bursting pressure for various thickness pipes welded with 0.6,0.8 and 1 kJ/mm
heat input. Pipe diameter 373 mm, and heat transfer coefficient of 300 W/m2.K.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 145

10. Model Validation Using the Gladstone Flow Loop

10.1 Background

The purpose of the trial was to obtain test data from in-service welds to compare with predictions
from numerical models. Much of the experimental data related to in-service welding comes from
simulated laboratory work using water as a coolant, and this does not allow a practical test of our
model predictions. Some flow loop data reported by EWI, is not complete, and it only includes weld
cooling rates, without other important parameters like the weld penetration or heat affected zone
(HAZ) hardness.

A flow loop test clearly includes the gas flow and heat loss due to flowing methane, and is the
closest possible simulation of operational conditions. Also it importantly provides specimens of the
finished weld for metallographic work and hardness measurement.

In February 1999 the management of Duke Energy gave their approval for the project to undertake
a welding trial at their Gladstone Gate facility. This trial was to consist of a number of
circumferential welds on a specially constructed length of 300 mm diameter pipe under operational
conditions.

The objectives of the trial were as follows:

1. To measure the arc energy, arc current and voltage for a range of in-service welding conditions.
2. To measure the welding speed.
3. To measure the thermal cycle for each weld from which t8/5 cooling times could be extracted e.g.
( the time in seconds to cool between 800° and 500°C).
4. To take a metallographic section of each weld and measure it’s:
• weld penetration
• hardness profile.

5. To then numerically simulate those welds and compare the predicted results with the measured
data. This simulation was to take two forms:
a) To directly model the weld conditions and compare predicted weld geometry, weld cooling
times and hardness with the measured data.
b) To utilise the ‘database’ approach for prediction of cooling times and HAZ hardness,
described in Section 11, and verify its acceptability.

10.2 Test Material


BHP (John Piper) kindly supplied the test pipe. This consisted of two, 3-metre lengths of 4.8 mm
and 5.6 mm thick 300 mm diameter X70, together with a length of 6.4 mm thick Ultrapipe. The
chemical compositions of these materials are given in Table 10.1. A length of 9 mm thick Ultrapipe
was also provided. This was saw-cut into 100 mm wide strips to be used as simulated sleeve
sections.

10.3 Test Description


The test weld consisted of a single root-bead, circumferential sleeve fillet weld, between top-dead-
centre 0° and bottom-dead centre 180° of the 300 mm diameter run pipe.

2.5 mm and 3.2 mm diameter E8018G, electrodes were used to produce single root, vertically-
down, circumferential welds. The weld preparation was 90°, achieved by a saw cut edge on the
simulated sleeve segment.

Welding was carried out using both the upstream, mainline pressure of approximately 5.5-6 MPa
and the reduced, downstream pressure of 2.5-3.0 MPa. Duke Energy recorded the flow and
pressure conditions at 1 minute intervals throughout the test period.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 146

A schematic diagram of the test procedure is given in Figure 10.1. For each weld the arc current
and voltage were recorded using a weld monitor, which logged data at a rate of 10 readings per
second. Previous work at AGL as described in see Section 7, had identified that the weld heat input
fluctuated due to variation in welding speed, so in these trials particular emphasis was placed on
recording welding speed. However in a manual operation continuously recording the welding speed
is clearly difficult. Hence a new technique has been employed. This involved recording the time at
which the weld passed a series of points made at 25 mm spacing on the surface of the pipe. This
was facilitated by recording a voltage signal on an additional channel of the weld monitor, and
manually switching the voltage recorded as the weld passed each 25 mm mark. The longitudinal
spacing of the resulting ‘square’ waveform allowed the calculation of welding speed. Weld cooling
cycles were determined from thermocouples which were ‘harpooned’ into the weld pool.

Forty-five test welds were carried out on the 9th. March, 1999, at Duke Energy’s Gladstone Gate
facility. Photographs of the site with the test pipe used and the test welds are shown in Figures 10.2
and 10.3.

10.4 Test Data Collection


The pipe was returned to CSIRO Manufacturing Science & Technology (CMST) in Adelaide where it
was cut up and the data collated.

For each weld a graph of heat input variation with distance was generated, similar to the examples
shown in Figure 10.4. From the cooling curves, see Figure 10.5 for examples, the t8/5 time was
taken. Harpooning thermocouples into a weld pool is not an easily controlled technique so there
was considerable scatter and some erroneous results were obtained.

For each weld, a metallographic section was taken near to the region where the thermocouple was
placed, which was approximately in the centre of the weld length. A photographic record of all
welds was taken (see Figure 10.6 for an example), and the dimensions of the weld and it’s HAZ
and fusion zone recorded.

To assess the HAZ hardness a line was established approximately in the centre of the weld bead,
perpendicular to the pipe surface (see Figure 10.6) and the hardness determined along that line
using indentations at approximately 0.5 mm spacing. Plots of hardness profile were established
and maximum hardness within the HAZ assessed.

A summary of all the test data is presented in Table 10.2, and a complete set of test results for each
weld is given in Appendix 10.1.

10.5 Data Analysis

10.5.1 Variability of Heat Input


There are two potential concerns relating to the heat input. These are:
• it’s general variability,
• the potential variation with position around the circumferential joint.

Previous laboratory work at AGL (Section 7) had identified considerable variation in heat input
during manual welding, but it had not detected any systematic variation around the welded joint.

The variation in heat input has been seen in Figure 10.4 and it is further shown in Figure 10.7
which simply plots the maximum and minimum heat input recorded for each weld. In each data set,
points are for welds with increasing angular position, with the majority of welds requiring three weld
runs to go from top-dead centre to bottom-dead centre of the pipe. It is apparent that there is
considerable variation in heat input, even over this relatively short weld length. There is no
systematic variation in heat input with different weld positions.

A relatively conservative estimate of the range of variation is ±15% of the nominal value but there
are cases where the maximum heat input is about 1.5 times the nominal.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 147

The variation in heat input for all welds is also presented in Appendix 10.1

10.5.2 Bead Shape


Numerical models require an estimate of the bead shape and size in order to accurately represent
the welding heat input and achieve the correct heat source orientation. This has been addressed in
this work by utilising simple linear relationships between heat input and the weld leg lengths. The
additional field data gathered in the Gladstone trial allowed the previously established relationships
to be further tested.

Figure 10.8 shows the relationship between the pipe leg length, the sleeve leg length, the deposited
area and the heat input, and compares the field data with the relationships established from the
AGL tests. Both the AGL laboratory tests and the results from the field trials showed some
considerable scatter, and because of this there is no conclusive evidence that the deposited area is
significantly different. There is however, some evidence that in the Gladstone trial the pipe leg
length was generally smaller, as indicated in Figure 10.8(b). The average ratio of the pipe leg
length to the sleeve leg length in the AGL tests was 0.97±0.05 whereas for the Gladstone trial it was
0.85±0.04. This variation is in weld angle or in the balance of heating between the sleeve and the
pipe. For the current trial there was an increased tendency for a longer weld on the sleeve side.

The data suggest that there will exist variation in ‘welder’ technique which will impact on bead
shape, however it is difficult to take this into account and the general scatter and natural variation in
results probably makes it unnecessary. Since in numerical models the leg lengths of the weld bead
control the angle of the heat source, the above variation would correspond to an average weld angle
of 46° to the sleeve, for the AGL welds, compared with an average angle of 50° for the Gladstone
trial. Although this is a small variation; in the Gladstone trials, it is sufficient to reduce the
penetration into the run-pipe.

10.5.3 Hardness Variation


The relationship between the weld cooling time, the pipe composition, and the resulting HAZ
hardness is also relevant to the development of a numerical simulation, since it is necessary to
deduce the HAZ hardness from the calculated weld cooling time.

There are a number of possible ways to evaluate the ‘best’ empirical relationship. Figure 10.9
compares a plot of measured hardness and t8/5 cooling times with three empirical relationships. It
illustrates that the trend in experimental data is clearly different to that suggested by the
relationships. It is clear that using the Yurioka empirical equations will overestimate hardness at all
cooling times, whilst the Beckert equation will generally underestimate the hardness at long cooling
times and overestimate hardness for low cooling times.

In an earlier assessment of AGL water cooled experiments (see Section 7) the mean absolute error
was obtained between the measured hardness and that calculated by a number of empirical
relationships. Table 10.3 compares the results from that work with the current Gladstone test data.
For the AGL tests the best agreement between predicted and measured maximum HAZ hardness
was found with the Yurioka-1 relationship. Applying the same procedure to the Gladstone results
the mean absolute error is 44 ± 7 VHN for Yurioka-1, and the best agreement is obtained with the
Beckert equation which gave a mean absolute error of 25 ± 7 VHN. The reason for this change in
behaviour is not clear. As indicated in Figure 10.10 the water-cooled AGL tests tended to produce
faster cooling times, and included pipe compositions of lower strength (X60) with higher carbon
levels, so this may explain the observed changes.

Finally the data can be considered as a plot of predicted hardness based on the measured t8/5
cooling time versus the experimentally measured hardness as shown in Figure 10.11. This gives
groups of data points related to three empirical equations. Whilst all equations have some error the
best general outcome from this would appear to be the Yurioka-2 equation, which generally gives an
over-predicted hardness value.

The argument in favour of continued use of the Yurioka-1 relationship is that it will produce a
conservative outcome: e.g. a required cooling time calculated from Yurioka-1, based on say 300
VHN, would need to be greater than 5 seconds, but from the experimental data this would produce
about 250 VHN.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 148

10.6 Testing Numerical Models

Numerical models of test conditions were generated to allow full and rigorous assessment of their
accuracy. This was achieved in two ways,

a) Individual numerical models were created using the measured bead shape, heat input and
welding conditions.
b) The ‘data-base’ approach described in Section 11 was also used to predict weld cooling times
and HAZ hardness.

10.6.1 Direct Numerical Simulation


Circumferential sleeve models were generated and used to predict weld cooling times, and weld
geometry, and pipe wall temperatures. Figure 10.12 shows typical predicted thermal fields from
which the relevant test data is extracted. Table 10.4 summarises the predicted values.

10.6.1.1 Comparison of calculated and measured weld cooling times


Figures 10.13 and 10.14 show the comparison between the predicted t8/5 and the experimentally
measured values. Simulations were carried out using the minimum, maximum and average
measured heat input. This generates a range of predicted t8/5 values, as indicated in Figure 10.14.
If the data points fall close to the 45° line across the centre of the graph this signifies agreement
between predicted and measured values. This does not always occur and it is clear that there is
significant scatter in the final results. Generally the agreement is reasonable for 4.8 mm thick pipe,
but for 6.4 mm there is a tendency for the numerical models to under estimate t8/5 cooling times
above 5 seconds.

Since the t8/5 cooling time occurs over a period of time during which the instantaneous heat input will
vary it is anticipated that the predicted value should be judged by comparing the results calculated
from the average value of heat input. This approach is used in Figure 10.13, where the maximum
and minimum t8/5 cooling times for the region of the weld within the pipe wall are plotted against their
corresponding measured values. For good agreement we may expect that the line of equality would
pass between the maximum and minimum values. This does not occur cleanly, but the trend is
established. Again it can be seen that there is a tendency for the numerical models to under-predict
t8/5 values for the thicker material and larger heat inputs.

With 90% confidence the mean error between experimental and predicted average t8/5 values is
between 4 and 16%, ie. the numerical models tend to under-predict t8/5 cooling times.

10.6.1.2 Comparison of the predicted and measured HAZ depth


Comparison between the predicted and measured HAZ depths is shown in Figure 10.15 and 10.16.
Figure 10.15 considers gives predictions for the minimum, average and maximum heat input
values, and plots these values against the measured value. Agreement is generally good for the 5.6
and 6.4 mm thick pipe. For 4.8 mm pipe the results are more scattered but still show reasonable
agreement. The increased scatter in 4.8 mm thick pipe is a consequence of the greater sensitivity
to heat input variation for welds on thinner walled pipes. Figure 10.16 compares the HAZ depths
predicted using the average heat input and shows a reasonable predictive capability, although as in
other sets of data, some scatter exists.

With 90% confidence the mean error between experimental and predicted HAZ depths, based on
the average heat input was between –18 and 0%, ie. the numerical models has a tendency to over
estimate HAZ depth.

10.6.1.3 Comparison of predicted and measured hardness


From the calculated t8/5 and the known composition of the steel pipe the hardness can be predicted.
Figure 10.17 shows the comparison between the measured hardness and the value calculated
using two empirical equations, that due to Beckert and Yurioka-1. Again there is scatter in the
results but reasonable agreement is achieved using the Beckert equation, whilst the Yurioka-1
relationship tends to over-predict the hardness level.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 149

With 90% confidence the mean absolute error between predicted hardness and measured values
when using Yurioka-1 is is between 20-30%, and for the Beckert equation between 8-16%.

10.7 Numerical predictions using the ‘hot-tap data-base’


The experimental data derived from the Gladstone experiment also allowed the opportunity to test
the hot-tap database described in Section 11. This was achieved by simply entering the test
conditions, pipe geometry, composition, gas flow rate and pressure in the ‘database’. This then
calculates the acceptable working window of heat input for a range of flow conditions, as illustrated
in Figure 10.18.

In addition selected heat inputs can be entered and the database used to calculate, t8/5 cooling
times and the HAZ hardness. In this case the average measured heat inputs were used and the
calculated values compared against the measured data to provide a test of the software’s
capabilities.

Figure 10.19 shows the prediction of t8/5. Whilst it shows some scatter this is certainly no worse,
and possibly a little better, than that obtained from a direct calculation through individual models.
So the database approach has not compromised the accuracy of prediction.

From Figure 10.20 it can be seen that there is a general tendency to over estimate the HAZ
hardness. This is due to the choice of empirical equation used within the database. This was set to
be the Yurioka-1 relationship based on the findings of earlier AGL experiments. However as
discussed (section 10.5.3) this was not supported by the Gladstone trial and better agreement
should be achieved with the Beckert equation. The result of using Yurioka’s equation in the
database would be to over estimate the lower heat input limit required to achieve a given hardness
thus restricting the operational window. To avoid this perhaps more attention has to be given to the
choice of composition-t8/5-hardness relationships.

The database calculations also use a standard empirical equation to determine deposited area and
weld bead leg lengths. As previously discussed, section 10.5.2, the Gladstone welds have shown
some variation from this. It is not considered that this would be significant to the prediction of weld
cooling times. However it may be necessary to take such variations into account for the calculation
of ‘effective cavity depth’ and burn-through limits since these will depend on the depth of penetration
which is sensitive to weld angle. Since the burn-through limits are not directly tested in these
experiments it is impossible to test this aspect of the database with the present data.

Further testing of the database concept is presented in Section 11.


CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 150

Monitor arc current


arc voltage & time Temperature from
harpooned thermocouple
Trigger
time
Weld length

pressure & gas flow


gas composition

Figure 10.1
Schematic layout of the Gladstone weld test.

Table 10.1

Chemical Analysis of the steel pipes used in the Gladstone tests.

%C %Si %Mn %Cu %Ni %Cr %Mo %V %B %Nb %Ntot %Ti


4.8 mm thick 0.065 0.34 1.42 0.014 0.029 0.023 0.107 0.005 0.001 0.076 0 0.02
X70
5.6 mm thick 0.065 0.31 1.32 0.01 0.024 0.011 0.106 0.005 0.001 0.062 0 0.019
X70
6.4 mm thick 0.137 0.12 1.03 0.018 0.031 0.009 0.01 0.005 0.001 0.005 0 0.005
Ultrapipe
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 151

Figure 10.2
Photograph of the test site at Gladstone, Queensland with the flow loop and test pipe in place.

Figure 10.3
Test welds completed on one section of test pipe.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 152

1.20

1.00
Heat Input (kJ/mm)

0.80

0.60

0.40

0.20
Test 31, 4.88 mm thick, X70

0.00
0 10 20 30 40 50
Time (secs)

1.20

1.00
Heat Input (kJ/mm)

0.80

0.60

0.40

0.20
Test 33, 4.88 mm thick
X70
0.00
0 5 10 15 20 25 30 35
Time (seconds)

Figure 10.4
Typical examples of the variation of instantaneous heat input during a single weld run.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 153

1800

1600 Thermocouple 1-t8/5=5.4secs


1400 Thermocouple 2-t8/5=4.7secs
Temperature (deg C)

1200

1000

800

600

400

200

0
1000 1100 1200 1300 1400 1500 1600 1700 1800

Time (0.1 seconds)

1800
Thermocouple 1- t8/5=4.2 secs
1600
Thermocouple 2 -Failed
1400
Temperature (deg C)

1200
1000

800
600
400
200
0
300 400 500 600 700 800
Time (0.1 seconds)

Figure 10.5
Typical examples of the temperature traces recorded from harpooned thermocouples. Non-smooth
recordings and occasional failures due to the thermocouple melting were not uncommon.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 154

Outer surface of pipe

Inner surface of pipe wall

Hardness measured along this line

Figure 10.6
An example of a polished and etched section through the weld. Sections were taken at
approximately the mid position of the weld runs and harness measured along the a line in the
middle of the weld beads.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 155

Table 10.2 A summary of all the experimental results for the Gladstone test welds

Weld Pipe Pressure Gas Flow Heat Input Heat Input Heat Input Experimental Weld Weld Pipe Deposited HAZ HAZ Depth Fusion
Number Thickness (minimum) (average) (maximum) T85 C Sleeve leg Leg Length Area Hardness Depth
length
mm MPa m/sec KJ/mm kJ/mm kJ/mm seconds mm mm mm^3 VHN mm mm
1 6.4 3.46 6.79 0.77 0.83 0.9 4.5 4.87 3.97 9.65 250 2.4 0.7
2 6.4 3.47 6.77 0.76 0.9 1.05 3.19 5.29 4.73 12.5 223 2.3 0.8
3 6.4 3.46 6.56 0.71 0.79 0.88 3.8 5.84 4.73 13.82 222 2.3 0.6
4 4.8 3.49 6.74 0.81 0.93 1.09 4 5.08 4.17 10.6 242 3.6 1
5 4.8 3.49 6.72 0.66 0.75 0.84 5.65
6 4.8 3.5 6.7 0.81 0.94 1.42 4.2 5.63 4.24 11.96 262 2.9 0.7
7 5.6 3.43 7.83 0.69 0.84 0.98 4.47 5.32 3.78 10.05 238 2.5 0.7
8 5.6 3.42 6.7 1.17 1.29 1.43 5.6 4.34 12.15 235 2.5 1
9 5.6 3.42 6.68 0.88 1.09 1.23 3.6 5.74 4.48 12.86 241 2.4 0.6
10 6.4 3.43 5.89 0.83 0.94 1.03 5.57 4.03 11.23 226 2.4 0.8
11 6.4 3.45 6.83 0.59 0.79 1.02 5.46 6.54 5.5 17.9 185 2.4 0.6
12 6.4 3.45 6.85 0.87 1.02 1.2 5.2 5.29 4.31 11.4 281 1.8 0.4
13 6.4 3.43 6.88 0.86 0.98 1.2 4.5 5.01 5.36 13.41 227 2.5 0.8
14 4.8 3.45 6.63 0.87 1.04 1.48 4.8 5.63 4.94 13.92 244 2.6 0.6
15 4.8 3.45 6.63 0.99 1.06 1.17 5.43 5.08 13.78 3.2 0.6
16 4.8 3.45 6.63 0.83 1.18 1.3 6.5 6.26 4.87 15.24 239 2.9 0.6
17 4.8 3.45 6.63 0.86 0.98 1.12 6.12 4.45 13.63 253 3.3 0.3
18 5.6 3.44 6.61 0.99 1.04 1.12 6.35 6.16 5.18 15.95 244 3.2 0.8
19 5.6 3.44 6.64 0.9 1.04 1.15 7.14 5.46 19.49 237 2.3 0.1
20 5.6 3.44 6.64 1.14 1.23 1.31 5.3 6.02 5.88 17.7 240 2.7 0.6
21 5.6 3.44 6.65 0.86 1.17 1.36 5.35 6.16 5.04 15.52 233 2.9 0.4
22 6.4 3.45 6.07 1.01 1.14 1.2 5.38 5.57 4.94 13.74 223 2.8 0.9
23 6.4 3.42 6.15 0.93 1.04 1.14 5.93 5.84 5.15 15.04 232 1.8 0.4
24 6.4 3.44 6.84 0.83 1.02 1.36 4.2 5.57 5.01 13.94 188 2.4 1.1
25 5.6 3.44 6.82 0.88 1.13 1.47 5.7 6.58 5.04 16.58 232 3.4 0.8
26 5.6 3.44 6.4 0.99 1.03 1.09 6.22 5.81 5.81 16.88 237 2.6 0.7
27 5.6 3.43 6.42 0.64 0.91 1.08 4.8 5.84 5.32 15.27 240 3.2 1.3
28 6.4 5.83 3.89 0.91 0.98 1.04 7.2 4.52 4.59 10.38 287 2.9 1
29 6.4 5.83 3.88 1 1.07 1.12 5 5.29 4.38 11.59 213 1.9 0.1
30 6.4 5.83 3.89 0.78 0.88 1.04 4.7 5.08 3.9 9.89 202 1.7 0.6
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 156

Table 2: A summary of the experimental results for the Gladstone Test Welds: Continued

Weld Pipe Pressure Gas Flow Heat Input Heat Input Heat Input Experimental Weld Weld Pipe Deposited HAZ HAZ Depth Fusion
Number Thickness (minimum) (average) (maximum) T85 C Sleeve leg Leg Length Area Hardness Depth
length
mm MPa m/sec KJ/mm kJ/mm kJ/mm seconds mm mm mm^3 VHN mm mm
31 4.8 5.83 3.88 0.7 0.94 1.03 5.2 5.7 4.66 13.29 242 3.8 0.6
32 4.8 5.83 3.88 0.87 0.94 1.16 5.8 4.66 4.73 11.02 254 3.1 0.7
33 4.8 5.84 3.88 0.55 0.87 0.98 4.2 5.36 3.48 9.32 247 2.3 0.63
34 5.6 5.8 3.88 1.2 1.52 1.78 4.63 3.92 9.06 234 3.5 0.7
35 5.6 5.84 3.89 0.83 0.99 1.15
36 5.6 5.84 3.89 1.04 1.14 1.23 4.5 6.09 4.41 13.43 239 2.5 0.7
37 6.4 5.86 3.89 1.38 1.56 1.7 5.6 5.5 4.73 13 284 3.1 0.8
38 6.4 5.86 3.89 1.02 1.15 1.29 5.84 4.66 13.62 251 1.8 1
39 6.4 5.84 3.89 0.95 0.98 1.12 4.2 5.15 4.66 12 246 1.7 0.5
40 5.6 5.85 3.89 0.98 1.1 1.17 6.73 5.04 16.93 233 3 0.7
41 5.6 5.85 3.89 0.99 1.13 1.31 6.88 4.34 5.18 11.24 230 3 0.6
42 5.6 5.86 3.89 0.94 0.97 1 4.9 5.67 13.89 244 3 0.7
43 6.4 5.85 3.89 0.91 1.2 1.81 7.5 7.23 5.15 18.62 210 2.6 0.6
44 6.4 5.85 3.9 1.31 1.57 1.65 6.96 6.68 23.23 319 2.7 0.4
45 6.4 5.85 3.9 1.28 1.46 1.62 8 7.93 6.82 27.03 265 2.4 0.4
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 157

2
1.8
1.6
H eat Input (kJ/mm)

1.4
1.2
1
0.8
0.6
0.4
Minimum Heat Input Maximum Heat Input
0.2
0
0 10 20 30 40 50 60
Weld Number

Figure 10.7
A summary of the variation of minimum and maximum heat input measured for each weld run. Each grouping represents one circumferential weld with
one pair of points for each weld run, which progressed from top-dead-centre to bottom-dead-centre (left to right within each group).
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 158

25

20

Weld Bead Area (mm^3)


15

10

5
Gladstone data
Relationship from AGL tests
0
0 0.5 1 1.5 2

Average Heat Input (kJ/mm) 10.8(a)

8
Weld Pipe Leg Length (mm)

2 Gladstone data
Relationship from AGL tests
1 Data points from AGL tests

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6 10.8(b)
Average Heat Input (kJ/mm)

8
Weld Sleeve Leg length (mm)

3
10.8(c)
2 Gladstone data
Relationship from AGL tests
1
Data points from AGL tests
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
Average Heat Input (kJ/mm)

Figure 10.8
A comparison between the empirical relationships determined to describe the weld bead
shape for E8018G vertical-down electrodes from laboratory tests at AGL (see section 7) and
the data generated from the Gladstone field trials.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 159

400

350

300
Vickers Hardness

250

200

150

100 Gladstone 6.4 Ultrapipe


Beckert Relationship
Yurioka -1
50 Yurioka-2

0
0 1 2 3 4 5 6 7 8 9

Experimental Weld Cooling Time t 8/5 (sec)

400

350

300

250
Vickers Hardness

200

150
Gladstone 5.6 mm X70
Gladstone 4.8 mm X70
100
Beckert Relationship
Yurioka -1

50 Yurioka -2

0
0 1 2 3 4 5 6 7 8 9

Experimental Weld Cooling Time t 8/5 (sec)

Figure 10.9
Comparison between the relationship between measured hardness and measured t8/5,
compared to that predicted by various empirical equations.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 160

Table 10.3

Comparison between the average errors determined for the prediction of hardness based on
measured t8/5 cooling times in the AGL tests (see Section 7) and in the current Gladstone trial.

Mean Absolute Error between Measured Hardness and Predicted


Hardness

VHN
Empirical Equation AGL-Tests Gladstone Tests

Yurioka - 1 14 ± 4 44 ± 7
Yurioka -2 20 ± 6 33 ± 7
Beckert 29 ± 8 25 ± 7
Dueren 24 ± 5 28 ± 8

450
Maximum HAZ Hardness VHN

400

350

300

250

200

150

100
AGL test data, X42-X80
50 Gladstone Data X70, Ultrapipe
0
0 1 2 3 4 5 6 7 8 9
Weld Cooling Time T85C (sec)

Figure 10.10
Variation of measured HAZ hardness with measured t8/5 cooling times for experiments carried
out on the Gladstone flow loop and on water cooled pipe at AGL.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 161

350
Measured HAZ Hardness VHN

300

250

200

150 Beckert
Yurioka-1
Yurioka-2
100
100 150 200 250 300 350

Predicted Hardness from Empirical Equations VHN

Figure 10.11

The accuracy of calculation of HAZ hardness from various empirical equations relating
composition and t8/5 weld cooling times.
Data collected during the Gladstone trial.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 162

(a)

(b)

Predicted average T85C


4.4 seconds

(c) Measured 4.7 seconds

Figure 10.12
Thermal field calculated from 3D quasi-steady state models; (a) general view of thermal field,
(b) maximum temperatures showing shape of fusion zone and HAZ, and (c) showing the
predicted t8/5 cooling times.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 163

Table 10.4
Predicted values of T85C and HAZ Depth Fusion zone depth and HAZ hardness from
Finite element models of individual joints

Test No Heat T85C HAZ Fusion VHN-Beckert VHN-Yurioka


Input depth depth
max min average min max average min max average
2 0.76 2.8 2.4 2.6 1.6 0.3 323 343 333 352 362 357
0.9 3.7 3 3.2 2 0.4 279 313 303 331 347 343
1.05 4.5 3.4 3.7 2.2 0.5 244 293 279 314 338 331

12 0.87 3.4 2.7 2.9 2.3 0.7 293 328 318 338 354 350
1.02 4.2 3.3 3.6 2.5 0.8 256 298 284 320 340 333
1.2 5.4 3.9 4.3 2.8 0.9 214 270 252 297 327 318

23 0.93 4.2 3.2 3.4 2.6 0.8 256 303 293 320 343 338
1.04 4.9 3.7 4 2.9 0.9 230 279 265 306 331 324
1.14 5.6 4.1 4.5 3.1 1 209 261 244 293 322 314

29 1 4.1 3.1 3.3 2.2 0.7 261 308 298 322 345 340
1.07 4.2 3.3 3.5 2.3 0.7 256 298 288 320 340 336
1.12 4.8 3.4 3.7 2.3 0.7 233 293 279 308 338 331

38 1.02 4.3 3.2 3.5 2.4 0.7 252 303 288 318 343 336
1.15 4.7 3.6 3.9 2.5 0.7 237 284 270 310 333 327
1.29 5.4 4 4.4 2.7 0.8 214 265 248 297 324 316

43 0.91 3.9 2.7 3.2 2.1 0.6 270 328 303 327 354 343
1.2 5 3.7 4.4 2.4 0.8 226 279 248 304 331 316
1.81 8.2 6.1 6.7 3.1 1 175 198 188 258 285 276

44 1.31 5.8 4.5 5 2.1 0.25 204 244 226 290 314 304
1.57 6.9 5.3 6 2.4 0.4 185 217 200 273 298 286
1.65 7.2 5.6 6.2 2.5 0.4 182 209 196 269 293 283

45 1.28 6.9 4.5 5.3 1.9 0.2 185 244 217 273 314 298
1.46 7.2 5.5 6.1 2.2 0.2 182 212 198 269 295 285
1.62 7.8 6.2 6.8 2.4 0.3 177 196 187 262 283 275
5.6mm thick x70

8 1.17 5.7 4.4 4.9 2.7 0.8 224 254 241 263 281 273
1.3 6 4.8 5.3 2.9 0.8 218 244 232 260 275 268
1.4 6.5 5.4 5.9 3.2 0.9 210 230 220 254 267 261

19 0.9 4.3 2.8 3.7 1.05 0 257 299 274 282 307 292
1 5.1 3.8 4.4 1.3 0 237 271 254 271 290 281
1.15 5.5 4.4 4.9 1.4 0 228 254 241 266 281 273

20 1.15 6.5 4.5 5.2 3.2 0.9 210 252 234 254 279 269
1.2 7 4.9 5.7 3.4 0.9 203 241 224 250 273 263
1.3 7.2 5.3 6.1 3.65 1 201 232 216 248 268 259

26 0.9 5 3.4 3.9 2.5 0.6 239 282 268 272 297 288
1 7.2 5.3 5.8 3.9 1.2 201 232 222 248 268 262
1.1 8 5.6 6.5 4 1.4 194 226 210 241 264 254

34 1.2 5.8 4.6 5.1 2.8 0.8 222 249 237 262 278 271
1.5 8.3 5.8 6.5 3.3 0.9 192 222 210 239 262 254
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 164

1.8 9 6.6 7.4 3.7 1.1 188 208 199 235 253 246

41 1 6 3.8 4.5 3.7 1.4 218 271 252 260 290 279
1.1 6.1 4.25 5 4.1 1.5 216 258 239 259 283 272
1.3 7.7 5 5.9 5 1.6 196 239 220 244 272 261

42 0.9 4.85 3.5 4.1 3.1 1 243 279 262 274 295 285
1 5.4 3.6 4.3 3.2 1 230 276 257 267 293 282
1.1 5.4 3.8 4.4 3.3 1.1 230 271 254 267 290 281

6.4 mm Ultrapipe
4 0.8 4.7 3.3 3.7 2.2 0.5 263 297 287 288 308 302
0.9 5.5 4 4.6 3.1 0.9 245 279 265 278 298 289
1.1 6.5 4.6 5.3 3.1 0.75 227 265 249 267 289 281

6 0.8 4.6 3.3 3.7 2.2 0.5 265 297 287 289 308 302
0.9 5.2 3.9 4.4 2.6 0.7 251 282 270 282 299 292
1.4 8.1 5.9 6.7 4.05 0.9 207 237 223 253 274 265

14 0.9 5.3 4.1 4.5 2.6 0.7 249 277 267 281 296 291
1 7.3 5.1 5.9 3.4 0.9 215 254 237 260 283 274
1.5 10.2 7.6 8.4 4.8 1.3 195 212 204 240 257 251

15 1 5.9 4.7 5.1 3.2 0.8 237 263 254 274 288 283
1 7 5.2 5.8 3.6 0.9 219 251 239 262 282 275
1.2 7.3 5.7 6.3 4.45 1 215 241 230 260 276 269

16 0.8 4.6 3.8 4.1 2.1 0.5 265 284 277 289 300 296
1.2 7.5 6.1 6.5 3.1 0.8 213 233 227 258 271 267
1.3 8.4 6.7 7.2 3.5 0.9 204 223 217 251 265 261

31 0.7 3.5 2.8 3.1 1.9 0.3 292 308 301 305 315 311
0.9 5.2 3.9 4.3 2.5 0.6 251 282 272 282 299 293
1 5.3 4.2 4.7 2.7 0.7 249 275 263 281 295 288

32 0.9 5.3 3.6 4.3 3 0.8 249 289 272 281 303 293
0.9 5.9 3.9 4.7 3.3 0.9 237 282 263 274 299 288
1.2 7.5 5 4.8 4.8 1.1 213 256 260 258 284 287

33 0.55 2.4 1.6 2 1.3 0.3 317 332 325 321 333 327
0.9 3.9 3.3 3.55 2 0.5 282 297 291 299 308 304
1 5 3.8 4.2 2.2 0.6 256 284 275 284 300 295
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 165

10 10
4.8mm thick X70
Predicted Weld Cooling Time T85C 9 9 5.6mm thick X70

Predicted Weld Cooling Time T85C


8 8

7 7
6 6
(sec)

(sec)
5 5

4 4

3 3

2 2

1 1

0 0
0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10

Measured Weld Cooling Time t 8/5 (sec) Measured Weld Cooling Time t 8/5 (sec)

10
6.4mm thick Ultrapipe
9
Predicted Weld Cooling Times

8
7
T85C (sec)

3
2

0
0 1 2 3 4 5 6 7 8 9 10

Measured Weld Cooling Times t 8/5 (sec)

Figure 10.13
Comparison between predicted and measured t8/5 cooling times for the Gladstone test welds

10
Predicted Weld Cooling Time T85C (sec)

7
6

4
3

2 minimum predicted t8/5 in pipe


1 maximum predicted t8/5 in pipe

0
0 1 2 3 4 5 6 7 8 9 10
Measured Weld Cooling Time t 8/5 (sec)

Figure 10.14
Comparison of the predicted and measured t8/5 cooling times. A spread of t8/5 values are
determined for the weld and the maxium and minimum values are plotted here.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 166

6 6
HAZ Penetration 4.8 mm X70

Predicted Depth of Heat Affected


5 5
Predicted Depth of the Heat
Affected Zone (mm)

4 4

Zone (mm)
3 3

2 2

1 1
HAZ Penetration 5.6 mm X70

0 0
0 1 2 3 4 5 6 0 1 2 3 4 5 6
Measured Depth of Heat Affected Zone (mm) Measured Depth of Heat Affected Zone (mm)
6

HAZ Pentration 6.4 mm Ultrapipe.


Predicted Heat Affected Zone Depth

4
(mm)

0
0 1 2 3 4 5 6
Measured Heat Affected Zone Depth (mm)

Figure 10.15
Comparison of predicted and measured heat affected zone depths for the Gladstone trials

5
Predicted HAZ depth (mm)

0
0 1 2 3 4 5 6
Experimental HAZ depth (mm)

Figure 10.16

Comparison of predicted and measured HAZ depths, using the average heat input values
within the numerical models.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 167

400

350
Measured HAZ hardness VHN

300

250

200

150

100

50 Beckert's relationship
Yurioka-1 relationship
0
0 50 100 150 200 250 300 350 400
Predicted HAZ hardness VHN

Figure 10.17

The accuracy of the HAZ hardness that can be predicted based on the t8/5 values from
numerical models, using either Beckert’s empirical relationship or Yurioka-1.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 168

Material glad4.8
2
1.9
Gas Pressure 3.8 3.8 1.8
Gas velocity 6.2 6.2 1.7 0.000468324 0

Gas temperature 20 20
1.6 0.000468324 3
1.5
Pipe Diameter 300 6 300 1
1.4
Thickness 4.8 4.8 3
1.3
Heat transfer coefficient 468 0.000468
1.2
Heat input kJ/mm

1.1
1
MAOP 10.4 10.4
0.9
Work pressure safety factor 1 1 1
0.8 dia t wtemp hvn ret% wsf
Safe maximum pressure 3.8 0.7 300 3.2 600 300 5 1
residual thickness 5 0.05 0.6 1 373 4.2 700 340 10 1.25

effective thick ness 0.24 0.24 0.5


423 4.8 800 350 15 1.5
0.4
Allowed wall temperature 1000 1273 5
473 5.6 900 380 20 1.75
0.3
523 6.35 1000 400 30 2
0.2
7 1100 50
0.1
Allowable Hardness 340 340 2
0
Required T85C cooling time 1.2 1.172125 0 0.0002 0.0004 0.0006 0.0008 0.001 0.0012
Heat transfer coefficent (W/mm2.K)

Target Heat Input 0.9 Pressure safe factor 1.80 0.66 Cooling time 7.2 Predicted HAZ Hardness 262

Inside wall temperature 790 Remaining thickness 0.50

Figure 10.18
The working zone for 4.8 mm thick X70 pipe under gas flow conditions of 3.8 MPa, and 6.2
metres/second (see Section 11 for an explanation of its derivation).
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 169

15
14
13

Predicted T85C cooling times (sec)


12
11
10
9
8
7
6
5
4
3
2
1
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
Measured T85C cooling times (sec)

Figure 10.19
A comparison of t8/5 cooling times calculated from the hot-tap database against the measured
values for the Gladstone trial. This data is for 6.4 mm thick Ultrapipe.

350

300
Predicted HAZ Hardness

250

200

150

100
100 150 200 250 300 350
Measured HAZ Hardness

Figure 10.20
A comparison of the HAZ hardness calculated from the hot-tap database with the measured
HAZ hardness from the Gladstone trial welds.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 170

11.0 Technology Transfer


11.1 Introduction
Numerical models of in-service welding have been developed using a combination of finite element
heat transfer and stress analysis as described in section 8. Heat transfer calculations can give
predictions of the weld cooling rates, and these have been used to estimate heat affected zone (HAZ)
hardness, which allows some assessment of the potential for weld cracking. The addition of a stress
analysis allows the risk of pipe bursting or burn-through to be established.

The drawback to this approach is that such an analysis takes appreciable computer time. For the
typical models used in this work (see Section 8) a thermal analysis may take approximately 20
minutes, and the stress analysis some 8 hours. From a user perspective these times are too long. In
addition the technique uses relatively expensive computer hardware, and requires some expertise to
efficiently operate the finite element software. These factors are therefore barriers to the use of such
simulations.

To address this difficulty an alternative approach has been developed. This has been termed ‘user
targeted databases’. The concept is straightforward. Firstly, we use the finite element software to
generate a data set for a range of assumed pipe, gas-flow and weld conditions. Then we can
determine the result for a specific pipe wall thickness, gas-flow and weld by interpolation from this
database. Calculating the data set does take time, but interrogation of the database is virtually
instantaneous.

The concept obviously relies on the fact that relatively smooth relationships exist between the
variables, and must necessarily only apply to a relatively small window of variation in gas-flow
conditions for example. However such constraints are relatively easily put into place. For example a
given pipeline will generally have a restricted range of material grades, pipe sizes etc., and it will have
a known variation in gas pressure and flow. Hence the concept of a targeted database developed
around a known set of conditions.

11.2 ‘User targeted’ Databases – The Principle


A simple example illustrates the concept as follows. Consider deriving information with respect to HAZ
hardness for a specific thickness pipe. The desired outcome is the relationship between the heat input
required to achieve a specified HAZ hardness and the heat transfer coefficient at the pipe wall
resulting from a specific gas-pressure and flow.

Firstly a database is developed based on a root-pass, circumferential sleeve weld, using E8018G
electrodes used in the vertical-down welding position. About 120 finite element calculations are
carried out for a fixed pipe diameter using a range of pipe wall thickness, heat transfer coefficients and
weld heat inputs. The four sets of results are shown schematically in Figure 11.1, each set is for a 3,
4, 5 & 6 mm thick pipe wall respectively, and within each set, values are calculated for 5 heat transfer
coefficients, 100, 150, 300, 500, 1000 W/m2.K, and for 6 heat input values, 0.4, 0.5, 0.6, 0.8, 1.0, and
1.2 kJ/mm. In this illustration the minimum t8/5 (the time to cool between 800° and 500°C) for the weld
is the calculated variable.

For a required pipe wall thickness an interpolated data set of t8/5 values is derived for the range of heat
inputs and heat transfer coefficients.

The necessary t8/5 to achieve a desired HAZ hardness can be derived for a specific pipe grade by
using an empirical relationship between composition, t8/5 and hardness. For this purpose the Yurioka
–1 relationship has been adopted.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 171

From the data set the heat input necessary to achieve this t8/5 can then be determined, again by
interpolating at a constant heat transfer coefficient. This gives a line which separates the heat inputs
which will give a hardness greater than that required from those that will give a hardness below that
required, see Figure 11.1. Put simply if we accept an hardness limit of say 350 VHN, then heat inputs
above the line are safe, whilst those below the line are unsafe.

The current pipe diameter, gas pressure and flow are used to calculate the current heat transfer
coefficient, which defines the current operating condition. This then allows the required minimum heat
input to be determined, as shown on Figure 11.1.

This procedure can be copied and applied to other critical parameters, for example:

• Maximum inside wall temperature of the pipe,


• Remaining effective thickness of the pipe wall,
• Safe bursting pressure based on the derived effective cavity in the pipe wall.

11.3 ‘User targeted Databases’ – Background


The successful development of a ‘user targeted database’ relies on a number of observations, as
follows:

a) Relatively smooth relationships exist between the variables.


Figure 11.2 and 11.3 illustrate that a typical data set is well behaved and exhibits smooth and
consistent relationships between inputs (e.g. weld heat input) and calculated values (e.g. t8/5,
maximum wall temperatures etc.).

b) The pipe diameter does not play a significant role in the calculated thermal cycles of the
weld.
For a given gas flow and pressure the pipe diameter influences the heat transfer coefficient from
the pipe wall. However, for a given heat transfer coefficient the cooling rate of a weld is essentially
independent of pipe diameter. This means that a specific combination of gas-flow, gas pressure
and pipe diameter can be used to derive the heat transfer coefficient. Then a data set from
numerical models based on a constant pipe diameter and a range of heat transfer coefficients can
be used to generate valid thermal data for a range of gas flows and pipe diameters.

c) The sleeve thickness does not significantly influence the weld’s cooling rate within the pipe
wall.
Therefore provided that the cooling rates within the sleeve are not of concern then calculated data
can be taken to generally apply to any sleeve thickness. However if the hardness of the sleeve’s
HAZ is of concern the composition of the sleeve must be provided and this may compromise this
simple approach. In such a case it may be necessary to generate a data set for a specific sleeve
geometry.

d) The relative strength of steel at elevated temperature is generally a function of temperature


and the original room temperature yield strength, i.e.,

σ y.elevated .temperature = σ y.room.temperature . factor (T )

This approximation allows the size of the ‘effective cavity’, as described in Section 9, to be
calculated without concern for the specific pipe grade, since the calculation depends on the
relative strength between low and high temperature. For a set of pipe steel grades it can be
observed that the temperature factor is approximately the same, independent of the pipe grade, as
shown in Figure 11.4. Hence the ‘effective cavity’ width and depth derived from a given thermal
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 172

field is the same for all pipe grades. To estimate the burst pressure of the pipe containing this
effective wall cavity requires knowledge of the specific pipe diameter and specified minimum yield
strength (SMYS), but this can be entered into the program that interrogates the database.

e) Thermal properties of steel are essentially independent of the pipe steel grade.
Hence cooling behaviour and the calculation of t8/5 cooling times can be considered to apply to all
pipe grades, only the relationship between the pipe grade and hardness and t8/5 will require input
of the pipe composition.

f) For a given electrode type, and electrode diameter the heat input is the factor which
controls the thermal field and weld cooling time.
For a specified electrode diameter and type the deposition rate and the weld bead shape and size
are related to the weld heat input. So weld heat input is the primary factor defining the weld
procedure.

11.4 User targeted database – its development


Based on the above concept a system has been developed and implemented using Microsoft Excel.
In this the following input variables are required:
• gas flow,
• gas pressure,
• pipe wall thickness (between 3 – 6 mm),
• pipe diameter,
• pipe composition,
• specified minimum yield strength,
• preheat temperature,
• required HAZ hardness,
• allowable maximum temperature at the inside of the pipe wall,
• allowable minimum effective wall thickness expressed as a percentage of the original wall
thickness.

The test data-set used to illustrate this process is restricted to circumferential fillet welds using
E8018G, 2.5mm diameter electrodes in the vertical-down welding position. For a set of input
conditions, namely:
• 300 mm pipe diameter,
• pipe wall thickness of 3, 4, 5 and 6 mm,
• a sleeve thickness of 6 mm,
• heat transfer coefficients of 100, 150, 300, 500 and 1000 W/m2.C,
• weld heat inputs of 0.4, 0.5, 0.6, 0.8, 1.0, 1.2 kJ/mm,
• preheat temperature of 30,50,100,150 and 200°C,
• a gas temperature of 27°C.

The following database values were calculated:

• the minimum t8/5 cooling time for the weld zone within the pipe region,
• the maximum wall temperature at the inside of the pipe wall,
• the minimum residual effective thickness,
• the width of the effective cavity along the pipe.

In addition the fusion zone depth and the HAZ depth were also calculated but these have not been
incorporated within the data-base since they are not normally used to assess either hydrogen assisted
cold cracking or burn-through.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 173

The analysis and interpolation procedures used with the different calculated variables were as follows:

(a) Cooling Time and Hardness


Firstly linear relationships were established between t8/5 cooling times and the level of preheat, for a
constant pipe wall thickness, constant heat transfer coefficient, and constant heat input, as shown in
Figure 11.5.

( t8/5 ) at a constant heat input, pipe wall thickness, heat transfer = { a( Tp) + b}
These relationships allowed t8/5 cooling times to be interpolated for a chosen preheat level. An
exponential relationship was then determined between, t8/5 values and pipe wall thickness at constant
heat input and heat transfer coefficient, as illustrated in Figure 11.6.
For chosen preheat level:

( t8/5) at a constant heat input, heat transfer = A exp(B. pipe thickness)


This enabled t8/5 values to be interpolated for a chosen pipe wall thickness. Second order polynomial
fits were then established to define the relationships between t8/5 and heat input at the chosen pipe
wall thickness and a constant heat transfer coefficient, see Figure 11.7.

( Heat Input = a + b t8/5 + c( t8/5 )2 ) at a constant heat transfer coefficient


From the Yurioka-1 relationship between hardness and t8/5 for the specified pipe composition the t8/5
cooling time needed to achieve a required hardness can be determined. The necessary heat input to
achieve this t8/5 can then be determined from the above equation set. The plot of ‘required heat input’
versus heat transfer coefficient is then determined, as shown in Figure 11.8.

(b) Maximum Temperature


The interpolation process for the maximum pipe wall temperature on the inside of the pipe follows a
similar procedure to that for t8/5 described above. Linear relationships between wall temperature and
the level of preheat, for a constant pipe wall thickness, constant heat transfer coefficient, and constant
heat input, as shown in Figure 11.9.

( Wall Temperature ) at a constant heat input, pipe wall thickness, heat transfer = { a( Tp) + b}
An exponential relationship determined between wall temperature and pipe wall thickness at constant
heat input and heat transfer coefficient, as illustrated in Figure 11.10. For chosen preheat level:

( Wall Temperature ) at a constant heat input, heat transfer = A exp(B. pipe thickness)
Second order polynomial fits were then established to define the relationships between wall
temperature and the heat input at the chosen pipe wall thickness and a constant heat transfer
coefficient, see Figure 11.11.

(Heat Input = a + b(wall temperature)+c(wall temperature )2) at a constant heat transfer coefficient
The necessary heat input to achieve a chosen maximum wall temperature can then be determined
from the above equation set. The plot of required heat input versus heat transfer coefficient is then
determined, as shown in Figure 11.12.

(b) Remaining effective thickness


The ‘remaining effective thickness’ is based on the idea that the local softening in the vicinity of the
weld pool can be considered as a reduction in the original pipe wall thickness at ambient temperature.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 174

The procedure for calculating the ‘remaining effective thickness’ has been described in Section 9. It
can be used as a parameter to qualitatively assess potential burn-through risk.

The interpolation system begins by establishing the heat input required to achieve a chosen
‘remaining effective thickness’ which is represented as a percentage of the original wall thickness, at
constant levels of preheat, pipe wall thickness and heat transfer coefficient.

Unlike the previous dataset the relationships illustrated in Figure 11.13 do not easily fit simple
mathematical forms. Therefore in this case piece-wise linear interpolation is used to generate a data
set of the ‘required-heat-inputs’ for a range of preheat temperatures, pipe wall thicknesses and heat
transfer coefficients. Approximate linear relationships between the required heat input and the
preheat temperature are used to interpolate a reduced data set at a chosen value of preheat
temperature, as illustrated in Figure 11.14. Again piece-wise linear interpolation is used to calculate a
‘required-heat-input’ dataset for a given thickness as illustrated in Figure 11.15.

From this procedure the heat input needed to achieve the specified percentage reduction in thickness
can be determined for the range of heat transfer coefficients, as illustrated in Figure 11.16.

(d) Effective cavity and the safe bursting pressure


This procedure seeks to provide information on burn-through risk by calculating the bursting condition
during welding. It relies on representing the weld as an effective cavity in the pipe wall. As described
in Section 9 the effective cavity is calculated from an estimate of the pipe’s strength at temperature
and the local thermal field around the weld. Then using the maximum cavity depth and the cavity
width along the pipe axis, the ratio of the safe maximum pressure to the maximum allowable operating
pressure, is calculated using a modified B31G procedure:

d
(1 + 0.85. )
SMP 65 t
= (1 + )
MAOP SMYS d 1
(1 + 0.85. . )
t M
Where, M is the Folias Factor which is given by:
1
1.255 L2 0.0135 L4 2
M = (1 + . − . )
2 D.t 4 D 2 .t 2
SMP is the safe maximum pressure
MAOP = maximum allowable operating pressure
SMYS = specified minimum yield strength of the pipe material
d = cavity depth
t =pipe wall thickness, and
L = cavity width
D = pipe diameter

The interpolation procedure is very similar to that described for the ‘remaining effective thickness’. It
begins by establishing the heat input required to meet a target safe maximum pressure (SMP) which is
expressed as a proportion of the maximum allowable operating pressure (MAOP), ie as SMP/MAOP.
The required safe maximum pressure is taken as the current working pressure multiplied by a chosen
safety factor.

In this case a piece-wise linear interpolation is used to generate a data set of the ‘required-heat-inputs
needed to match a specified safe maximum pressure’, for a range of preheat temperatures, pipe wall
thicknesses and heat transfer coefficients, see Figure 11.17. Approximate exponential relationships
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 175

between the required heat input and the preheat temperature are used to interpolate a reduced data
set at a chosen value of preheat, as illustrated in Figure 11.18.

Exponential relationships are fitted to the variation of ‘required-heat-input’ and pipe wall thickness, as
illustrated in Figure 11.19.

From this procedure the heat input needed to meet the target safe maximum pressure can be
determined for the range of heat transfer coefficients, as illustrated in Figure 11.20.

11.5 User targeted database Software - The Implementation

The procedure described in section 11.4 has been implemented within a Microsoft Excel spreadsheet.

Program input and output takes place on one screen as shown in Figure 11.21

The pipe composition and specified minimum yield strength must be known. The pipe’s diameter and
wall thickness are entered via drop-down boxes, and the gas pressure and gas flow is input. The
output is in the form of a graph of heat input versus heat transfer coefficient, shown in Figure 11.22.
For the current pipe diameter, gas pressure, and gas flow velocity a heat transfer coefficient is
calculated and identified on the graph as a vertical line. This represents the current pipe operating
condition. Various lines, which are the heat inputs required to achieve specified outcomes, are also
plotted.

The red line indicates the heat input needed to achieve a specified hardness as entered in a drop-
down box to the left of the graphical output. Heat inputs above this line are predicted to give a
hardness below the specified value. The other lines relate to various parameters used to evaluate the
risk of burn-through. Heat inputs below these lines are ‘safe’, whereas heat inputs above the line are
‘unsafe’ and exceed the criteria related to that line. For example, the blue line indicates the heat input
that would create the specified maximum inside wall temperature on the pipe, again entered in a drop-
down box to the left of the graphical output. Heat inputs below the blue line are predicted to give
maximum wall temperature below that specified.

Setting the required hardness to 350 VHN, and the maximum temperature to 980°C, is equivalent to
the limits used by the Battelle program, in that heat inputs above the red line are predicted to avoid
weld cracking and those below the blue line are predicted to avoid burn-through. Therefore, put
simply, the zone between the burn-through limit line and the required hardness line is the safe working
zone.

The lower black line indicates the heat input that would result in the ‘percentage remaining effective
wall thickness’ set within the drop-down box to the left of the graphical output.

The magenta line relates to the heat input that would result in a safe maximum pressure equal to the
working pressure times a specified safety factor. This is again set in the drop-down boxes. If this
value exceeds the maximum allowable operating pressure (MAOP) the pressure limit is set to MAOP
and the printout at the side of the graph indicates this.

The upper black line represents ‘full penetration’ of the pipe wall which is also included. This is based
on the heat input necessary to achieve a 1500°C temperature on the inside of the pipe wall.

Finally along the base of the screen there is a cell to allow the user to input a target heat input. For
the current operating conditions the program then calculates the predicted t8/5 cooling time and the
HAZ hardness, together with the safe bursting pressure. The ratio of the safe bursting pressure to the
current working pressure is then expressed as a pressure safety factor.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 176

11.6 User targeted databases – testing & evaluation


Various strategies have been adopted to validate the current database and the interpolation
procedures used within it.

11.6.1 Comparison with direct calculation.


To confirm that the interpolation routines were not dramatically smoothing or introducing errors into the
predicted values, finite element models were used to directly calculate, t8/5 , maximum inside wall
temperatures, and critical SMP/MAOP ratios. These were then compared with values derived from the
database. Figures 11.23 & 11.24 show that good agreement was obtained for both t8/5 and maximum
temperatures. For the SMP/MAOP ratio the database appears to generate some reduction in values
for the 4mm thick pipe tested, as shown in Figure 11.25. This may be associated with the restricted
data set, or due to interpolation smoothing. Database predictions under estimate the SMP/MAOP
from direct calculation and therefore are generating an increased conservatism in the assessment of
burn-through.

11.6.2 Comparison with EWI cooling data


In Figures 11.26 & 11.27 the data base calculations of t8/5 are compared with experimentally
determined values published by EWI (1). Whilst there is some scatter the results show general
agreement. Some errors were generated when entering weld heat input values that were significantly
outside the current data-set range of 0.4-1.2 kJ/mm. It was clear that excessive extrapolation should
to be avoided and restrictions were placed on entry values, particularly in heat input values (restricted
to 0.2-1.4 kJ/mm) and in the pipe wall thicknesses ( 2.5-6.35 mm).

11.6.3 Comparison with Battelle Predictions


In Figure 11.28 and 11.29, calculated t8/5 values from the Battelle program (2) are overlaid on Figures
11.26 and 11.27. Good agreement has been achieved, with no indication that the database approach
is worse than the Battelle program.

11.6.4 Comparison with burn-through data


There is a restricted amount of experimental data relating to burn-through so it is not possible to
rigorously test this aspect of the database predictions.

Figure 11.30 shows a comparison between the failure pressures calculated from the database and the
experimental measurements of Wade (3). Whilst the predicted trend is in broad agreement the failure
pressures do not directly compare with Wade’s data. It is clear that the predicted values are a
conservative estimate of the burn-through condition.

In Table 11.1 some further experimental values of pressure / heat input burn-through limits are
compared with values predicted from the database. Again the values are reasonable, but they
underestimate the limiting heat input quoted by Bruce et al (4) and overestimate those quoted by Bout
& Gretskii (5).

11.6.5 A Comparison with measured data from Gladstone Flow Loop Experiments
The experimental data derived from the Gladstone experiment (see Section 10) also allowed the
opportunity to test the hot-tap database. This was achieved by simply entering the test conditions, pipe
geometry, composition, gas flow rate and pressure. Then selected heat inputs were entered and the
database used to calculate t8/5 cooling times and the HAZ hardness. The experimental monitoring of
heat input for each weld showed some degree of scatter so in this case the average measured heat
input was used and the calculated values compared against the measured data to provide a test of the
software’s capabilities, as shown in Figure 11.31.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 177

Figure 11.32 (a,b,c) attempts to allow for the observed variation in heat input during these
experiments. Significant variation occurred, hence the agreement between predicted and measured
t8/5 must allow for this. It was concluded that the prediction of t8/5, whilst it shows some scatter, it is no
worse than that obtained from a direct calculation through individual models. It appears that the
database approach has not compromised the accuracy of t8/5 prediction.

Figure 11.33 shows that when using the Yurioka-1 relationship to calculate hardness there is a
tendency to over-predict the HAZ hardness. The choice of the Yurioka-1 relationship was based on the
findings of earlier AGL experiments. However as discussed in Section 10.5.3, this was not supported
by the Gladstone trial and better agreement may be achieved with the Beckert equation. The effect of
using Yurioka’s equation in the database would be to overestimate the lower heat input limit required
to achieve a given hardness and hence would be more conservative than the Beckert one, however
this may restrict the operational window. To avoid this perhaps more consideration has to be given to
the choice of composition- t8/5 -hardness relationships.

The database calculations also use a standard empirical equation to determine deposited area and
weld bead leg lengths. As previously discussed in Section 10.5.2, the Gladstone welds have shown
some variation from this. This is not considered to be significant to the prediction of weld cooling
times. However it may be necessary to take such variations into account for the calculation of
‘effective cavity depth’ and burn-through limits since these will depend on the depth of penetration
which is sensitive to weld angle. Since the burn-through limits are not directly tested in these
experiments it is impossible to test this feature of the database using the present data.

11.7 Conclusions

1. The process of generating useful information from a previously calculated set of results has
proved to be a useful strategy to provide immediate user information about the safe window for in-
service welding.

2. Some potential errors from excessive extrapolation beyond the data set were identified. However
within the data set relatively smooth and simple interpolation schemes could be implemented with
a reasonable accuracy.

3. A prototype software package was created. This is currently restricted to pipe wall thickness
between 2.5-6.5 mm, and to a heat input range of 0.2-1.4 kJ/mm, using 2.5 mm diameter, 27P
electrodes in the vertical-down welding position. However there appears to be no restriction in
generating further sets of data and managing these with the same software protocols.

4. Within the constraints of inherent variability in experimental data, and from the restricted amount
of information relating to burn-through limits, the predictions of the hot-tap database have been
validated.

5. The t8/5 cooling times have been calculated with reasonable accuracy. However there is some
difficulty in reliably determining hardness. Initial experimentation supported the accuracy of the
Yurioka-1 relationship between composition, t8/5 and hardness, but its accuracy was not
substantiated within the Gladstone flow loop experiments. Currently the hot-tap database uses
Yurioka-1, but it may prove necessary to modify this or to include an option to allow the user to
choose a preferred relationship for a given material.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 178

References

1. Bruce W. A. & Threadgill P. L., Effect of Procedure Qualification Variables for Welding Onto In-
service Pipelines, American Gas Association Report J7141, July 1994.

2. Kiefner J.F & Fischer R. D. Models Aid Pipeline Repair Welding Procedure,
Oil & Gas Journal March 1988, pp41-47.

3. Wade J.B., Hot Tapping of Pipelines, Australian Welding Research Association Research Report,
Snowy Mountains Corporation1978.

4. Bruce W.A., Holdren R.L. & Mohr W. C., Repair of Pipelines by Direct Deposition of Weld Metal –
Further Studies, Final report Edison Welding Institute, EWI Project J7283, November 1996.

5. Bout V.S. & Gretskii Yu. Ya., Arc Welding Application on Active Pipelines, Pipeline Technology,
Volume 1, Ed. R. Denys, pub Elsevier Science BV. pp550-558.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 179

Pipe wall thickness=t 1 Pipe wall thickness=t 2 Pipe wall thickness=t 3 Pipe wall thickness=t 4

Heat input
Heat input

Heat input

Heat input
T85 C T85 C T85 C T85 C

Heat transfer coefficient Heat transfer coefficient Heat transfer coefficient Heat transfer coefficient

Required pipe wall thickness Pipe grade, composition


Pipe wall thickness=t Required HAZ hardness
Heat input

Relationship between T85C & HVN


T85 C

Heat transfer coefficient Required T85 C

Required heat input versus heat transfer coefficient


Current gas flow, pressure, pipe diameter
Pipe wall thickness=t

Heat transfer coefficient


Heat input

HVN < required

HVN > required

Heat transfer coefficient

Resulting value of the minimum required Heat Input


in order to achieve a HAZ of require hardness

Figure 11.1
Schematic illustration of the way that a data base of results can be used to derive the critical
value of heat input required to achieve a given HAZ hardness for a specific pipe wall
thickness, steel grade, pipe diameter and gas flow condition.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 180

25
3 mm thick pipe

t 8/5 Cooling Time (sec) 20 6 mm thick pipe

15

10

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4
Welding Heat Input (kJ/mm)

Figure 11.2
Two examples of the observed smooth relationships between heat input and t8/5 cooling
times, in this case for a 100°C preheat and a heat transfer coefficient of 300 W/m2K.

2000
Maximum Temperature on Inside Wall (deg K)

1800

1600

1400

1200

1000

800

600

400 3 mm thick pipe


5 mm thick pipe
200

0
0 50 100 150 200 250
Preheat Temperature (deg C)

Figure 11.3
Two examples of the observed smooth relationships between preheat and wall temperature,
In this case for a heat input of 1kJ/mm and a heat transfer coefficient of 100 W/m2K.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 181

1.2

1 X70
X42
Normalised Yield strength

X60
0.8 X52
Average value

0.6

0.4

0.2

0
0 200 400 600 800 1000 1200 1400
Temperature (C)

Figure 11.4
The variation of yield strength with temperature for a range of pipe steels. The data has been
normalised and a common curve approximated to fit to the data.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 182

3.5

t8/5 cooling time (secs) 2.5

1.5
1.2 kJ/mm
1
1 0.8
0.6
0.5 0.5
0.4

0
0 50 100 150 200 250
Preheat (degree C)

Figure 11.5
The variation of t8/5 cooling times with preheat temperature, for a range of weld heat inputs
and a pipe wall thickness of 3 mm, and a constant heat transfer coefficient of 100 W/m2C.

18

1.2 kJ/mm
16
1
0.8
14
0.6
t8/5 cooling time (secs)

12 0.5
0.4
10 y = 23.714e-0.2331x

0
0 1 2 3 4 5 6 7
Pipe Thickness (mm)

Figure 11.6
The variation of t8/5 cooling time with pipe wall thickness, for constant levels of heat input at a
fixed heat transfer coefficient of 100 W/m^2.K. Data points interpolated at a chosen preheat
level of 30°C.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 183

1.4

1.2
Welding Heat Input (kJ/mm)

0.8

0.6
100 W/m^2K
150
0.4 300
500
1000
0.2
y = -0.003x 2 + 0.122x + 0.2567

0
0 2 4 6 8 10 12 14
t8/5 Cooling Time (secs)

Figure 11.7
The variation of t8/5 cooling time with welding heat input at fixed heat transfer coefficients.
This data set is interpolated for a chosen pipe wall thickness of 4 mm.
The lines in the figure are second order polynomial fits to the data.

1.6
Welding Heat Input to Achieve Required

1.4

1.2

1
t8/5

0.8

0.6

0.4

0.2

0
0 0.0002 0.0004 0.0006 0.0008 0.001 0.0012

Heat Transfer Coefficient (W/mm^2. K)

Figure 11.8
The heat input needed to achieve a hardness of 350 HVN for a pipe wall thickness of 4 mm.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 184

1800

Maximum Temperature on the Inside of the


1600

1400

1200

Pipe ( deg K)
1000

800
0.4 kJ/mm
600 0.5
0.6
400
0.8
200 1
1.2
0
0 50 100 150 200 250
Preheat Temperature ( deg. C )

Figure 11.9
The relationship between the maximum temperature at the inside of the pipe and the preheat
temperature for a range of weld heat inputs. This data is for a, 4mm thick pipe and a heat
transfer coefficient of 100 W/m2°C.

2000
Maximum Temperature on the Inside of the

1800

1600

1400
Pipe (deg K)

1200

1000 0.4 kJ/mm


0.5
800
0.6
600 0.8
1
400
1.2
200
y = 2327.1e-0.1909x
0
0 2 4 6
Pipe Wall Thickness (mm)

Figure 11.10
The relationship between maximum temperature on the inside of the pipe and the pipe wall
thickness. Data is for a chosen preheat temperature of 30°C and a heat transfer coefficient of
150 W/m^2.K. Each line indicates an exponential fit to the interpolated data points.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 185

1600

Maximum Temperature on the Inside of


1400

1200

the Pipe (deg k)


1000

800
100 W/m^2K
600 150
300
400 500
1000
200
y = -173.26x 2 + 1027.7x + 284.46
0
0 0.2 0.4 0.6 0.8 1 1.2 1.4
Welding Heat Input (kJ/mm)

Figure 11.11
The variation of wall temperature with welding heat input at fixed heat transfer coefficients.
This data set is interpolated for a chosen pipe wall thickness of 4 mm.
The lines are 2nd order polynomial fits to the data.
1
Required Wall Temperature (kJ/mm)
Heat Input Required to Achieve

0.9

0.8

0.7

0.6

0.5
0 0.0002 0.0004 0.0006 0.0008 0.001 0.0012
Heat Transfer at Inside of Pipe (W/mm^2.K)

Figure 11.12
The heat input required to achieve a chosen maximum inside wall temperature of the pipe
wall.
This for a value of 980°C, for a pipe wall thickness of 4 mm.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 186

1.2

Reduction in thickness / Original Thickness


1

0.8

0.6

0.4
3 mm wall
4
0.2 5
6

0
0 0.5 1 1.5 2 2.5 3
Welding Heat Input (kJ/mm)

Figure 11.13
Variation in the reduction of thickness ratio with welding heat input, for a constant preheat
level of 50°C, a range of pipe wall thicknesses, and a constant heat transfer coefficient.

0.9
Heat Input Required to Meet Thickness

0.8

0.7
Reduction (kJ/mm)

0.6

0.5

0.4

0.3 100 W/m^2.K


150
0.2 300
500
0.1
1000

0
0 50 100 150 200 250
Preheat Temperature (deg C)

Figure 11.14
Variation of the welding heat input required to achieve a specified thickness reduction ratio
with preheat temperature, for a range of heat transfer coefficients and
an original wall thickness of 4mm.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 187

5
100 W/m^2.K
4.5

Required Weld Heat Input to meet reduction in


150
300
4
500
1000
3.5
Thickness (kJ/mm)
3

2.5

1.5

0.5

0
0 1 2 3 4 5 6 7
Pipe Wall Thickness (mm)

Figure 11.15
The heat input required to meet a specified thickness reduction ratio for different pipe wall
thickness at a chosen preheat level of 50°C.

1.2
Heat Input Required to Meet Thickness

1.1

1
Reduction (kJ/mm)

0.9

0.8

0.7

0.6
0 0.0002 0.0004 0.0006 0.0008 0.001 0.0012
Heat Transfer Coefficient (W/mm^2.K)

Figure 11.16
Required relationship between the heat input to achieve a specified thickness reduction ratio
of 0.95 and the heat transfer coefficient. Data set is for a chosen preheat temperature of
50°C and a pipe wall thickness of 4 mm.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 188

1.4

1.2

Ratio of SMP/MAOP
0.8

0.6

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1 1.2 1.4 1.6
Welding Heat Input (kJ/mm)

Figure 11.17
Variation of calculated safe maximum pressure (based on the cavity model) with welding heat
input. This data for a pipe wall thickness of 4mm, a preheat temperature of 50°C and a heat
transfer coefficient of 100 W/mm^2.K.

1.2
Required Heat Input to give target SMP/MAOP

1
Bursting Pressure (kJ/mm)

0.8

0.6

100 W/m^2.K
150
0.4
300
500
1000
0.2
y = 1.0617e-0.002x

0
0 50 100 150 200 250
Preheat Temperature ( deg C)

Figure 11.18
Variation of ‘heat input required to meet a set bursting pressure’ and the preheat temperature.
This data for a fixed pipe wall thickness and various heat transfer coefficients.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 189

Heat Input Required to Meet Target SMP/MAOP


100 W/m^2.K
3.5 150
300
3 500

Bursting Condition (kJ/mm)


1000

2.5
y = 0.1353e0.5328x

1.5

0.5

0
0 1 2 3 4 5 6 7
Pipe Wall Thickness (mm)

Figure 11.19
The variation of the value of ‘heat input required to meet a set bursting pressure’ and the pipe
wall thickness. This data corresponds to a fixed preheat temperature and heat transfer
coefficient.

1.6

1.4
SMP/MAOP Bursting Condition (kJ/mm)
Heat Input Required to Meet Target

1.2

0.8

0.6

0.4

0.2

0
0 0.0002 0.0004 0.0006 0.0008 0.001 0.0012
Heat Transfer Coefficient ( W/mm^2.K)

Figure 11.20
Relationship between the heat input needed to achieve a specified safe bursting pressure and
the heat transfer coefficient. Data set is for a chosen preheat temperature of 50°C and a pipe
wall thickness of 4 mm.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 190

2
Material
? Material 4X70 4
1.9
GasPressure
Gas Pressure (MPa) 2 2
1.8
Gasvelocity
Gas velocity (m/s) 6 6 0.000261357 0
1.7
Gastemperature
Temperature (C)
Gas 20 20 1.6 0.000261357 3

PipeDiameter
Pipe Diameter (mm) 6 373 1.52
373

Thickness (mm)
Thickness 4 4 1.43
1.3
C Heat transfer c'f (W/m^2.K) 261 0
Heat input kJ/mm

1.2
Preheat Level (C)
C Preheat 50
50 1.12

MAOP (MPa) 6.9 6.9


1
0.9
Workpressure
C Work Pressuresafety
Safe factor
factor 1 1
1 mat T dia t wtemp hvn ret% wsf
0.8
Safe max pressure (MPa) 2
0.7 X42 30 273 2 600 350 5 1
% remaining
C residual thickness
thickness 5.00 1 X52 50 373 3.2 700 360 7.5 1.25
5
0.6
effective thickness (mm) 0.2 0.2 0.5 X60 100 423 4 800 370 10 1.5
Max wallwall
? Allowed temperature (C)
temperature 980 0.45
980 X70 150 473 4.88 900 380 15 1.75
0.3
X80 175 523 5.2 980 400 30 2
0.2 agl6 200 6.35 1100 50

Allowable HV max
0.1
C Allowable Hardness 350 350 1 agl7
0
Req'd t8/5C cooling time (s) 1.6 1.6
0 0.0002 0.0004 0.0006 0.0008 0.001 0.0
Heat transfer coefficent (W/mm2.K)
TargetHeat
? Target heat input
Input (kJ/mm) 1 SMP/MAOP 0.28 Inner wall temperature (C) 969 Cooling time t8/5 (sec) 8.3
Pressure safe factor 0.98
Safe max pressure (MPa) 1.96 Rem'ing thickness (mm) 0.07 Predicted HAZ Hardness 274
CRCWS

Cooperative Research Centre for Welded Structures: Pipeline Program.

Figure 11.21
A view of the screen produced by a Microsoft Excel program emulating the hot-tap data-base
usage for calculation of in-service welding outcomes.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 191

2
1.9
1.8
1.7
1.6
1.5
1.4
Burnthrough limits
1.3
Heat input kJ/mm

1.2
1.1
1
0.9
0.8
0.7
0.6
Hardness limit line
0.5
0.4
0.3
0.2
0.1
0
0 0.0002 0.0004 0.0006 0.0008 0.001 0.0012
Heat transfer coefficent (W/mm2.K)

Figure 11.22
The final graphical representation of a safe working zone between the heat input to achieve
350 VHN, and the chosen burnthrough limit. This is for an X60 pipe with a 4 mm pipe wall
thickness at a preheat temperature of 50°C.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 192

18

t8/5 values calculated from data-base


16

14

12

10
(sec)
8

0
0 2 4 6 8 10 12 14 16 18
t8/5 values directly calculated from FEA model (sec)

Figure 11.23
A comparison between t8/5 cooling times calculated directly from a finite element model and
those derived by interpolation from the data-base.

1400
Maximum Wall Temperature calculated

1200
from data-base (deg C)

1000

800

600

400

200

0
0 200 400 600 800 1000 1200 1400
Maximum Wall Temperature directly calculated from
FEA models (deg C)

Figure 11.24
A comparison between values of maximum inside wall temperature calculated directly from a
finite element model and those derived by interpolation on the data-base.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 193

1.2

SMP/MAOP derived from data-base


1

0.8

0.6

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1 1.2

SMP/MAOP directly calculated from FEA model

Figure 11.25
A comparison between the estimated safe bursting pressure calculated directly from finite
element results and the value derived by interpolation on the database.

12
4.88 mm pipe wall thickness
Predicted t 8/5 cooling time (sec)

10

2 Hot-tap database

0
0 2 4 6 8 10 12
Measured t 8/5 cooling time (sec)

Figure 11.26
A comparison between t8/5 cooling times determined from the data-base and that measured
in flow loop tests by EWI (1).
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 194

14
6.35 mm pipe wall thickness

12

Predicted t 8/5 cooling time (sec) 10

2 Hot-tap database

0
0 2 4 6 8 10 12 14
Measured t 8/5 cooling time (sec)

Figure 11.27
A comparison between t8/5 cooling times determined from the database and that measured in
flow loop tests by EWI (1).
12
4.88 mm pipe wall thickness
Predicted t 8/5 cooling time (sec)

10

Hot-tap database
2
Battelle Program

0
0 2 4 6 8 10 12
Measured t 8/5 cooling time (sec)

Figure 11.28
The comparison in Figure 11.26 is overlaid with predictions from the Battelle program.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 195

14
6.35 mm pipe wall thickness

12
Predicted t 8/5 cooling time (sec)

10

Hot-tap database
2
Battelle Program

0
0 2 4 6 8 10 12 14
Measured t 8/5 cooling time (sec)

Figure 11.29
The comparison in Figure 11.27 is overlaid with predictions from the Battelle program.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 196

14

Bursting Pressure (MPa) 12

10

2 Measured by Wade - Burst


Measured by Wade - bulge > 1 mm
Burst Pressure from Data-Base
0
0 0.5 1 1.5 2
Heat Input (kJ/mm)
Figure 11.30
Wade’s (3) experimental data for burn-through conditions in 5 mm thick pipe, compared to
the predicted envelope of safe maximum pressure from the hot-tap data base.

Table 11.1 Comparison of experimentally determined critical levels of heat input to cause
burn-through and values determined from the cavity model.

Source Wall Gas Pressure Critical Predicted


Thickness Heat Input Heat Input
(mm) (MPa) (kJ/mm) (kJ/mm)
Bruce et al (4) 3.2 * 6.2 0.54 0.41

4.0 * 6.2 0.78 0.69

Bout & Gretskii (5) 3.0 4.0 0.48 0.50

3.0 3.0 0.51 0.54

* data determined during weld repair procedure testing therefore this is remaining thickness
in restricted region.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 197

14
4.8 mm thick pipe
5.6 mm
12 6.4 mm
Predicted t 8/5 Cooling Time (sec)

10

0
0 2 4 6 8 10 12 14
Measured t 8/5 Cooling Time (sec)

Figure 11.31
An indication of the accuracy in prediction of t8/5 cooling times. Three pipes with different wall
thickness were considered, and the t8/5 cooling time was based on a measured average heat
input for the weld run.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 198

14

5.6 mm thick pipe


12

Predicted t 8/5 Cooling Time (sec)


10

0
0 2 4 6 8 10 12 14

Measured t 8/5 Cooling Time (sec)

(a)
14

4.8 mm thick pipe


12
Predicted t 8/5 Cooling Time (sec)

10

0
0 2 4 6 8 10 12 14
Measured t 8/5 Cooling Time (sec)
(b)

14

12
Predicted t 8/5 Cooling Time (sec)

10

2 6.4 mm thick pipe

0
0 2 4 6 8 10 12 14
Measured t 8/5 Cooling Time (sec)

(c)

Figure 11.32
An indication of the accuracy in prediction of t8/5 cooling times. This data is for a pipe with (a)
4.8 mm (b) 5.6 mm and (c) 6.4 mm wall thickness, and the calculated t8/5 cooling times were
based on the minimum, maximum and average measured heat input for the weld run.
CRCWS Project 96:34, In-Service Welding on Gas Pipelines: Final Project Report 199

400

Predicted Hardness Based on Average Heat


350

300

250

Input (HVN)
200

150

100

50 Hardness calculated from Yurioka -1

0
0 50 100 150 200 250 300 350 400
Measured Maximum Hardness in HAZ on Weld Centreline (HVN)
(a)
400
Predicted Hardness Based on Average Heat

350

300

250
Input (HVN)

200

150

100

50 Hardness calculated from Beckert's Equation

0
0 50 100 150 200 250 300 350 400
(b)
Measured Maximum Hardness in HAZ on Weld Centreline (HVN)

400
Predicted Hardness Based on Average Heat

350

300

250
Input (HVN)

200

150

100

Hardness calculated from Beckert's Equation


50
Hardness calculated from Yurioka -1

0
0 50 100 150 200 250 300 350 400

(c) Measured Maximum Hardness in HAZ on Weld Centreline (HVN)

Figure 11.33
An indication of the accuracy of hardness prediction. Predicted values based on the average
measured heat input for a given weld and on carbon equivalent relationships by:
(a) Yurioka –1 (b) Beckert, (c) combines data on to one graph.

You might also like