Download as pdf or txt
Download as pdf or txt
You are on page 1of 197

659

Transformer Thermal Modelling

Working Group
A2.38

June 2016
TRANSFORMER THERMAL
MODELLING
WG A2.38

Members
J. Lapworth, Convenor (UK), P. Picher, Secretary & Task Force Leader (CA)
Task Force leaders: J. Channet (FR), J. Kranenborg (SE), H. Nordman (FI), Z. Radakovic (RS), O. Roizman
(AU), K. Spoorenberg (NL), D. Susa (NO)
F. Berthereau (FR), H. Campelo (PT), S. Chen (FR), M. Cuesto (ES), V. Davydov (AU), G. Fleck (AT), T.
Gradnik
(SI), N. Gunter (ZA), W. Guo (AU), J.‐K. Kim (KR), J. Lee (KR), A. Portillo (UY), N. Schmidt (DE), E. Simonson
(UK), S. Tenbohlen (DE), F. Torriano (CA), F. Trautmann (DE), W. Van der Veken (BE), Z. Wang (UK), J.
Wijaya (AU), G. Wilson (UK), W. Wu (UK), S. Yamamura (JP).

Copyright © 2016

“All rights to this Technical Brochure are retained by CIGRE. It is strictly prohibited to reproduce or provide this publication
in any form or by any means to any third party. Only CIGRE Collective Members companies are allowed to store their copy
on their internal intranet or other company network provided access is restricted to their own employees. No part of this
publication may be reproduced or utilized without permission from CIGRE”.

Disclaimer notice

“CIGRE gives no warranty or assurance about the contents of this publication, nor does it accept any responsibility, as to the
accuracy or exhaustiveness of the information. All implied warranties and conditions are excluded to the maximum extent
permitted by law”.

ISBN: 978-2-85873-362-0
TRANSFORMER THERMAL MODELLING

ISBN: 978-2-85873-362-0
TABLE OF CONTENTS

Chapter 1:  Introduction .............................................................................................................. 1 


1.1  Importance of thermal design and development of thermal modelling ....................................... 1 
1.2  Objectives of the Working Group ................................................................................................ 2 
1.3  Basic transformer thermal modelling concepts ............................................................................ 2 
1.4  Definitions ................................................................................................................................... 4 
Chapter 2:  Temperature rises in steady state.......................................................................... 9 
2.1  Introduction .................................................................................................................................. 9 
2.1.1  General .................................................................................................................................... 9 
2.1.2  Thermal build of liquid-filled power transformers .................................................................. 9 
2.1.3  Standard design practice ........................................................................................................ 10 
2.1.4  Installation altitude influence ................................................................................................ 11 
2.2  Description of Q, S and H factors .............................................................................................. 11 
2.3  Eddy losses in windings ............................................................................................................. 14 
2.4  Stray losses in structural parts.................................................................................................... 17 
2.4.1  General .................................................................................................................................. 17 
2.4.2  Typical thermal experiences in the core and the metallic structural parts ............................. 17 
2.4.3  Operating conditions.............................................................................................................. 18 
2.4.4  Calculation of loss distribution .............................................................................................. 18 
2.4.5  Calculation of temperature distribution ................................................................................. 20 
2.4.6  The practical approach........................................................................................................... 20 
2.4.7  Conclusion ............................................................................................................................. 21 
2.5  Thermal-Hydraulic Network Model (THN) .............................................................................. 22 
2.5.1  Basics of thermal-hydraulic model ........................................................................................ 22 
2.5.2  Understanding the global pressure equilibrium in transformer ............................................. 25 
2.5.3  Basics of thermal model of parts of solid materials............................................................... 31 
2.5.4  Compiling hydraulic networks of complete transformer ....................................................... 34 
2.5.5  Final results of complete calculation based on detailed THN ............................................... 37 
2.5.6  Recommended concept of the software based on THN......................................................... 38 
2.5.7  Example of constructions which can be covered by software ............................................... 39 
2.5.8  Input data for the calculation ................................................................................................. 43 
2.5.9  Critical points and parameters in application of detailed THN ............................................. 46 

i
2.5.10  Illustration of use of THN as a calculation and analysis approach ........................................ 48 
2.6  Computational Fluid Dynamics ................................................................................................. 50 
2.6.1  Introduction ........................................................................................................................... 50 
2.6.2  CFD basic concepts and modelling approach ........................................................................ 51 
2.6.3  CFD application in transformer thermal modelling............................................................... 54 
2.6.4  CFD in winding heat transfer analysis................................................................................... 55 
2.6.5  CFD and THN validation ...................................................................................................... 57 
2.6.6  CFD and thermal model improvement .................................................................................. 59 
2.6.7  Global winding pressure drop and heat transfer correlations ................................................ 59 
2.6.8  Radiator modelling ................................................................................................................ 59 
2.6.9  CFD in tank wall computation............................................................................................... 62 
2.6.10  CFD and the design process .................................................................................................. 62 
2.6.11  Conclusion ............................................................................................................................. 62 
Chapter 3:  Benchmark of numerical tools ............................................................................. 65 
3.1  Introduction ................................................................................................................................ 65 
3.2  Description of the transformer ................................................................................................... 65 
3.3  Modelling specifications ............................................................................................................ 68 
3.4  Loss calculation ......................................................................................................................... 70 
3.4.1  Results ................................................................................................................................... 70 
3.4.2  Eddy losses calculations on a second transformer geometry ................................................. 71 
3.4.3  Discussion.............................................................................................................................. 73 
3.5  Temperature calculation............................................................................................................. 75 
3.5.1  Temperature calculation using uniform losses ...................................................................... 75 
3.5.2  Temperature calculation using non uniform losses ............................................................... 76 
3.5.3  Discussion.............................................................................................................................. 77 
Chapter 4:  Dynamic thermal modelling .................................................................................. 79 
4.1  Introduction ................................................................................................................................ 79 
4.2  Review of the state-of-the-art .................................................................................................... 79 
4.3  Loading Guide Dynamic Thermal Models (DTMs) .................................................................. 82 
4.3.1  General .................................................................................................................................. 82 
4.3.2  IEC 60354 model ................................................................................................................... 83 
4.3.3  IEC 60076-7 model ............................................................................................................... 83 
4.3.4  IEEE C57.91 Annex G model ............................................................................................... 84 
4.4  In-service accuracy evaluation of Loading Guide Dynamic Thermal Models .......................... 85 
ii
4.4.1  Description of the transformer under investigation ............................................................... 85 
4.4.2  Comparison of the three Loading Guide Dynamic Thermal Models .................................... 86 
4.4.3  Discussion and conclusion..................................................................................................... 93 
4.5  Transformer Dynamic Thermal Rating evaluation .................................................................... 93 
4.6  Standard exponents and constants of loading guide dynamic models ....................................... 95 
4.6.1  General .................................................................................................................................. 95 
4.6.2  IEC 60354 and IEC 60076-7 ................................................................................................. 96 
4.6.3  IEEE Annex G ....................................................................................................................... 97 
4.7  Uncertainties in Exponents Determination ................................................................................ 98 
Chapter 5:  Direct measurements .......................................................................................... 101 
5.1  Introduction .............................................................................................................................. 101 
5.2  Fiber-optic measurements in core-type transformers ............................................................... 101 
5.2.1  General ................................................................................................................................ 101 
5.2.2  Number of sensors ............................................................................................................... 103 
5.2.3  Location of sensors .............................................................................................................. 104 
5.2.4  Installation techniques ......................................................................................................... 107 
5.3  Fiber-optic measurements in shell-type transformers .............................................................. 112 
5.3.1  General ................................................................................................................................ 112 
5.3.2  Number of sensors ............................................................................................................... 112 
5.3.3  Location of sensors .............................................................................................................. 113 
5.3.4  Installation techniques ......................................................................................................... 115 
5.4  Measurement of the duct oil temperature next to the hotspot .................................................. 118 
5.5  Routing of the fiber-optic probes ............................................................................................. 119 
5.5.1  Through the phase insulation ............................................................................................... 119 
5.5.2  Through the cleats and leads ............................................................................................... 119 
5.6  Measurement of local temperatures in core and structural parts by thermocouples ................ 120 
5.6.1  General ................................................................................................................................ 120 
5.6.2  Number and location of thermocouples ............................................................................... 120 
5.6.3  Location of Thermocouples in Yoke Clamps ...................................................................... 121 
5.6.4  Location of thermocouples in flitch-plates and outer core packets ..................................... 122 
5.6.5  Location of thermocouples in the top yoke ......................................................................... 125 
5.6.6  Core temperature ................................................................................................................. 125 
5.6.7  Installation technique of thermocouples .............................................................................. 126 
5.6.8  An experience with fiber-optic installation in the yoke clamps .......................................... 129 
iii
5.7  Measurement of top-oil temperature with magnetically mounted sensor ................................ 131 
Chapter 6:  Shell-type transformer thermal modelling ........................................................ 133 
6.1  Introduction .............................................................................................................................. 133 
6.2  Design aspects .......................................................................................................................... 133 
6.3  CFD models of shell-type units ............................................................................................... 138 
6.3.1  General ................................................................................................................................ 138 
6.3.2  CFD simulations of the global phase ................................................................................... 139 
6.3.3  CFD simulations inside coils ............................................................................................... 141 
6.3.4  CFD flow field validation with PIV .................................................................................... 144 
6.4  Leakage flux and electromagnetic losses calculation .............................................................. 146 
6.4.1  General ................................................................................................................................ 146 
6.4.2  In coils ................................................................................................................................. 146 
6.4.3  In other parts (tank, T-beams, shunts…) ............................................................................. 150 
6.5  Hotspot calculation (coupled CFD simulations) ...................................................................... 152 
6.6  Direct measurements (FO location) ......................................................................................... 154 
6.7  Hotspot evidences .................................................................................................................... 156 
6.8  Summary remarks .................................................................................................................... 158 

References ............................................................................................................................................... 159 


Annex A Installation altitude influence on transformer thermal modelling ........................................... 167 
Annex B Guideline to CFD simulations ................................................................................................ 175 
Annex C Examples of fiber-optic measurements ................................................................................... 183 

iv
SUMMARY

CIGRE Working Group A2.38 ‘Transformer Thermal Modelling’ was set up in 2008 with the primary
objective of investigating the estimation of transformer winding hotspot temperatures by calculation and
direct measurement.

The ‘state of the art’ regarding transformer thermal modelling has been investigated practically by inviting
Working Group members to perform hotspot temperature calculations for a test case involving a disc type
winding for a core-type transformer cooled by directed ‘zig-zag’ cooling arrangement under natural oil
flow (ON) conditions.

Calculations of stray winding losses induced by leakage fields have been carried out by members for the
defined winding geometry using proprietary transformer design tools. This has revealed an unexpectedly
large variation in stray loss predictions, by about ±40% of the average calculated value, the reason for
which is being investigated. Clearly this represents a significant source of potential uncertainty in
predicted hotspot temperatures.

Calculations of winding temperature distributions for an agreed loss distribution and inlet cooling
conditions have been carried out by members using both thermal-hydraulic network (THN) and
computational fluid dynamics (CFD) models. THN modelling uses a lumped parameter network type of
representation of the arrangement of cooling ducts, while CFD modelling uses a very fine spatial
discretisation which allows details of flows within ducts to be modelled. Not surprisingly, when stray
losses were included, it was unanimously agreed that the hotspot temperature would occur in the top disc,
and there was good agreement about hotspot temperature rise. For this test case the hotspot temperature
rise above ambient was about 105K compared with an average winding rise of only 62K, because of the
unusually high hotspot loss factor (Q = 2.5). There was however a significant difference between THN
and CFD models regarding predicted oil flow distributions, with CFD models predicting lower oil flows
towards the tops of passes. This is attributed to the so-called ‘hot streak’ effect, in which CFD models
predict that hot oil exiting the cooling ducts stays close to the outer face of the winding rather than mixing
with colder oil already in the outlet duct, thereby experiencing a higher hydraulic resistance.

Therefore, it would appear that there is the possibility of a fundamental difference between ON and OD
cooling for such directed ‘zig-zag’ cooling oil flow arrangements: that for ON flow conditions, minimum
flows and maximum temperatures could occur in discs towards the top of each pass rather than towards
the middle of each pass, as expected for OD. Moreover, rencent CFD investigations seemingly show that,
for OD cooling with high flow velocities, a stagnant flow or even a reverse flow can be observed in the
first horizontal duct of a pass. This phenomenon can induce a significant overheating of the bottom disc in
the pass which sometimes becomes the hottest disc in the winding.

Both THN and CFD modelling require a large amount of effort even when automatic model generating
facilities are available and are not reasonably practicable for day to day design. THN models require
considerably less effort, but whereas they can model the distribution of oil flows between cooling ducts,
they cannot model the details of oil flows within those ducts. A compromise approach may be to include
into THN models empirical correlations for pressure losses and heat transfer which take into account the
effects of flow details on overall cooling.

A comprehensive and well‐referenced introduction to state‐of‐the‐art of dynamic thermal modelling is


given. Loading Guide type algorithms for calculating transient temperatures have been reviewed, in

v
particular the new formulation proposed in the latest edition of IEC 60076‐7, which includes a new ‘over‐
shoot’ gradient function and a change to the recommended winding exponent. It is concluded that further
research and development is needed to improve the existing loading guide models, in particular, to
increase the modelling accuracy during sub‐zero ambient condition where the oil viscosity effect is
dominant.

Recommended practice for installing fibre optic temperature sensors to measure hotspot temperatures has
been described in terms of method of installation, location and number. Examples of results obtained
from such direct measurements have been given. It is shown that in addition to checking hotspot
temperature rises during temperature rise tests, valuable information about other key thermal parameters,
e.g. time constants and exponents, can be obtained. It is suggested that in addition to checking steady
state hotspot temperatures, transient temperature changes should be recorded and analysed. In particular,
with fibre optic sensors fitted there is the possibility of obtaining valuable extra information during the
initial five hour heating up period when full load losses are supplied to establish steady state oil
temperatures.

Recommended practice for installing thermocouples to measure local temperatures in the core and
structural parties has also been described. This practice is assumed to happen rarely and only on selected
transformers as a means to safeguard quality and to tune the calculation programs.

The state of the art for thermal modelling of shell type transformers has also been investigated. Such
transformers have fundamentally different winding geometries to core-type units. Calculation of stray
losses and oil flows is more complicated, requiring 3D and CFD modelling, and examples are described.
Typically winding hotspots occur on the inside of the winding pancakes, particularly above the upper ‘T’
beams, because of the possibility of local oil flow minima, and the location of maximum stray losses
caused by leakage fields. An experimental investigation using optical fibre sensors installed in a large
single phase generator transformer with OD cooling has indicated hotspot factors of 1.3 to 1.9.

Experience of calculations and direct measurements for large core or shell type transformers suggests that
the hotspot factor H can often be in the range 1.5 to 2.0, or even higher, i.e. rather higher than the typical
range of 1.1 to 1.3 suggested in a previous version of IEC 60076-2, but it should be possible to bring this
below 1.5 with good design. Some of the commonly used solutions for good thermal designs are
discussed. Perhaps the most important factor affecting hotspot factor is strand height.

Lastly, although not the primary objective of the Working Group, the management of hotspot
temperatures in other structural steel parts, e.g. core, clamping and tank, has also been reviewed and
general recommendations given.

In conclusion it can be said that today techniques exist to enable transformer winding hotspot temperatures
to be predicted with good accuracy. Hotspot factors can be estimated by detailed mathematical models
and verified experimentally by direct measurement using fibre optic temperature sensors. In view of the
time and effort required, it is not reasonable to expect such modelling to be carried out for day-to-day
designs, but it seems reasonable that manufacturers should be using such techniques to verify hotspot
factors for their mainstream designs and any significant design variants.

vi
Chapter 1: Introduction

Chapter 1: Introduction

1.1 Importance of thermal design and development of thermal modelling

Good thermal design is a very important aspect of overall design since it determines insulation ageing and
life. Unfortunately, largely because it is critically dependent on local performance at the winding hotspot,
it is not an aspect which can be entirely adequately checked at the design or manufacturing stage, and as a
result problems may only become evident after a premature insulation failure or when a transformer is
eventually scrapped, when insulation samples can be taken for DP analysis. Good thermal design is also
in the interest of the manufacturer, since an inefficient design will represent a poor use of material,
resulting in a transformer with a higher thermal rating than required and consequently an over expensive
product.

Traditionally, maximum temperature rises have been specified by standards and checked by measuring
mean winding temperature rises by resistance measurements at the end of factory heat run tests. Thermal
design practice had been essentially limited to deriving key temperature rises by extrapolation from
existing designs. Hotspot temperatures were estimated using assumed hotspot factors. Unfortunately, as
transformers become larger and more sophisticated, the details of design and manufacturing become more
critical. Fortunately, with the development of powerful computers and sophisticated modelling programs
it has become possible to improve the detailed design. Recently there has been evidence of a new failure
process, namely copper sulphide deposition, which although is primarily due to chemical activity, has a
strong temperature dependence, and as a consequence predominantly occurs in heavily loaded
transformers with a particular cooling mode, and particularly at certain locations in the windings. Also,
there has been concern about new manufacturers entering the market with less experience in the field of
thermal modelling.

When checking the thermal design it is the steady state temperature rises that are of primary concern, in
particular the average winding rise above oil temperature or ‘gradient’. However, utilities are also
interested in transient performance and overload capability, particularly for transmission and distribution
transformers which see significant daily load variations, and also for transformers which have switchable
cooler modes. So-called ‘loading guide’ algorithms have been agreed for calculating transient
temperatures, based on component winding and oil temperature rises derived during factory heat run tests
and additional parameters such as time constants and exponents to be derived.

Whereas it is desirable to use a common approach to thermal modelling for all transformers, in practice
the influence of detailed design, which can vary considerably, has to be taken into account. For instance,
transformers can be one of two fundamentally different geometries: core type and shell type. All shell-
form transformers have essentially the same winding and cooling geometry, although the pancake coils
may be horizontally or vertically disposed. Core-type transformers can have a variety of winding
arrangements, essentially either ‘layer’ type in which the windings are wound in a series of concentric
cylinders around the core separated by axial cooling ducts or ‘disc’ type in which the windings are wound
in a series of radial discs separated by radial cooling ducts. Aside from winding geometry, another
important factor affecting cooling is whether steps are taken to control or ‘direct’ the flow of cooling oil
entering the tank through the windings. Last but not least, the oil flow may be ‘forced’ by an external
pump or set up by ‘natural’ internal buoyancy forces, and the performance of external coolers may be
enhanced by forced cooling.

Recently, efforts to improve thermal designs have concentrated on improving the estimation of the hotspot
factor, either by using detailed mathematical models which take into account the detailed distribution of
1
Chapter 1: Introduction

losses and cooling, or by the use of fibre-optic sensors to measure the temperature at the assumed hotspot
location.

1.2 Objectives of the Working Group

In the Terms of Reference the objectives of the WG were defined as:

 examine and discuss the ‘state of the art’ in transformer thermal modelling, in particular the
estimation of winding hotspot temperature and identify critical parameters which affect the
accuracy of modelling;

 describe practical examples of thermal modelling and limitations

 derive typical hotspot factors for various common cooling arrangements

 identify any fundamental differences between naturally (ON) and forced (OF) cooled, whether or
not directed oil flow is used and critical factors

 discuss dynamic thermal models for calculation of transient temperatures for transformers subject
to conditions of variable load and temperature of cooling medium: including consideration of
protection, monitoring, estimation of overloading possibility, estimation of ageing based on
temperature values, etc.

 identify good practices for direct measurements of hotspot temperatures and illustrate use for
checking temperature rises and deriving other thermal parameters required for modelling

 include a consideration of shell-type transformers, for which there are fewer technical publications

 make recommendation for standard improvement

1.3 Basic transformer thermal modelling concepts

The basic thermal diagram for oil cooled transformers is shown in Figure 1.1.

2
Chapter 1: Introduction

A Top-oil temperature derived as the average of the tank outlet oil temperature and the tank pocket temperature
B Mixed oil temperature in the tank at the top of the winding
C Temperature of the average oil in the tank
D Oil temperature at the bottom of the winding
E Bottom of the tank
gr Average winding to average oil (in tank) temperature gradient at rated current
H Hotspot factor
P Hotspot temperature
Q Average winding temperature determined by resistance measurement

y-axis representing relative position (height) and x-axis representing temperature

Figure 1.1: Thermal Diagram from IEC 60076-7 standard [21].

The temperature rise of a winding above ambient is built up from three component temperature rises: the
temperature rise of the inlet cooling oil above ambient, the temperature rise of the cooling oil as it passes
through the transformer, and the temperature rise of the winding above the cooling oil; which are derived
from three key measured transformer temperatures: inlet and outlet oil temperatures and mean winding
temperature.

The temperature increase of the oil on passing through the transformer can be approximated by the
difference between inlet and outlet oil temperatures, but this does not take into account the fact that there
are usually more than one parallel oil paths within the transformer: at least one for each winding, each
with their own temperature rise dependent on heat removed. Whereas the parallel oil paths will have a
common inlet temperature, their outlet/exit temperatures will be different until they merge and mix,
adopting the common ‘top oil’ or cooler inlet temperature. Temperature pockets in the top of the
transformer tank are usually provided to enable this temperature to be measured, but the measurement of
oil temperature in the cooler inlet manifold is perhaps better. In addition to the inevitable uncertainty in
top oil temperature arising from the fact that there is usually more than one winding, another common

3
Chapter 1: Introduction

reason is that there can also be significant parallel paths that bypass the windings completely, e.g. for non
directed forced (OF) cooling.

The temperature rise of the winding above the oil is referred to as the ‘gradient’, g. A mean value for this
is obtained from the difference between the mean winding temperature, usually determined by resistance
measurements, and an estimate of the mean oil temperature for the winding (which is subject to the
abovementioned uncertainty in winding exit oil temperature). The gradient can vary with position because
of local variations in winding losses and cooling effectiveness. The highest winding temperature, the so-
called ‘hotspot’ temperature, is usually at the top of the winding because this is where the cooling oil
passing through the winding is hottest, and because at the ends of the windings there are extra losses due
to eddy currents set up by leakage fluxes, and there may be extra insulation or cooling oil flow
differences. Because of these factors the temperature rise at the hotspot (h) above the local cooling oil
is expected to be greater than the mean winding gradient (g), as expressed by a hotspot factor, H, defined
by

h = o + Hg (1.1)

Recently, H has been further expressed as the product of two factors:

H = Q.S (1.2)

where

Q: hotspot factor related to additional loss

S: hotspot factor related to efficiency of cooling

These sub-components will be discussed in more detail later.

The fundamental objective of transformer thermal design is to be able to accurately predict winding
temperatures and component temperature rises to ensure they are within specification. The primary
requirement is to meet the specified mean winding temperature rise above ambient, usually 65 K, which is
checked by winding resistance measurements made at the end of a factory heat run test. Traditionally,
hotspot temperatures were estimated based on mean winding gradients derived from heat run test
measurements and assumed hotspot factors. Increasingly nowadays, customers expect manufacturers to
meet predictions made before the transformer is tested, and to have good evidence for hotspot factors,
either based on hotspot temperature measurements made using point temperature sensors, or by
calculation. Therefore, the two most important thermal parameters for the transformer designer are
winding gradient, g, and hotspot factor, H.

1.4 Definitions

Thermal model: mathematical description of temperature distribution within transformer insulation and
current conducting systems.

Thermal modelling: process of creating a thermal model.

4
Chapter 1: Introduction

THN (or THNM or TNM): thermal-hydraulic network model.

Computational Fluid Dynamics (CFD): branch of fluid mechanics that uses numerical methods and
algorithms to solve and analyse problems that involve fluid flows and heat transfer.
Finite Element Method (FEM): numerical technique for finding approximate solutions to partial
differential equations (PDE) and their systems

Note: One of the discretization methods used in CFD. The other common methods are finite volume and
finite differences methods.

Temperature rise: difference between the temperature of the part under consideration and the
temperature of the cooling medium at the intake of the cooling equipment, for air-cooled or water-cooled
transformers or reactors.

Temperature rise test (heat run test): procedure for determination of temperature and temperature rise
values during factory testing of a transformer by supplying a rated losses for oil temperature
determinations and rated current for winding temperature determinations.

Top-oil temperature: temperature of the top layer of the insulating liquid in a transformer, representative
of the temperature of the top liquid in the cooling flow stream.

Top-oil temperature rise: arithmetic difference between the top-oil temperature and the ambient
temperature.

Ambient temperature: temperature of the medium such as air or water into which the heat of the
transformer is dissipated.

Average winding temperature: average temperature of a winding as determined from the resistance
measured across the terminals of the winding.

Average winding temperature rise: arithmetic difference between the average winding temperature of a
winding and the ambient temperature.

Bottom-oil temperature: temperature of the liquid in a liquid-immersed transformer as measured at an


elevation just below the bottom of the coils or in the oil flowing from the liquid cooling equipment into
the transformer.

Bottom-oil temperature rise: arithmetic difference between the bottom-oil temperature and the ambient
temperature.

Hotspot: if not specifically defined, “hotspot” means the hottest-spot temperature of the winding.

Hotspot temperature rise: arithmetic difference between the hotspot temperature and the ambient
temperature.

Hotspot factor (H factor): dimensionless factor to estimate the increase of the average winding gradient
due to the local increase of additional loss and variation in the liquid flow stream.

Note: H factor is obtained by the product of the two following Q an S factors.

5
Chapter 1: Introduction

Q factor: dimensionless factor to estimate the increase of the average winding gradient due to the local
increase of the additional loss.

S factor: dimensionless factor to estimate the local increase of the average winding gradient due to the
variation in the liquid flow stream.
Core loss: power dissipated in a magnetic core subjected to a time-varying magnetizing force. Core loss
includes hysteresis and eddy-current losses of the core.

Eddy current loss: power loss in conductors resulting from magnetic leakage field in windings and from
the flow of eddy currents in parallel windings or in parallel winding strands.

Note: There is no test method to determine individual winding eddy loss or to separate transformer stray
loss from eddy loss. The total stray and eddy loss is determined by measuring the total load loss during the
impedance test.

Stray loss: power loss that is due to the stray leakage flux, which introduces losses in the core, clamps,
tank, and other structural parts.

Note: There is no test method to determine individual stray loss or to separate transformer stray loss from
eddy loss. The total stray and eddy loss is determined by measuring the total load loss during the
impedance test.

Load loss:

a) of two-winding transformers (for the principal tapping):

active power absorbed at rated frequency when rated current is flowing through the line
terminal(s) of one of the windings, the terminals of the other winding being short-circuited, and
any winding fitted with tappings being connected on its principal tapping

b) of multi-winding transformers, related to a certain pair of windings (for the principal tapping):

active power absorbed at rated frequency when a current flows through the line terminal(s) of one
of the windings of the pair, corresponding to the smaller of the rated power values of both
windings of that pair, the terminals of the other winding of the same pair being short-circuited,
any winding of the pair fitted with tappings being connected on its principal tapping and the
remaining winding(s) being open-circuited

No-load loss: active power absorbed when a given voltage at rated frequency is applied to the terminals of
one of the windings, the other winding(s) being open-circuited.

Total losses: sum of the no-load loss and the load loss.

Note: For multi-winding transformers, the total losses refer to a specified loading combination.

ONAN/ONAF: cooling class for a transformer having its core and coils immersed in insulating liquid and
having a self-cooled rating with cooling obtained by the natural circulation of air over the cooling surface,
and a forced-air-cooled rating with cooling obtained by the forced circulation of air over this same cooling
surface.

6
Chapter 1: Introduction

Note: ONAN/ONAF was previously termed OA/FA. For this cooling class, there is a natural
thermosiphon oil flow through cooling equipment and in windings

ODAF: cooling class for a transformer having its core and coils immersed in insulating liquid and cooled
by forced circulation of the insulating liquid utilizing directed flow

Note: The insulating liquid is cooled by external insulating liquid-to-air heat-exchanger equipment
utilizing forced circulation of air over its cooling surface. ODAF was previously termed FOA. For this
cooling class, there is a forced oil circulation through cooling equipment and forced oil circulation into at
least the main windings.

OFAF: cooling class for a transformer having its core and coils immersed in insulating liquid and cooled
by forced circulation of the insulating liquid utilizing non-directed flow

Note: The insulating liquid is cooled by external insulating liquid-to air heat-exchanger equipment
utilizing forced circulation of air over its cooling surface. OFAF was previously termed FOA. For this
cooling class, there is a forced oil circulation through cooling equipment and thermosiphon oil flow in
windings.

Shell-type transformer: transformer in which the laminations constituting the iron core surround the
windings and usually enclose the greater part of them.

Core-type transformer: transformer in which those parts of the magnetic circuit surrounded by the
windings have the form of legs with two common yokes.

7
Chapter 2: Temperature rises in steady state

Chapter 2: Temperature rises in steady state

2.1 Introduction

2.1.1 General

Different types of models are available for modelling temperatures in power transformers. The type of
transformer where it is to be applied and its different design phases influence the choice.

Different types of transformers can then be eligible for a thermal analysis:

 New transformers

 Retrofit transformers where some modifications were made on the active part or the oil

 Analysis of existing transformers to confirm or increase the loading capability

The amount of design information may be very different for these cases. For example, a retrofitted
transformer where the oil is replaced may only have heat-run test information available and thus only a
simple thermal model will apply. For a new unit being manufactured the full design is known and then it
is important to use state-of-the art models to pursue accurate temperature predictions.

Furthermore, in case of new transformer, the design process can be divided in the quotation phase and the
design phase, each phase having different requirements:

 Quotation phase: the transformer can be built economically satisfying the temperature limits. In
this phase the amount of cooling equipment needed is determined. Hotspot calculations are not
performed in detail since the design details necessary for an accurate prediction are not yet
known.

 Design phase: all design details should be accounted for and calculated accurately. The
determination of all temperatures have to be guaranteed according to standards, including hotspot
temperature.

As a consequence, thermal models are applied in different levels of detail being the major subject of
Section 2.1.3 the suitable and then recommended approaches for accurate temperature calculation.

Whatever the particular application is, a proper start of the discussion is the IEC standard, since it relates
the thermal parameters that play a role in the heat run test – Section 2.1.2.

2.1.2 Thermal build of liquid-filled power transformers

The thermal build inside liquid-filled power transformers can be well understood by analysing the thermal
diagram as in IEC60076-7 (Figure 1.1). This diagram is understood to be a simplified representation of a
more complex temperature distribution.

The measurable quantities during a heat run test comprise the three dotted quantities B, D and Q. With
these three measurable quantities the average winding thermal gradient, gr, can be estimated. Afterwards

9
Chapter 2: Temperature rises in steady state

the hotspot temperature is obtained by multiplying the thermal gradient, gr, by an empirical hotspot factor
- H.

During many years, and due to the practical limitations of embedded temperature sensors, IEC did not
recommend their direct installation inside the windings.

More recently the safe operation of temperature sensors inside power transformers became possible and
measurements can be made directly in the winding in the area where winding hotspot is identified with
design tool – point P – Figure 1.1.

With direct measurements of hotspots and, as reported in the previous CIGRE WG 12.09, cases were
revealed where the standard empirical factors have been overruled, exhibiting values ranging up to 2.1.

This also motivated the present WG to analyse comprehensively the best methodologies being put in
practice in order to enhance power transformer thermal designs and furthermore thermal prediction
capabilities.

Since an accurate prediction is needed of directly measured hotspot temperatures, this Brochure also
includes a Chapter on guidelines for the placement of winding optical fibre sensors – Chapter 4.

2.1.3 Standard design practice

In order to calculate the temperature distribution inside power transformers, a manufacturer may typically
rely on three main methods:

 correlation methods,

 Thermal Hydraulic Network Models (THN), and

 CFD (Computational Fluid Dynamics).

A key requirement for a design tool in the design process is that it implements a proper balance of
accuracy and simulation speed. The design of a complex product like power transformers involves a large
number of design rules to be checked, covering the various technical areas (electrical insulation, thermal,
mechanical, noise, etc.). This implies that in most cases the thermal calculations should be performed in a
time of the order of only a few minutes at most.

With the current level of computer technology this requirement implies that both correlation methods and
THN are applicable for the general design practice. The THN approach is dealt in detail in Section 2.5.

CFD is a relatively new approach, mostly focussing on the active part. In this approach the geometry of
the active parts is modelled in very fine detail (much finer than the geometry dimensions) and the
equations for heat and mass transfer are expressed and solved accordingly. Consequently a very detailed
resolution can be obtained that is much finer than THN but comes at the cost of a much larger calculation
time (typically of the order of hours) and requires specialist knowledge in order to be applied. The CFD
approach, its usefulness and its relation to other methods is discussed in detail over Section 2.6.

The three methods represent a top-down list regarding detail level however these methods are:

10
Chapter 2: Temperature rises in steady state

 not exclusive: in most of the cases they are complementary as they might be used in different
design process phases (i.e. correlation methods for quotation and THN for the design phase);

 neither independent: namely THN and even correlation methods, as they can include information
extracted from CFD ‘experiments’.

It is clear that, due to its mathematical and numerical robustness, CFD is the most powerful method
available to analyse the cooling of power transformer windings. However its direct applicability in this
moment in the design process is unfeasible and not envisaged in the mid-term due to the conclusions
explained in Section 2.6.11.

Customer’s requirements and expectations must be also taken into account, especially for Design Review
documentation that may involve performing specific temperature calculations or simulations.

There are, however, practical limits to the accuracy that can be obtained with a model. Transformer
manufacturing involves not only an analytical design process but also has to cope with uncertainties and
tolerances in materials, manufacturing constraints, etc. Consequently, even models that are based on the
best mathematical descriptions of the underlying physical processes will always need to incorporate a
certain degree of empiricism and tuning factors (based on the manufacturer’s experience) to keep the final
temperature differences between measured and calculated values within acceptable limitations.

Notwithstanding this fact, the methods mentioned and herein extensively analysed in this technical
Brochure – namely CFD and THN - compose a significant step forward towards a better understanding of
the thermal performance of power transformers, thus towards its optimized design and operation.

2.1.4 Installation altitude influence

There are a lot of countries with transformers operating at sites above sea-level. The industry is involved
in develop better winding hotspot estimations. The oil temperature is an important factor in these
estimations and this temperature is greatly affected by installation altitude of the transformer equipped
with AN or AF coolers.

The transformer thermal models must be able to represent the installation altitude influence in the values
of calculated temperatures.

The thermal models must have an input for the installation altitude of the transformer and if we change the
installation altitude from 0 meters above sea-level to, for instance, 3000 meters above sea-level, the
calculated temperature values must change.

The installation altitude influence on transformer thermal modelling is further discussed in Annex A.

2.2 Description of Q, S and H factors

The IEC 60076-2 (2011) [1] has defined the hotspot factor H, as “a dimensionless factor to estimate the
local increase of the winding gradient due to the increase of additional loss and variation in the liquid
flow stream”. IEC 60076-2 notes also that the hotspot factor H is obtained by the product of the Q and S
factors.

The Q and S factors are dimensionless factors described in this standard as:

11
Chapter 2: Temperature rises in steady state

 Q is “a dimensionless factor to estimate the increase of the average winding gradient due to the
local increase of additional loss.”

 S is “a dimensionless factor to estimate the local increase of the average winding gradient due to
the variation in the liquid flow stream.”

According these definitions Q should be calculated by modelling the winding with the correct heat loss
distribution, but with uniform oil velocity. The Q factor is then the ratio of the maximal winding to local
oil gradient over the average winding to average oil gradient.

On the other hand S should be calculated by modelling the winding with uniform heat losses and with the
correct oil velocity inside the winding. The S factor is the ratio of the maximal winding to local oil
gradient over the average winding to average oil gradient.

However after calculating Q and S in line with the IEC definition, the hotspot factor H cannot be
calculated directly as the product of Q and S factor [129], as mentioned in IEC standard, because :

- Disc with maximal Q factor and disc with maximal S factor can (and will probably be) different
discs

- When modelling the winding with correct heat loss distribution and oil flows, Q and S factor will
not be independent from each other as explained later.

Because of above reasons the working group proposes more practical definitions of Q, S and H factor.

The H factor can be derived out of Figure 1.1 and is:

h  o
H
o  b (2.1)
w 
2

This formula (2.1) is different from the current IEC 60076-2 definition, because the hotspot temperature is
referenced to the mixed top oil, while increase in local winding to oil gradient refers to local winding oil.

Formula (2.1) will be used in the remainder of the document to calculate the hotspot factor.

This formula for H has the following advantages:

- H-factor can be calculated directly out of the calculation results (calculated with correct loss and
oil flow distribution)
- This is the correct hotspot factor to predict the hotspot temperature out of the temperature rise
results, obtained in the standard heat run test.
- This hotspot factor can also be checked in case fibre optic measurements are made (Thotspot is
known)
- Hotspot is not always located at the top of the winding. However this formulation is a practical
solution to overcome this issue.
The Q-factor used in this document is a dimensionless factor as a ratio of two losses, and in cylindrical
coordinates be defined as:

12
Chapter 2: Temperature rises in steady state

Q  Qr , z,  , T  Qave (2.2)

where Q(r,z, φ, T) [W/m3] is the local loss density at a location where r is the radial position, φ is the angle
in circumferential position, z is the axial position, T is the local temperature at (r, z, φ) and Qave is the
average loss of the winding at average temperature.

For calculation purposes, one can redefine that for a disc winding in which each disc has several
conductors in radial direction and consists of numerous discs in axial direction, as:

Q  Qconductor _ number _ in _ disc, disk _ number,  , T  Qave (2.3)

The Q factor is a scalar function [2] and is based on the steady state condition of a defined loading at a
defined tap position (if applicable).

It is important to note that the Q-factor in this definition is not a ratio of temperatures but a ratio of losses.

Finally the S factor used in this document is defined as :

H (2.4)
S
Q

which can be easily calculated as soon as H and Q are known.

This S factor is a size for the local cooling inefficiency. Higher S factor means higher local temperature
gradient thus worse cooling efficiency.

According to the current IEC calculation this S factor should be calculated as the ratio of local hotspot
gradient over winding gradient with constant heat losses. With this definition the S factor is proportional
with the ratio of two thermal resistances, resulting in:

S  S r , z,  , T  S ave (2.5)

where S(r, z, φ, T) [K/W] is the local cooling resistance and Save [K/W] is the average cooling resistance.

We should note that heat transfer can be in different directions. The (overall) local heat transfer consists of
series and parallel parts, such as:

 The insulation between the neighbouring conductors, that are in direct contact with each other, in
radial direction.

 The insulation paper and oil boundary layer between conductor and the oil flow in axial direction.
Note that heat transfer functions for oil boundary layer are often a function of the heat flux.

 The copper (which almost can be neglected) in the tangential direction.

This implies that the Q and S factors are not fully independent, because they are linked by temperature,
heat flux, etc. E.g. if the local losses are higher, the local temperature will also increase and will influence
the local flow stream and the local convection heat transfer coefficient from conductor to oil. This is
illustrated by the THN calculations on the 66 MVA transformer winding described in Chapter 3.
13
Chapter 2: Temperature rises in steady state

 Remark 1: It should be noted that the size of one (paper insulated) conductor is the smallest
element in which the losses are calculated. There is the same temperature inside each element, so
thermal resistances inside the element are neglected. In the case the Q factor is based on a number
of conductors in one (or sometimes even 4) top discs, it increases in essence the element size to a
large extend and it neglects the temperature distribution between conductors in the disc (and even
between discs), which results in a too low estimate of the hotspot. The approach of using one or
more discs as smallest element results in a too low estimate of the hotspot temperature and must
be rejected.

 Remark 2: In the case of a high Q factor in a transformer, one is able to limit the hotspot factor by
creating locally more cooling surface and so design for a low S factor at that location, see Figure
2.13 (winding types). This principle is easy to do by adding an axial cooling channel inside a
radial spacer disc or by adding a radial spacer inside a winding with axial cooling channels. The
location of the hotspot does not necessarily have to correspond with the location of the maximum
losses.

Example: In case that the upper disc – with highest Q factor - of the LV winding (Chapter 3 – radial build
50 mm, axial height 15 mm) would have had one extra axial cooling duct of 4 mm inside the disc, the
surface heat flux density would have decreased with a factor [2*(50+15) ]/ [ 2*(50+15)+2*15] = 0.82. The
S-factor would then be reduced and if the losses per disc remained the same, then the hotspot factor of this
disc would be lower and probably the hotspot would move to another disc (Table 3.4).

2.3 Eddy losses in windings

The first step prior to performing winding temperature calculation consists in determining the amplitude
and distribution of electromagnetic losses. The accuracy of loss calculations may depend on the level of
details used in the modelling approach. Moreover, it is not possible to formally validate the simulation
results using load-loss test data because eddy-current losses in the windings cannot be separated easily
from stray losses in other metallic parts (tank, core clamps, etc.).

To calculate the magnetic field is a straight forward mathematical procedure, especially when one can
neglect the magnetic field due to eddy currents in all the conducting parts. These conducting parts consist
of the copper of the windings, the steel-press construction of the windings, the core and the tank wall
including the magnetic shunts or aluminium/copper shields.

The Finite Element Method (FEM) is quite commonly used in the transformer industry for calculating the
eddy losses. Approximation of the 3D electromagnetic field of the transformer using a 2D model requires
selecting a projection plane and a co-ordinate system (axisymmetrical or cartesian).

The windings are divided into many rectangular sections with a uniform ampere-turn distribution. The
conductivity value is not defined for these sections since it is assumed that the eddy currents do not
influence the leakage field (assumption valid for thin conductors). The values of the axial and radial flux
densities for each conductor can be obtained from the FEM solution, whereupon the axial and radial
components of the eddy losses are calculated for each conductor using Equation (2.6).

 2 B 2T 2
P (2.6)
24 

14
Chapter 2: Temperature rises in steady state

where B is the peak leakage flux density in V·s·m-2,  = 2πf where f is the frequency in s-1, T is the
conductor dimension perpendicular to the direction of the leakage flux density in m and ρ is the resistivity
in V·A-1·m. The axial and radial flux densities are assumed to be constant over a single conductor and
equal to the value at the center of the conductor. The total eddy loss for each winding is calculated by
integrating the loss components of all its conductors.

It should be noted that Equation (2.6) was derived assuming a small conductor size compared with the
depth of penetration. Proprietary software from the manufacturers may use an alternative equation for the
losses calculation based on their own experience.

The following describes how the boundary conditions selection can influence the results. A transformer is
a 3D construction, in which the windings are (almost) rotationally symmetric, but the surrounding steel
parts (core and tank) are not (see Figure 2.1). In this case the inner low voltage winding has a big pitch,
and is certainly not rotationally symmetric.

Figure 2.1: 3D picture of a three phase transformer with 5-legged core and LV winding with large pitch.

The magnetic field in the winding can be determined more accurately by making a number of 2D
calculations. Looking at the top view of this active part, four 2D cross sections defined as A, B, C and D
can be made (Figure 2.2). It is allowed to make these 2D calculations because the dimensions that
determine the radial magnetic flux density at the top of the winding are much smaller than some other
dimensions, like winding height and average diameter. The 2D projections are shown in Figure 2.3.

15
Chapter 2: Temperature rises in steady state

Figure 2.2: Top view of active part between the two tank walls.

Cross section A Cross section B Cross section C Cross section D

Figure 2.3: Four 2D cross sections of the active part.

The critical parameters to determine the Q-factor, apart from the selected cross section, are:

1. The winding heights due to manufacturing tolerances.

2. The location of the tap winding, because it can be located:

 between the inner main winding and the core

 in the main gap between inner- and outer winding

 at outside of the outer winding

 as taps at the face of the outer winding

3. The type of tap winding arrangement, like plus/minus regulation or coarse/fine regulation.

4. The location and the value of the max hotspot temperature in the transformer depends on the tap
position and can even shift from one main winding to the other, especially in the case of
autotransformers
16
Chapter 2: Temperature rises in steady state

2.4 Stray losses in structural parts

2.4.1 General

The literature on stray field losses in structure metal parts of a transformer is very extensive. However to
conduct a literature review would be elaborate and time consuming for reasons such as:

 The constructions and assumptions, described in the collective works, are often very specific and
not all the information is disclosed.

 Quite often material parameters or boundary conditions are altered to match calculation with test
results.

 Results are usually presented as temperature rise over ambient. The latter is less sensitive
compared with the temperature rise over local oil temperature that potentially has a larger
measurement error.

The loss distribution involves a scalar function P(r,φ,z) [W/m3] over the metal parts in the transformer
volume. The cooling can be considered as a vector function C(r,φ,z) [K/(W/m2)] over the total
transformer volume [2]. These loss and cooling calculations require a great deal of (computing and
modelling) effort.

A more practical approach is based on tacit experience and know-how on the relevant parameters and the
(usually known) location of the hotspot by performing measurements by means of thermal couples or fibre
optic temperature sensors to determine the hotspot temperature rise

The loss distribution in the tank wall, either due to the stray field of the windings or the high current leads,
is not taken into account in this Brochure. Localised high temperatures of the tank wall due to the lack of
shielding or the absence of non-magnetic material in the tank cover in the case of high current leads are
easy visible by an infra-red scan.

2.4.2 Typical thermal experiences in the core and the metallic structural parts

The requirements in [1] as described in the following sentence should be noted:

“No numerical limits are specified for the temperature rise of magnetic core, bare electrical connections,
electric or magnetic shields and structural parts in the tank. However, a self-evident requirement is that
they shall not reach temperature which will cause damages to adjacent parts or undue ageing of the
insulation liquid”.

In [3], a good overview is presented on many aspects of stray flux, however the overview as presented
here, focuses more on the location of the metal part in relation to the winding set as described hereafter.

a) At the inner diameter of the winding set assembly:

 stray flux field penetrating on the core leg, perpendicular to the core sheets;

 stray flux field penetrating on the tie-plate at the core (where applicable).

b) Above and under the winding set assembly:


17
Chapter 2: Temperature rises in steady state

 Stray field penetration on the press frame. The press frame is a steel construction above
and below the phases and can be a preferred path for the magnetic stray flux between the
phases. The press frame sometimes is often contracted of one element (yoke-beam) for
all phases or sometimes each phase has its own individual press frame. In the latter case,
the metal parts are mechanical and magnetically independent.

c) At the outer diameter of the winding set assembly:

 Stray field penetrating on the tie rods of the press frame between two phases (where
applicable).

2.4.3 Operating conditions

Temperature rises in the core and other metallic structural parts are caused by the combination of the
voltage related losses (no-load) and the current related losses (stray field). The following operating
conditions influence the localised flux density and related local losses:

 the power factor of the transformer load;

 step-up or step-down operation;

 the tap position of the on-load tap changer;

 normal or over-excitation condition caused by over-voltage, frequency swings of faults on the


network.

All of these parameters make an analysis complicated, but based on the loading requirements and the
design of the active part one is usually able to determine the most severe operating conditions. The
calculation can be limited to these cases. By performing a back-to-back temperature rise test, it is possible
to check the effect of the combination of voltage and current loading [4].

The impact of a zero sequence current, as in the case of geomagnetic induced current (GIC), where DC
current is injected into the transformer could result in a partial to full saturation of the core. A GIC event
is more relevant for transformers connected to long overhead lines in certain areas over the world. High
temperatures in the core and core support depend on the core geometry, the clamping structure and the
material used [5].

2.4.4 Calculation of loss distribution

The aim is to calculate the temperature rise (hottest spot) of the structural metal parts of the transformer.
As a first step, the loss distribution in these metal parts should be calculated. The loss distribution is a
scalar function P(r,φ,z) [W/m3] over the volume of these metal parts. As a second step the temperature
rise has to be calculated, based on this loss distribution taken into account thermal conductivity of the
metal, the thermal conductivity of the electrical insulation (if applicable) and convective cooling due to oil
flow. The cooling can be considered as a vector function C(r,φ,z) [K/(W/m2)] over the total transformer
volume, because heat transfer is in three directions [2].

To model a transformer for the calculation of the losses, the following statements on electromagnetic
properties of metals have to be considered.

18
Chapter 2: Temperature rises in steady state

 Non linearity and saturation of steel and core steel, so no superposition possible.

 Large anisotropy of core steel because “high grain oriented” material has different magnetic
properties in rolled direction compared to non-rolled direction.

 Skin depth is small (tens of millimetres for copper, aluminium and stainless steel but parts of
millimeters for construction steel and core steel).

 Large uncertainty, because steel (both normal construction steel and stainless steel) are
mechanically well defined but magnetically they are not so well defined.

From an electromagnetic point of view the following items regarding the transformer construction have to
be considered.

 Large dimensions compared to skin depth.

 Circular windings (which does not imply rotational symmetry [2]) in addition to long structural
metal parts along all other phases, such as yoke beams.

 Flux has to close around a path and this requires a proper choice on the location of the boundary
of the model and the boundary conditions.

 Magnetic field is determined by a vector potential, so 3D models are almost inevitable.

 The core consists of 0.3 mm thick sheets, which are electrically insulated from each other. A very
fine mesh is required in the perpendicular direction of the core sheets.

The following summarized the conclusions on modelling to determine the localised losses.

 Material properties are strong non linear, anisotropic and element size for modelling must be
much smaller as the skin depth. For accurate results, most literature emphasise that at least two
mesh elements are necessary inside the skin depth [6].

 Virtually the complete construction needs to be modelled.

 Some parts require a large number of very small mesh elements, which will result in a very large
number of elements.

 To determine the local losses, one needs a 3D model of the whole construction, which requires a
very large number of small elements in a non-linear AC calculation.

 Many spatial (multi-dimensional) simulation packages do not have the non-linearity properties of
the elements for eddy current problems. In most cases a transient simulation is considered to take
non-linearity into account, rather than an eddy current solution which uses an AC harmonic
solver.

 Applying the correct meshing techniques to the model plays a vital role in achieving an accurate
result.

19
Chapter 2: Temperature rises in steady state

2.4.5 Calculation of temperature distribution

To model a transformer to calculate the temperature, the following statements regarding the thermal
properties of all materials have to be considered.

 Thermal conductivity for different materials is usually well defined. Reasonable approximations
are possible for the core material.

 Viscosity of oil is important, which makes the temperature also an important parameter.

From an oil flow point of view in the case of ON, the thermo-siphon flow also has to be modelled to
include the various effects of the coolers. In the case of the windings, the thermo-siphon flow can be
simulated by assuming a very low oil velocity at the inlet of the cooling ducts in the windings. This
velocity is deduced from the temperature gradient over the height of the windings.

From a heat transfer point of view, the following statements regarding the transformer construction have
to be considered.

 Free oil-convection in “open” space, which is not well defined.

 Heat transfer takes place in the boundary layer, which requires small elements in the CFD
modelling and so large computer capacity both from memory as processor side.

 Defining the most optimal element or meshing technique is vital to avoid numerical instability and
inaccuracies in the result.

 The tolerances regarding the metal structure parts are large and usually inconsistent with respect
to the relevant dimensions for the oil velocities.

In conclusion, the cooling surface conditions (oil velocities, surrounding oil temperature, parallel high
current leads and electrostatic shielding) are not so well defined and the results of the calculations will
have a large tolerance.

2.4.6 The practical approach

As discussed in Sections 2.4.4 and 2.4.5, it is possible to determine the loss distribution and temperature
distribution but this approach could potentially be a computational time consuming one. Based on the
temperature distribution results, it is then possible to determine the location of the hotspot and the hotspot
temperature rise and that should then be validated by measurements. A well defined handshake with
simulation software is made by actual direct measurements.

If one takes a step back, one is in essence only interested in the hotspot temperature rise in the structural
metal parts and that location can be well predicted by testing. By performing measurements on these
spots, one can determine the relations between magnetic field and the relevant parameters.

For example:

 The location of the maximum stray flux density on the tie plate (and smallest core plates) is
known, so the location of the maximum losses is known.

20
Chapter 2: Temperature rises in steady state

 The yoke beam is in the stray field and the location of the maximum flux density on the surface of
the beam is known.

 In the case the yoke beam works as a magnetic shunt between phases, one other possible location
for the hotspot will be between the phases due to magnetic saturation of the steel.

 The magnetic field generated due to the stray field of one or two phases on the tie rod can be
calculated and the area of measure be established.

To limit the temperature rise in the structural metal parts, the technical solutions can be simple at
relatively low cost [7], [8], [9]. These temperatures can be influenced by:

 difference between winding heights;

 moving the metal parts away from the flux concentration, like increasing the core to winding
height;

 factory tolerances;

 splitting of tie plates;

 splitting of core plates;

 shielding (by aluminium or copper) or shunting (by magnetic material);

 choice of construction material (steel or stainless steel).

In the design stage the above mitigation techniques are used to lower the temperature rise in the structural
metal parts during the life of the transformer. In the case that the required solutions are not implemented
properly, it will result in unacceptably high temperatures (< 150-250 ºC). Temperatures in this range
might not be detected during the normal temperature rise test at nominal current, but they reduce the
lifetime of the transformer considerably due to paper and oil degradation.

An alternative solution to address the challenges of flux penetration and temperature rises is the use of
(non-magnetic) stainless steel. Stainless steel has a lower yield strength than steel and that might result in
larger dimensions, but in many cases, if applied correctly, can offer a successful alternative. The
compromise is that the thermal conductivity of stainless steel is approximately a quarter that of steel.

Measurement of temperatures in structural metal parts is far less complex than in the windings. All
structural metal parts of the transformer are grounded and direct temperature measurement is feasible with
thermocouples or fibre optic sensors. The thermocouples and leads can be fixed at locations where there
is almost no electrical field stress (see Chapter 5 on direct measurements). Measurements give very
valuable data, but one has to realize that one has to determine the temperature difference between the
object and the surrounding oil. If this is not the case, this might result in large measurement errors [2].

2.4.7 Conclusion

To simulate the loss distribution in the structural metal parts of a transformer requires at first a non-linear
magnetic AC calculation model with a very large number of small mesh elements. The validity of the
result will be highly dependent on the mesh size and placement to ensure numeric stability. An additional
21
Chapter 2: Temperature rises in steady state

consideration is whether the problem is a 2D problem or a spatial multidimensional one, which requires a
3D simulation.

To calculate the temperature rise distribution in the structural metal parts, this loss distribution serves as
an input to a CFD calculation, requiring a much higher resolution mesh density to ensure result accuracy.

The location of the hotspots in the metallic parts can be estimated in an FEA environment and in addition
measurements can be made to complement the results obtained from these estimates.

It should be understood that simulation models deal with exact dimensions and well defined parts. In
reality the tolerances on the geometry can have a significant influence and it is up to the engineer to adjust
the simulation results to the actual construction.

It is up to each manufacturer to generate adequate models and validate the simulations with testing, to
ensure compliance with customer specification or standards.

2.5 Thermal-Hydraulic Network Model (THN)

2.5.1 Basics of thermal-hydraulic model

2.5.1.1 General

Detailed THN modelling relies upon basic conservation principles: a) conservation of heat, b)
conservation of mass and c) momentum conservation. Therefore, it implies a spatial subdivision of the
domain of interest into a set of smaller elements, wherein each one the conservation principles are
iteratively applied and must be finally observed (as a condition of convergence).

Unlike CFD modelling, which is a distributed parameter approach where the same conservation principles
are described as complex system of partial differential equations, THN describes them by simple algebraic
equation sets, and the consequential time-to-solution is found to be considerably shorter.

These principles may be applied to model the temperature distribution and oil flow in several parts of the
transformers (windings, core, coolers).

The detailed global THN model of core-type windings comprises two interdependent models: the
hydraulic network model; and the thermal network model:

1. The hydraulic model describes the oil flow distribution through a network of ducts and junctions.
The flow through the network is modelled using the analogy with an equivalent electric circuit
where the flowrate through the ducts and the pressure drop correspond to the electrical current
across the resistance branches and to the voltage between nodes, respectively.

2. The heating of the oil throughout the coils due to energy dissipation on the discs is solved by
coupling the thermal model (that includes all the heat transfer mechanisms) with the hydraulic
model.

Published THN approaches differ on how much detail they employ in the subdivision of the spatial
domain of interest as well as how they model the transfer of heat and pressure from element to element of
the global network.

22
Chapter 2: Temperature rises in steady state

Early models were simplified to reduced complexity necessitated by restricted computational resources,
the manner of the simplification often based on the purpose of the model. It is worth noting some of the
simplifications that were made in developing these numerical techniques without describing a complete
overview. In the Oliver approach [10] the axis-symmetrical cylindrical geometry of the power
transformer is approximated to rectangular ducts with constant cross-sectional area while in Declercq et al.
[75] the temperature of each surface of the conductor disc is assumed to be constant. All of these models
additionally consider that oil streamlines are parallel and the flow is fully developed. However recent
CFD observations [131], [132], reported the existence of low-frequency eddies due to sharp angles,
particularly near the entrance of the radial ducts.

More sophisticated models take into consideration the variation of temperature along the radial and axial
directions and give importance to minor losses that actually have dominant influence on the flow
distribution and pressure loss. These methods use empirical correlations for the friction factor and heat
transfer coefficients relying on isolated experimental measurements or indirect measurements to validate
the simulation results. Other authors include correlations extracted directly from CFD simulations
showing how these two approaches can be combined [133].

Ideally, the analysis should be done for the whole transformer, including the radiators. However, the
complexity of the problem and size of the computational domain can be reduced using heat-run test data.
In fact, using the bottom-oil temperature rise, winding average temperature rise and losses, it is possible to
estimate the total oil flow rate in the winding. This value is then used as an initial condition in a detailed
THN model of the winding and the oil flow rate is iterated until an average winding temperature matching
the heat-run test result is obtained. Once the exact value of the flow rate is thus determined, the hotspot
temperature can be readily calculated.

2.5.1.2 Presentation with hydraulic networks

The basic equations of pressure drop equilibrium are given by Bernoulli. They can be set for each of the
independent oil loops. Such an approach, where Bernoulli equations are set and afterwards solved, is used
in [10] and in [44]. The alternative is to perform an intermediate step: first to ‘visualise’ the mathematical
model of oil flow, i.e. to make the hydraulic network, and afterwards to solve such a non-linear circuit.
Such visualisation contributes to better understanding of the physics and the ratio of characteristic values.
Such a concept also helps in forming the hydraulic concept of the complete transformer: closing oil loops
consisting of an inner active part and an outer cooling part.

Hydraulic networks may be illustrated using the example of a disc winding with oil barriers (zig-zag oil
flow); a winding part is shown in Figure 2.4. The hydraulic network shown in Figure 2.5 consists of 8
nodes (4 nodes on each of the entrance and exit sides), with one node on each side corresponding to the
point where the oil enters or exits the network while the remaining 3 nodes of each set correspond to
branching or merging. The physical quantity associated with the nodes is the sum of static and dynamic
pressures. Friction-related pressure drops are described by frictional hydraulic resistances (in axial
cooling channels on left side – Raf1, Raf3, Raf5 and Raf7, axial cooling channels on right side – Raf2, Raf4, Raf6
and Raf8 and radial cooling channels – Rf1, Rf2, Rf3 and Rf4). Local pressure drops are described by
hydraulic resistances as either branching (RC,S’ and RC,St’), merging (RC,S” and RC,St”) or corner (RC)
resistances. Gravitational pressures are described by “pressure generators” with downward pressure. The
quantities “flowing” through the branches are the oil flows. Thus the analogy with with electric circuits
may be seen by equating the sum of static and dynamic pressures to a potential; the hydraulic resistances
to electrical resistance;, “pressure generators” (for gravitational pressure) to voltage generators; and oil
flow to current. It should be noted that the local and frictional pressure drops are calculated for the oil

23
Chapter 2: Temperature rises in steady state

flows and re-calculated for local oil temperatures. The hydraulic network is non-linear since the hydraulic
resistances depend on the flow and has to be solved by some numerical iterative procedure.

Figure 2.4: Illustration of disc winding with barriers.

in1 Temperature of oil entering the first pass (°C)


Out1 Temperature of oil exiting the first pass (°C)
in2 Temperature of oil entering the second pass (°C)
Out2 Temperature of oil exiting the second pass (°C)

The elements of the hydraulic network depend on the oil temperatures, which depend on oil mass flows
and heat transferred from the solid elements from which it is generated (windings, core, and construction
parts – due to stray flux). The influence of oil flow on heat transfer from the solid elements to the oil
should also be taken into account; this affects the oil temperature and consequently the hydraulic
resistances and gravitational pressures. So, hydraulic networks cannot be solved independently from
networks describing heat transfer from solid elements to the oil, these are explained below. For certain
elements of the transformer it is possible to introduce some approximations (causing negligible errors) and
to decouple hydraulic and thermal calculations for example a disc winding with barriers [51], but for other
elements this is not the case, such as a layer winding with axial cooling channels [51]. For the cases when
the decoupling is not possible the computational procedure becomes more complicated and more time
consuming.

24
Chapter 2: Temperature rises in steady state

Figure 2.5: Hydraulic network of one pass of disc winding with barriers.

Solving the hydraulic network of the elements of the transformer gives the distribution of oil flow and
total pressure difference between the nodes of entrance and exit. This can only be achieved if the flow of
oil into the element and the oil temperature are known. The elements are thus “prepared” for construction
of the complex network of the transformer.

The oil flow through each of the elements of a transformer is determined by an iterative procedure of
solving the complete complex network of the transformer. Within the network, the total pressure drop
across each element is calculated and this depends on the flow through it. Applying the iterative
procedure, the flows should be distributed in a way to achieve pressure equilibria in each of the
independent oil loops.

2.5.2 Understanding the global pressure equilibrium in transformer

There is an alternative approach to forming a network model of a transformer than that described in the
section above. Arguably it offers a better understanding of the physics involved and is described below.

Oil circulation results from a balance of thermal forces, pump driving forces and pressure losses in the
flow network. The thermal driving force appears as the consequence of change of oil density [3].
Pressure drop results from frictional and other losses due to oil flow throughout the elements of the oil
25
Chapter 2: Temperature rises in steady state

circulation path. Both the driving pressure and the pressure drop depend on oil flow. In the steady state,
the oil flow, Q, represents an equilibrium of driving pressure and total pressure drop.

Figure 2.6 shows a simple single oil loop. Oil is heated in the winding (AB), it then flows through the
space above the winding (BC) before being cooled in the radiator (CD) and finally flows through the
space between the exit from the radiators and the bottom of the winding (DA) to start the loop again. In
the regions BC and DA, heat exchange exists (for example cooling on connecting tubes) but is negligible
comparing to heat exchanged in windings and radiators; consequently any change of oil temperature is
also negligible in these regions.

Figure 2.6: Change of oil temperature along the simplest oil loop.

a ambient temperature (0C)


ob bottom oil temperature (0C)
ot top oil temperature (0C)
oa average oil temperature (0C)

The thermal driving force (pT) is equal to the integral of the weight per unit volume along the oil loop, and
may be expressed as:

pT    g d l    g cos dl (2.7)

where  is the oil density [kg/m3], g is the gravity vector (9,81 m/s2),  is the angle between velocity and
gravity vector and l is the path vector [m].

Figure 2.7 shows change of oil density along the oil loop. For simplified representation, convenient for
basic understanding, expression (2.7) can be simplified [46] to the form:

pT  r g  Ol H (2.8)

where r is the oil density at a reference temperature [kg/m3],  is the volume expansion coefficient of
the oil [1/ºC], Ol is the vertical temperature gradient (ot - ob) [K] and H is the height difference
[m] between the vertical center of the radiators and that of the winding (at this point certain refinement is
possible [3]).

26
Chapter 2: Temperature rises in steady state

Figure 2.7: Change of density of oil along the simplest oil loop.

When the pumps run it directs the oil to the windings (OD cooling) and the total produced pressure (pprod)
is equal to the sum of pT and pressure produced by the pump (pP), where pP is much larger then pT:

p prod  pT  p P (2.9)

The pressure drop in the oil loops is equal to the sum of individual pressure drops in elements through
which the oil passes. There are two types of pressure drops:

 dispersed pressure drop – flow through straight tubes of constant cross-sections);

 local pressure drop – elements where oil changes oil streaming lines).

As has already been stated, pressure drop depends on oil flow; oil flow also influences the value of
produced pressure, both thermal driving force and pressure produced by the pump. From equations (2.7)
and (2.8) and Figure 2.6/Figure 2.7 it can be seen that thermal driving force depends on change of
temperature along the oil circulation loop. The energy balance equation for the winding is given by:

P   cP Q Ot Ob  (2.10)

where P is the power losses in winding [W],  is the oil density at ob [kg/m3], cP is the specific heat of
the oil [J/(kg K)] evaluated at the mean oil temperature in the winding and Q is the oil volume flow [m3/s].

From equation (2.10) it is clear that increased oil flow leads to a reduction of the temperature gradient
(ot – ob) and consequential decrease of thermal driving force (equation (2.8)).

In order to calculate produced pressure using the simplified equation (2.8) the temperature gradient (ot –
ob) is needed. The two equations for pressure equilibrium (2.9) and energy balance (2.10) contain three
unknowns: flow (Q) and the two oil temperatures (ot and ob). A third equation describes the cooling in
the radiator. The differential equation of energy balance (Figure 2.8) of oil in the radiator is:

k P O O ( x)  a  dx    cP QO dO ( x) (2.11)

where kP is the total heat transfer coefficient (HTC) [W/(m2 K)] defined as the reciprocal value of the
thermal resistance from one fluid to another attaching a wall between them, O is the circumference of the
27
Chapter 2: Temperature rises in steady state

outer radiator cross-section [m]; for a circular tube O =  D, where D is outer diameter of tube, O(x) is
the oil temperature at position x [ºC] and QO is the oil flow through the radiator [m3/h].

The previous equation is similar to the well-known basic equation of heat exchangers [50], with the
difference that the temperature of one of the fluids (in this case the air) is constant.

Figure 2.8: Energy balance of oil in the radiator

For a one-dimensional wall (a commonly used approach in the theory of heat exchangers, meaning that the
high thermal resistance through the tube in direction of fluid flow causes low heat transfer through the
tube in this direction), kP can be taken as constant equal to

1
kP 
1  Ri 1 (2.12)
 
O R i  a

where O is the convection HTC from the oil to the radiator [W/(m2 K)],  (Ri/Ri) is the sum of ratios
thickness to thermal conductivity of radiator paintings layers and of iron wall [W/(m2 K)] and a is the
convection HTC from the radiator to the air [W/(m2 K)].

Constant values of O and a are taken as average convection HTC over the surface of the radiator. This
is an acceptable approximation since in this case convection HTC does not change over the radiator
significantly – the entry region is small in comparison to the total length. Accepting this assumption, the
solution of the equation (2.11) is

kp O
 x
O ( x) a  Ot  a  e (2.13)
 cp Q

By integrating cooling power along the radiator,

LR

P   (O ( x)  a ) k P O dx (2.14)
0

28
Chapter 2: Temperature rises in steady state

we get the following:

 k O LR
 P 
P   c P Q Ot   a 1  e  cP Q  (2.15)
 
 

If in this simple illustration we assume that there is no cooling on the tank surfaces and no core losses, the
power of heat transferred through the radiator is equal to power losses in the windings (P = P).
Consequently, from equations (2.15) and (2.10) and condition (2.9), two unknown temperatures and the
oil flow can be determined. The thermal driving force can be calculated using either the simplified
equation (2.8) or the exact equation (2.7), where the real shape of temperature, defined by (2.13), is used.

The next example, contributing to a better understanding of the physics is the example of two windings
and radiators, shown in Figure 2.9. The following is assumed:

 only one phase exists,

 there are no core losses,

 no oil flow through the core,

 no cooling at the tank surfaces,

 no by-pass of the oil (between the windings and the tank).

In the case from Figure 2.6, the temperature of the top oil has the value by which the cooling power on the
radiators is equal to the power losses in the winding. For the case from Figure 2.9 the oil mixes and the
corresponding mathematical model becomes more complicated.

Initially the oil flows through winding 1 (Q1) and winding 2 (Q2) are assumed; but they will actually be
calculated later as the result of equilibrium of produced pressure and total pressure drop in two oil loops
(one is formed by oil flowing through winding 1 and the radiators and another through winding 2 and the
radiators). From energy balance for both windings and for the radiators the following equations may be
written:

P 1   c P Q1 Otw1 Ob  (2.16)

P 2   c P Q2 Otw 2 Ob  (2.17)

P 1  P 2   c P Q1 Q2 Otr Ob  (2.18)

where  is the oil density at ob [kg/m3], cP is the specific heat of the oil [J/(kg K)].

29
Chapter 2: Temperature rises in steady state

Figure 2.9: Simplified example of oil flow in transformer with two windings.

ob Bottom oil temperature (0C)


otw1 Oil temperature at the top of winding 1 (0C)
otw2 Oil temperature at the top of winding 2 (0C)
otr Oil temperature at the top of radiators (0C)
P1 Power losses in winding 1 (W)
P2 Power losses in winding 2 (W)
Q1 Oil flow through winding 1 (m3/s)
Q2 Oil flow through winding 2 (m3/s)
Qr Oil flow through radiators (m3/s)
hs1 Hotspot temperature rise of winding 1 (K)
hs2 Hotspot temperature rise of winding 2 (K)
av1 Average temperature rise of winding 1 (K)
av2 Average temperature rise of winding 2 (K)

Oil volume flows are related to bottom oil temperature (mass flow is constant) whereas specific heat is
taken at average oil temperature (ob + ot) / 2; for ot, in (2.16) otw1 is taken, in (2.17) otw2 and in (2.18)
otr.

As in the case of the simplest single oil loop, the equation describing cooling on the radiator is needed. It
is the same as (2.15), where sum of oil flows Q1 and Q2 replaces Q, sum of P1 and P2 replaces P and otr
replaces ot. The procedure for calculating four unknown temperatures and two flows is iterative: with
supposed oil flows first the temperatures are calculated, afterwards oil pressure components, followed by
check of pressure equilibrium. If pressure drop equilibrium in one or both oil loops is not fulfilled, the oil
flows are changed and the calculation procedure of oil temperatures and pressures repeated.

So, the final results of the described procedure are oil flows through the windings and the radiators, as
well as oil temperatures at the inlet and outlet of these parts. Oil flow and inlet oil temperature are
important input values for the calculation of distribution of temperatures in the windings.

At this point it is possible to split the equation for pressure equilibrium starting from the previously
described procedure of equalizing produced pressure and total pressure drop in complete oil loops. The
integral describing thermal driving force (2.7) may be split into two integrals – one corresponding to oil

30
Chapter 2: Temperature rises in steady state

flow through the tank and one to oil flow through the radiators (if the pump exists, the pressure produced
by the pump (pPump) is added to this second integral); for the loop related to winding 1 (Figure 2.9) it
yields the following:


p prod  
 1

DAB C

 CD

 g cos ds     g cos ds  p Pump  (2.19)

Pressure drops (pD) can be grouped into two sums: the first is the sum of pressure drops inside the tank,
including winding 1 – pDTank1 and the second in the radiators (including entrance and exit from them) –
pDRad. The pressure equilibrium equation:

p prod  pDTank1  pD Rad (2.20)

may be now written as:

 
   g d l  pPump   pD Rad  pDTank1    g d l (2.21)
 
 CD  DAB1C

For winding 2 the corresponding equation yields:

 
   g d l  p Pump   p D Rad  p D Tank 2    g d l (2.22)
 
 CD  DAB2C

This splitting is an important principle also used for forming hydraulic networks of parts of transformer.
From equations (2.21) and (2.22) it is clear that the sum of gravitational (terms of integrals of densities)
and frictional pressures in parallel branches between two points must be equal. Systematically making use
of this fact simplifies the structure and convergence of the calculations for individual transformer parts
and the whole transformer.

In fact, this observation provides the same conclusions as already expounded in Section 2.5.1.2.

2.5.3 Basics of thermal model of parts of solid materials

2.5.3.1 General

Section 2.5.1 deals with the determination of oil distribution: both globally (flows through transformer
parts) and inside parts of the transformer (throughout the cooling ducts in the windings, for example). The
oil distribution has an influence on the oil temperatures and convection heat transfer coefficients between
solid surfaces and the oil (from the active part to the oil and from the oil to the surface of the coolers
(radiators/compact coolers). Consequently, oil flow has an influence on the temperature of the solid
surfaces. This part of the complete model for the thermal calculation of an oil power transformer is
described in this Section. As the oil distribution depends on the distribution of temperatures in the solid
insulation and in some cases these calculations cannot be split, i.e. an iterative procedure for the
calculation of oil flows and the temperatures of solid surfaces has to be applied.

31
Chapter 2: Temperature rises in steady state

Basically, there are two approaches to the thermal calculations of solid parts of a transformer. The first is
applied to the radiators, as already illustrated by equations (2.11) – (2.15). The second is applied to those
parts of a transformer having different elements where the heat is generated. These elements exchange
heat through the solid insulation between them and with the surrounding oil. The heat exchange between
these elements and the oil consists of two components: thermal conduction through the insulation and heat
convection from the insulation surface to the oil. An approach that can be used for such parts of a
transformer is to build thermal networks. The briefly described principle will now be illustrated on one
disc of a disc winding with barriers.

The thermal network of one disc between radial cooling channels is shown in Figure 2.10. To simplify the
explanation, the case of one conductor in an axial direction between two radial cooling channels is shown
and will be considered. It is a reasonable simplification to assume that the temperature of the oil in all
axial cooling channels on the entry oil side is the same (in), likewise on the exit oil side (Out). It means
that heat transfer from the conductor to the oil in the axial cooling channels is not taken into account by
determining distribution of flow and temperature of oil. Such a simplification is acceptable since it does
not noticeably influence either elements for local pressure drops or friction in Figure 2.5 neither does it
affect the conductor temperature distribution. A great benefit of the simplifications that non-linear
equations for calculating oil flows through the channels and equations for calculating temperature of
conductors can be established and solved independently (first for flows and then for temperatures).

The model corresponding to the thermal network from Figure 2.10 takes into account the power loss in
each of the conductors, resistances to heat conduction through the insulation of the conductor, the change
of oil temperature along radial cooling ducts and the local convection heat transfer coefficient on the
surfaces of each of the conductors attaching oil (values in radial cooling duct change along the duct). As a
reasonable approach, each conductor is represented by one node, i.e. the difference of power losses and
resistance to heat conduction inside the conductor is neglected. The error is negligible even with CTC
conductors, where the enamel insulation between individual conductors of CTC exists. The thermal
resistance of enamel (thickness of enamel equal to the thickness of enamel on the individual conductor) is
added to the thermal resistance of insulation over all individual conductors.

32
Chapter 2: Temperature rises in steady state

Figure 2.10: Thermal network of one disc of disc winding with barriers.

A further example of thermal networks using a two-dimensional network is given in [51], for a case with
layer windings with axial cooling channels.

2.5.3.2 Discussion about thermal resistances

Heat transfer on each of the surfaces of the elements of transformer parts can be described by simple
expressions for thermal resistance (K/W), valid for one-dimensional heat transfer. This does not mean the
thermal models are one-dimensional (they are mainly two-dimensional); only that the resistance, for an
axial or radial surface, is taken as for one-dimensional heat transfer. Due to realistic dimensions (of the
thickness of insulation comparing to dimensions of copper of the conductor, for example), the error from
such a simplification is negligible. For the convection, the fundamental equation for the temperature
gradient between the surface (S) and the fluid (f) by transferring the heat q through surface S to the fluid
is

1
S   f  q (2.23)
S

33
Chapter 2: Temperature rises in steady state

where  is the convection HTC from surface to the fluid [W/(m2K)].

The following convection heat transfers are of interest: from the surfaces of active parts to the adjacent oil,
from the oil to surfaces of the cooling elements (radiators), from the cooling elements (radiators, compact
coolers) to the outer cooling fluid (air, water). In the literature there are equations for convection HTC for
a certain number of configurations appearing in oil transformers, such as: a) one side heated channels, b)
both side heated channels (based on known fluid velocity: for the oil inside transformer and for the air by
AF cooled radiators), c) both side heated channels (for the air by AN cooled radiators), d) vertical wall
(for tank surface to oil and to air side), e) horizontal wall (four different variants – fluid under / above the
wall, heat transferred to / from the fluid – used for tank and surfaces on the bottom and on the top of
windings and core), f) inclined channels (for radial cooling channels without guiding oil).

Similar to the pressure drop coefficient , equations for  have been published, but sometimes they do not
really match the situation in the transformer and they have to be improved or completely new equations
are to be established, based on experiments on the models, based on CFD computations, but also based on
results of heat run tests of transformers.

The error of not taking radiation into account, i.e. neglecting radiation heat transfer, is negligible. The
biggest relative impact it would have is for AN cooled radiators, but it is also very small due to the small
view factor between two facing plates forming a radiator. Only at the outer surface of the last plate of
radiator radiation heat transfer would be comparable with the convection heat transfer to the air, but the
effect to the cooling power of the complete radiator is small.

2.5.4 Compiling hydraulic networks of complete transformer

The complete hydraulic network consists of inner and outer parts. The inner part has three parallel
branches: with windings, with core and of oil by-pass (flow through free space between the active parts
and the tank). The outer part contains the elements outside the tank.

2.5.4.1 Hydraulic branches inside Tank / the branch with winding

In branches of oil flow through the windings, the components of pressure drop in the following elements
exist:

A. Oil distribution channel taking oil from the pipe coming from radiators and distributing oil to the
holes under the windings. Note that this pressure drop is averaged for the limbs since the
difference of pressure drop for different limbs is negligible compared with total pressure drop.

B. Oil flow splitting, where oil leaves the distribution channel and enters the hole under the winding.

C. Friction (dispersed pressure drop) in holes under one or more of the windings. These holes are
one of the elements used to adjust the oil flow distribution between various windings.

D. Oil channels, or orifices, for producing pressure drops which are added to the pressure drop in
holes (item C); these are also the elements for adjusting distribution of oil flow.

E. The same as C and D, but for single windings – for example when more windings are supplied
from one system of holes.

34
Chapter 2: Temperature rises in steady state

F. Insulation system under the windings (from the point of view of hydraulic resistance (pressure
drop) pressboard electrical insulation represents a complex system of oil channels).

G. Winding itself; in this zone there is oil temperature and oil density change.

H. Insulation system over the windings (similar to the system under the winding).

I. Space under the pressing rings, through which the oil from different windings exits to free space
in tank.

Note: in ON and OF cooled transformers, elements A.-E usually do not exist.

2.5.4.2 Hydraulic branches inside Tank / the branch with core

If the core is OD cooled, which is actually difficult to implement in practice, the components A. – D. of
the branch with windings remain the same.

There are parallel oil flows through the parts of the core. For example, with three limb cores there are
three equal oil flows through the limbs and two equal oil flows in the yokes. Oil temperature changes
while oil flows through cooling channels between sheets of the core. The total pressure between the point
at the bottom and the top of core for both oil paths, with flows Q1 and Q2, have to be same. This means
that flows Q1 and Q2 have the values at which the total pressure drops are equal.

2.5.4.3 Hydraulic elements outside tank

Compared with the size and complexity of the hydraulic network inside the tank, the hydraulic network of
elements outside the tank is small and simple. It consists of pipes, valves and radiators (or compact
coolers), which in a hydraulic sense represents relatively long tubes of small diameter. There are different
ways of connecting the radiators with the tank: in smaller transformers they are mounted directly on the
tank, sometimes with a goose neck; in other designs radiator batteries are connected with tubes (pipes) –
for cold oil and for hot oil – to the tank.

Again, a hydraulic network can be generated automatically from the following:

 configuration, number, diameter and lengths of piping for hot and cold oil;

 position and type of valves;

 number of bands;

 number of radiators,

 number of plates/tubes per radiator,

 type and length of the plates/tubes.

2.5.4.4 Components of oil flow

On Figure 2.11 the components of global oil flow are shown, for the example of an OF cooled transformer
and for the case when the oil is cooled on tank surface.

35
Chapter 2: Temperature rises in steady state

Figure 2.11: Components of global oil flow for OF cooled transformer.

Qrad Oil flow through the radiators (compact cooler) (m3/s)


Qtank Oil flow downward the tank (m3/s)
Qap i Oil flow through each of active parts (windings and core) (m3/s)
Qby-pass Oil flow in space between the windings and the tank (by-pass of oil) (m3/s)
Qtot Total oil flow (m3/s)
Qtot = Qrad + Qtank =  (Qap i) + Qby-pass
bo, r Temperature of oil exiting the radiators (0C)
bo, t Temperature of oil at the tank bottom (0C)
bo, mix Temperature of mixture of oils of temperatures bo, r and bo, t (0C)
ap, in Temperature of oil entering the active parts (0C)
ap, out i Temperature of oil exiting each of active parts (0C)
to, mix Temperature of mixed oils exiting active parts (0C)
to, r Temperature of oil entering the radiators (0C)

The details of the hydraulic network of a complete transformer and the algorithm for the calculation of
components of global oil flow can be found in [51]; OD construction is considered in [52].

Figure 2.12 shows the hydraulic network for the previously discussed OF cooling of a three-phase two-
winding transformer. The flow through the cooler is equal to the sum of the flows through the windings
multiplied by the number of phases and the flows through the core and the flow of oil by-pass.

36
Chapter 2: Temperature rises in steady state

Bottom
Entrance core – Exit Top
LV LV
to LV Bottom from winding –
winding winding
winding winding winding top core
QLV + + +

(per phase)
pdEnLV pgBC-BLV pdLV pgLV pdExLV pgTLV-TC
Bottom
Entrance core – Insulation Insulation Exit Top
to HV bottom below HV HV above from winding –
winding winding winding winding winding winding winding top core
QHV + + +

(per phase)
pdEnHV pgBC-BHV pdIBHV pdHV pgHV pdIAHV pdExHV pgTHV-TC
Entrance
to core Core Core
+

Qcore
pdEC pdC pgC
Entrance to By-pass on
by-pass core height
+

pdEOBP Qby-pass
pgOBP

Cold oil Hot oil


Piping Coolers Coolers Pumps
piping piping
+ + +

Qcooler
pdPO pgPCO pdCC pgCC PP pgPHO

Figure 2.12: Hydraulic network of OF configuration.

2.5.5 Final results of complete calculation based on detailed THN

The result of the described THN method is a very detailed distribution of oil flow (global and inside the
parts of transformer) and distribution of solid insulation and oil temperatures. Such a level of detail may
seem unnecessarily complicated but the non-uniform power losses and oil velocities throughout the
winding make it only way to attain accurately the critical (the highest) values of the temperatures, which
are needed in practice.

In addition to the quoted main results of the calculation, the method also delivers pressures in different
parts of transformer. This, combined with detailed distribution of oil flows, oil velocities in the oil ducts,
oil temperatures, convection heat transfer coefficients on each of the cooling surfaces, temperature of each
of the conductors, combined with calculated losses in each of the conductors, offers huge potential for
optimizing design in order to avoid high temperatures due to high local losses or bad local heat transfer.
The method also gives information about some poor global cooling parameters: for example low total oil
flow due to weak pumps or too much oil by-passing the windings.

During the heat run test, top oil, average winding and winding hotspot temperature rises are checked and
have to be lower than guaranteed values. The designer needs to be able to predict these values and it is
most easily done from obtained results of THN.

37
Chapter 2: Temperature rises in steady state

The hotspot temperature is equal to the maximum conductor temperature. The average winding
temperature is calculated through the DC resistance of the complete winding, being a sum of resistances of
each of the turns (turn “i”), of cross-section Si and diameter Di at a calculated temperature of the turn i.

Di
 Si
(235  i )
avh  i
 235 (2.24)
Di
i S
i

From THN it is clear that the oil temperatures at the top of different windings are not equal; they also
differ from the oil temperature at the top of radiators. Consequently, average oil temperatures in different
windings are also different. In a standard heat run test, the average winding temperature minus the
average oil in the radiators is considered as the temperature gradient winding to oil. So, not only the
conduction and convection heat transfer in the winding, but also oil flow distribution between the
windings (see Figure 2.11) are the components that influence the temperature gradient winding to oil,
which is measured in a standard heat run test. The hotspot factor obtained from THN and used to
determine hotspot temperature from the results of heat run test (2.25) - similar as (2.1).

hs to, mix


H
to, mix  bo, r (2.25)
av 
2
An illustration of the thermal calculation of a winding (with oil guiding elements) is given in [51].

2.5.6 Recommended concept of the software based on THN

The THN model may be implemented in software for use in thermal analysis and design. In that case the
following can be recommended for the software:

 It requires data about distribution of the losses (in each of the conductors of the windings and in
the parts of the core). It is therefore convenient to integrate or link it with the software for the
calculation of power losses. Averaging the losses over the winding is not the correct approach
since the hotspot temperature is the consequence of non-uniform losses (Q factor) and the non-
uniform cooling (S factor); averaging the losses would mean equalizing the Q factor with 1.

 It should hold information on real transformer designs, including all relevant details (illustrations
are given in Section 2.5.7).

 It should be integrated: include the calculation of inner heating and for the outer cooling.

 It should be general, i.e. cover all common cooling modes (OD, OF, ON, with compact coolers or
radiators, AF, AN, OF) [130].

 Since the thermal calculations are just a part of engineering design tools, this tool should be
integrated in a software environment for the design of complete transformer: this affects input
data and the presentation of the results. Good visualization certainly helps to identify critical

38
Chapter 2: Temperature rises in steady state

points and to improve the design (the aim is efficient cooling system, meaning design of a cooling
system to meet the required temperature rise limits while minimizing the overall cost).

2.5.7 Example of constructions which can be covered by software

2.5.7.1 General

As described in Section 2.5.4, a complete hydraulic network of a transformer consists of hydraulic


networks of individual parts of a transformer. The hydraulic networks of individual parts depend on their
constructions and can significantly differ. It is obvious that the software should cover all constructions
met in practice. The aim of this Section is to give a brief overview of constructions of interest being used
in each of the transformer parts.

2.5.7.2 The windings

Values for total pressure drop, oil velocity in cooling channels, distributions of conductors and oil
temperatures for different windings found in practice should be developed and included in the software.
Two dimensional representations of 9 such winding types are presented in Figure 2.13.

Note: There is a substantial difference between layer windings with axial cooling channels formed by
strips (type 1) and disc windings with axial cooling channels formed by clack bands (type 3) regarding
pressure drop and convection heat transfer coefficients. The reason being that with strips the oil flows
straight from bottom to top and with clack bands, the clacks represent randomly positioned disturbance at
each of the discs.

39
Chapter 2: Temperature rises in steady state

40
Chapter 2: Temperature rises in steady state

Figure 2.13: Winding types.

2.5.7.3 The radiators

There are two types of radiators: plate (Figure 2.14) and tubular (Figure 2.15). Pressure drops on the air
side are completely different for the two types, i.e. different forms of equations are used to calculate them.
The same is true for convection heat transfer coefficients on the air side. In addition, the arrangements of
the cooling fans for different types of radiators are not the same: with plate radiators, the fans blow
horizontally, vertically or horizontally and vertically; for tubular radiators, the standard is for the fans to
blow horizontally.

Figure 2.14: Plate radiators.

Figure 2.15: Two types of tube radiators.

There are different possibilities for positioning the fans blowing horizontally and vertically along the plate
radiators. Consequently, there are zones of the surface being cooled by natural air flow and by forced air
flow. In AN and AF cooled zones, convection heat transfer coefficients differ greatly and the software
needs to identify these zones for each of the plates. Convection heat transfer coefficients for AF-cooled
zones depend on air velocity. The air velocity is determined as the velocity by which there is an
equilibrium of the pressure produced by the fan and the pressure drop in the air stream flowing over the
radiators. This pressure drop depends strongly on the type of radiator and the position and the
arrangement of the fans.

2.5.7.4 The insulation system

There are two types of insulation pressboard elements. The first is between different windings: one or
more cylinders with oil channels between them. If there is more than one cylinder, they should be sealed
at the bottom so that oil does not flow between them. This means that oil between cylinders has
practically no influence on the cooling; more cylinders with blocked oil between can be considered as
adiabatic walls. If the sealing is not good, oil can flow between the cylinders and this can jeopardize the

41
Chapter 2: Temperature rises in steady state

cooling of an OD-cooled transformer (effective oil flow for the cooling of active parts reduces) and this is
one of the important quality issues in production.

The second type of pressboard elements appear between the windings and the yoke. This part, including
stress rings (potential rings), is designed by the electrical insulation designer and, for cooling
considerations, represents a labyrinth causing a pressure drop in the oil. The structure of these insulation
elements (example given on Figure 2.16) is the input data for the calculation of the pressure drop.

One result of solving the equations corresponding to an automatically generated hydraulic network is the
pressure drop in the insulation below or above the winding, while oil velocity in each of the channels is an
intermediate result; the oil does not change temperature along this labyrinth.

Figure 2.16: Example for insulation system under the winding.

2.5.7.5 The elements for adjusting oil flows through the winding

With no special consideration for cooling, the oil flow between the windings would be distributed
according to frictional and local pressure drops and according to the losses – these influence oil
temperature and consequently gravitational pressures. In such a case some of the windings could have too
little oil, which would result in high oil temperatures at the top of the winding and an excessively high
hotspot temperature. This is prevented through control of the quantity of oil through each of the windings
with the cooling design. Standard solutions are the application of the pressboard elements producing
additional pressure drop: the aperture, the ring causing zig-zag oil flow and opening of controlled diameter
and the length in support brackets (openings for injection of the oil from the oil distribution channel to the
42
Chapter 2: Temperature rises in steady state

windings). All of these solutions can be applied for a group of windings and/or for each of the windings.
If these elements are a part of the construction of a transformer, the possibility to consider their influence
via calculated pressure drop on them has to exist in the software. Example of such element is shown in
Figure 2.17.

Openings

Rings

Aperture

Oil distribution channel


Figure 2.17: Elements for controlling oil flow into the individual windings (typical OD construction).

2.5.8 Input data for the calculation

It should be clear that a very detailed model of oil flow and heat transfer is proposed above. It is based on
basic laws of physics, i.e. every characteristic component of the oil flow and heat transfer are
mathematically described. For the application of basic laws of physics, it is always necessary to specify
exact geometry and physical properties of materials. Consequently, these data are used in the detailed
THN model and have to be defined as inputs.

43
Chapter 2: Temperature rises in steady state

The required input data may be grouped as follows:

 Detailed data about the construction. Practically this means complete geometry of the transformer
elements, i.e. of the complete path of oil flow – channels, spacers (obstacles), elements where oil
changes the direction etc. An illustration, given for some data about the winding, is given in
Figure 2.18 and Figure 2.19. Also, the geometry data about solid insulation (paper, enamel, etc.)
should be specified (an example for the conductor is given in Figure 2.20).

Figure 2.18: The axial distribution of radial ducts.

Figure 2.19: The distribution of oil guiding elements inside the winding.

44
Chapter 2: Temperature rises in steady state

Figure 2.20: Example of a description of the conductor.

 It is important to note that some aspects of the construction (geometry) may be defined with high
accuracy and reliability (for example thickness of the tank or radiators and the paintings of their
steel walls) but for others (for example real thickness of the insulation and the width of cooling
channels after pressing) this is not straightforward. This can result from some characteristics
changing during manufacture; the changes depend on the applied technology of the manufacturer.
However, these few parameters can be identified (discussed in Section 2.5.9) and can be defined
based on empirical experience of the manufacturer.

 Since the model responds to physical properties of the outer cooling medium (via convection heat
transfer coefficients), the ambient temperature should be specified as an input.

 Specification of applied materials and their physical properties, for example: oil type, type of
insulation, type of radiator paint etc. The software should be organized to avoid direct
specification (meaning that the designer enters the data) of physical properties. Instead, the
program should contain certain sets of characteristics for materials which should be read when the
material is specified by the designer. Naturally, this set of characteristics should be easily
extended, i.e. adding new materials such as a new type of oil appearing on the market.

 Specification of the type of the equipment from external suppliers which would be built into
transformer (examples: pumps, fans, compact coolers). The best concept is to form the database
of potential suppliers and their products; similar to the previous point (physical properties of the
materials) characteristics of the equipment (for example dependence of oil flow of the pump on
produced pressure) should be stored in the database and imported to the program after selecting
the device.

45
Chapter 2: Temperature rises in steady state

 Finally, the distribution of losses is needed as an input. For the winding, the best way is to specify
additional losses in each of the conductors at a specified temperature. The losses due to the DC
component of resistance can be easily calculated. With these specified losses, the value of the
losses in each of the conductor at its final temperature can be calculated. This is the most accurate
definition for the losses to be used in a thermal calculation. There are different levels of
simplification: a) recalculating the losses to average winding temperature, b) taking the losses at
the temperature of the winding as equal to the guaranteed value (it does not differ much from the
actual value), c) taking equal losses in each of the conductors of one disc, d) taking constant
additional losses all over the winding. The inaccuracy caused by each simplification increases
from a) to d). Simplification d) can be is not suitable if a detailed THN model is applied and
simplification c) is not recommended.

2.5.9 Critical points and parameters in application of detailed THN

2.5.9.1 General

This Section deals with the difficulties in implementing a transformer design tool based on detailed THN
modelling.

2.5.9.2 The pressing factor influencing the duct width

Getting the correct dimensions for the radial cooling channels after pressing can be a challenge. The
reduction factor depends on the pressing of the conductor insulation and of the spacer. The manufacturer
will usually have values for the pressing factor, depending on the pressing force applied. The factors are
defined based on experience (winding height before and after the pressing) for transformers produced by
the manufacturer and it is assumed that they are reliable.

2.5.9.3 Bulging

Bending CTC conductors can cause the insulation around the cable to change shape and partially close the
radial channel, this is called bulging and has an even more unpredictable effect on getting the correct
dimensions than that mentioned above. There are two consequences that arise from bulging: 1) the cross-
section of oil flow is reduced – consequently the hydraulic frictional resistance is increased and 2) the
thermal resistance to conduction heat transfer through the insulation is increased by the presence of static
oil.

It is necessary to define the geometry of bulging in conductor insulation in order to define properly the
hydraulic and thermal resistances. There is an equation for estimating this effective increase of thickness
of the insulation but it does not offer sufficient accuracy required for detailed THN. The only way to
describe the bulging accurately is if the manufacturer carries out a systematic study and develops their
own equations. These equations may be formed by performing measurements of the real width of radial
ducts. The equation can be easily made by correlating the width of real radial ducts with the parameters of
the winding: thickness of the insulation, width of the conductor, height of the conductor, producer of the
conductor, winding diameter and spacer distance.

Bulging is a very important quality issue and it depends on the quality of the conductor. There have been
problems with some conductor producers resulting in large bulging and, in the extreme, reduction of the
radial cooling channels and an increase in conductor temperature. Significant bulging, if not taken into
account, can have a profound effect on the results of THN studies.

46
Chapter 2: Temperature rises in steady state

2.5.9.4 Local pressure drops


As already explained there are coefficients for local pressure drop  pressure drop   V 2 for which a
2
certain calculation procedure is needed (based on equations, graphics, tables etc.). Although such
procedures exist in the literature (for example [48]), as a rule they do not match well with the geometry
and properties of construction (size and fluid) in transformers. Typical basic elements of the complete
hydraulic network are oil branching, oil merging, corner, vertical T element from bottom, vertical T
element to top and oil crossing.

The only way to establish a procedure (if the published methods are not adequate) to determine accurate 
coefficients is through experimentation or numerical CFD calculation, which is now more popular as
technology makes it possible, but often verification through experimentation is a reasonable approach.

There are some research activities in this direction at universities and research centers, but to date there
are no published methods for reliably establishing  coefficients for essential elements.

2.5.9.5 The convection heat transfer coefficient in radial cooling channels without oil
guiding

There are equations for convection heat transfer coefficients for internal oil flow of known velocity, which
cover the majority of the cooling surfaces inside the windings. Results from hydraulic calculations are the
velocities in all axial cooling ducts and in radial cooling ducts if there is any kind of oil guiding (types 2,
4, 5 and 7 from Figure 2.13); but the oil velocity is not calculated in radial ducts in the winding of type 6
because the oil flow path is unknow as it is unguided.

To determine heat transfer from the winding to the oil in radial cooling ducts without oil guiding, the most
reasonable approach is CFD modelling (supported by verification experiment) and determination of
coefficients in the equations for the convection heat transfer coefficient.

2.5.9.6 The total heat transfer coefficient on the radiator surface

Experience shows that the rules and equations for the outer cooling used in practice are of high accuracy.
Typically, transformer manufacturers have accurate methods for the determination of the cooling power of
the radiators. From these rules, the average total heat transfer coefficient kP can be determined. In the
theoretical considerations related to equations (2.13) to (2.15) it is supposed that convection heat transfer
coefficients (HTC) from oil to the radiator wall and from radiator wall to the air and total heat transfer
coefficient kP do not change along the radiators; this approximation influences the shape of the profile of
oil temperature along the radiator and consequently to the shape of change oil density. A practical
consequence is the influence on the produced pressure in this part of oil loop – see equations (2.19) to
(2.22) - being a very important quantity for the determination of the oil flow, especially in ON mode. The
fine tuning of the method requires this issue to be properly treated; equations (2.15) should be improved
by considering a variable heat transfer coefficient along the radiator.

2.5.9.7 Space distribution of air flow

Figure 2.21 shows the surfaces of force cooled and naturally cooled plates for specified positions of the
fan. Air velocity for force cooled plates is obtained from the equilibrium of pressure produced by the fans
and pressure drop in the ducts existing between the plates. This rule can be applied in the following
approximate manner: produced pressure is obtained from the fan characteristic (dependence of effective
47
Chapter 2: Temperature rises in steady state

pressure on air flow) and pressure drop is calculated for the channel a x b: if the air flow between two
plates is Q1 (in the channel a x b) the pressure produced by the fan is determined for the value of air flow 4
x Q1.

The adopted approximation is equivalent to the situation that the fan blows uniformly through the channel
a x b. The air stream is conical with a circular base equal to the diameter of the fan (the distribution of air
flow over the cross-section of the fan circle, as the characteristic of the fan, is not uniform). Furthermore
some of the air does not flow over the plates then through the space between two radiators. These effects
can only be investigated using CFD software or experimentally.

AN b
AN
AF
AF
AF
AF
AN
AN

Figure 2.21: Position of the fan blowing vertically in respect to the plates of the radiators (view from bottom).

2.5.9.8 Radiators with vertically and horizontally blowing fans

The methodology of equalizing the pressure produced by the fans and pressure drop on the radiators can
be relatively easily applied if there are no big obstacles in the air path. This is the case when the fans are
blowing only in one direction: vertically or horizontally.

If there are fans blowing both vertically and horizontally, there will be a zone where the two air streams
clash. In this zone strong turbulence appears, causing higher pressure drops and increased convection heat
transfer coefficients. CFD software may be used to obtain high accuracy results for the pressure drop and
convection heat transfer coefficients (supported by a set of experiments). A similar situation exists when
there are fans blowing horizontally from opposite sides.

2.5.10 Illustration of use of THN as a calculation and analysis approach

An example is given for the transformer: three-phase five-limb, 360 MVA, 235 kV / 15 kV, YNd5, short-
circuit voltage 12.46 %, with constructive details specified in [52].

The total losses amount to: load losses in the HV and LV windings at a rated load of 935 kW and no-load
losses in the core of 210 kW.

The illustrative results of the calculations, for the rated load in normal operation, are given in this Section
for 3 cases:

 OF: LV winding with no zig-zag flow and OF cooling mode.


 OD: LV winding with no zig-zag flow and OD cooling mode.

48
Chapter 2: Temperature rises in steady state

 OD, zig-zag: LV winding with zig-zag flow (17 passes with 8 radial cooling channels and
one pass with 7 radial channels) and OD cooling mode.

Table 2.1 contains the calculated values of components of global oil flow distribution and Table 2.2 at
some characteristic oil temperatures.

Table 2.1: Oil Flows (m3/h).


HV windings LV windings Core Oil by-pass Compact cooler
OF 3 x 23.45 3 x 18.09 38.05 293.3 455.9
OD 3 x 57.27 3 x 36.40 174.5 - 455.6
OD, zig-zag 3 x 62.49 3 x 34.51 164.6 - 455.6

Table 2.2: Oil Temperatures  (°C).


bo, r out LV out HV out core to, r
OF 67.18 78.66 84.58 77.76 71.78
OD 67.41 72.21 76.25 69.77 72.17
OD, zig-zag 67.41 71.81 76.73 69.91 72.17

bo, r Temperature of the bottom oil exiting the radiator


out LV Temperature of the oil exiting the LV winding
out HV Temperature of the oil exiting the HV winding
out core Temperature of the oil exiting the core
to, r Temperature of the top oil entering the cooler

The temperatures (hotspot, being the critical value, and average winding, measured in a standard heat run
test) and characteristic values (winding to oil temperature gradient and hotspot factor) for LV and HV
windings are given in Table 2.3 and Table 2.4, respectively. The average winding temperature was
calculated from the value of DC resistance of the complete winding, as described in [51]. The location of
the hotspot was the same for all calculated cases: for the LV winding at the inner top conductor and for the
HV winding at the third conductor in the top disc.

Table 2.3: Characteristic Values for LV Winding.


Based on out LV Based on to, mix
hs LV av LV av LV–ao LV Hr LV av LV–ao, mix Hs LV
OF 103.8 91.01 18.18 1.385 21.62 1.483
OD 92.27 87.43 17.62 1.422 17.64 1.423
OD, zig-zag 82.16 75.16 5.56 1.864 5.38 1.859

ao LV Average oil temperature in the LV winding


ao, mix Average oil temperature in the tank (below and above active part)
Hr Real Hotspot factor of winding 1: (hs LV – out LV) / (av LV – ao LV)
Hs Hotspot factor based on mixed top oil: (hs LV – to, mix) / (av LV – ao, mix)

Table 2.4: Characteristic Values for HV Winding.

Based on out HV Based on to, mix


49
Chapter 2: Temperature rises in steady state

hs HV av HV av HV–ao HV Hr HV av HV–ao, mix Hs HV


OF 106.2 90.86 15.07 1.434 21.47 1.603
OD 97.13 88.51 16.68 1.252 18.72 1.334
OD, LV zig-zag 97.86 88.84 16.78 1.260 19.05 1.348

The meaning of the symbols is the same as for Table 2.3, but related to HV winding instead of LV
winding.

The most significant conclusions of this example are:

 Due to the pump being too strong there is huge oil by-pass in the OF cooling mode. The
consequences are high temperatures of the oil at the top of the windings (also at the top of the
core) and high hotspot and average winding temperatures

 Change from OF to OD construction, of supposed perfect sealing (no oil by-pass) leads to
reduction of quoted high values by OF cooling

 Change from OF to OD does not lead to significant reduction of temperature gradient winding to
oil. The reason is that the increased oil flow does not contribute to better cooling of radial
surfaces of the conductor: since there is no oil guiding, the cooling of radial surfaces remain
similar by OD cooling to the cooling by OF construction

 Introducing zig-zag oil flow decreases significantly the temperature gradient winding to oil:
hotspot and average winding temperature of zig-zag cooled LV windings are reduced

 As expected, the hotspot factor strongly depends on the construction: for the LV winding,
introducing zig-zag oil flow increased the hotspot factor from 1.42 to 1.86 (Q factor remained the
same)

More detailed description of the transformer and more comprehensive results are published in [52].

2.6 Computational Fluid Dynamics

2.6.1 Introduction

The advent of increasingly powerful computing resources has facilitated the approach of tackling
numerically (i.e. “discretizing”) a physical problem on a complex geometry by directly solving the
underlying governing equations without the necessity of drastically simplifying the model. The resulting
set of equations can then be solved on powerful computers based on numerical algorithms that can handle
the very large matrices arising from the discretization process.

Since the study of large-scale heat transfer problems on complex 2D and 3D geometries is of primary
importance in many technological and industrial applications, the mentioned developments have lead to
the establishment of the so called Computational Fluid Dynamics (often simply referred to as CFD).
Since many aspects are involved in defining, implementing, solving, presenting and analyzing a CFD
problem, this Section gives an overview of these aspects and their interrelation. Annex B presents
additional guidelines on CFD simulations.

50
Chapter 2: Temperature rises in steady state

2.6.2 CFD basic concepts and modelling approach

2.6.2.1 Modelling approach

The CFD approach is based on the solution of the governing equations (Navier-Stokes) which state the
conservation of mass, momentum and energy for a fluid flow. This set of partial differential equations
cannot be solved analytically, except in very special cases, and therefore to obtain a solution it is
necessary to use a discretization method which converts the set of differential equations to a system of
algebraic equations. The latter can then be solved on a computer and a solution at discrete points in space
can be obtained.

In general, a numerical solution is an ‘approximation’ of reality and the sources of error can be subdivided
in three categories:

1. Modelling errors

2. Discretization errors

3. Iteration errors

The modelling errors are associated with the accuracy of the mathematical model that is employed to
describe the flow phenomena. For example, in solving high Reynolds flows, turbulence models (k-ε, k-,
SST, etc.) are used to mimic the effect of turbulence on the main flow without actually solving the
turbulent motions. Another example of modelling error may be result from using a two-dimensional
model for analyzing a flow that is 3D. Consequently, the mathematical model is the starting point of a
CFD method and it is crucial to select the appropriate model for the desired application.

The second type of numerical errors relates to the discretization method that is chosen to convert the
differential equation to a set of algebraic equations. The three main discretization techniques are finite
element, finite volume and finite difference. The discretization errors are also due to the level of
refinement of the grid on which the algebraic equations are solved. The grid, or mesh, is a discrete
representation of the solution domain and it is made of a set of elements or control volumes. Generally, a
finer grid will yield a more accurate numerical solution. The generation of a mesh is often one of the most
challenging steps of a CFD method, particularly for complex geometries, and it is a key element of the
solution process.

The last source of numerical error is the iteration or convergence error owing to the use of an iterative
method for solving the discretized set of equations. In fact, when solving a numerical problem, a
convergence criterion is often specified for the residuals of the governing equations and this implies that
the exact solution of the algebraic equation is not formally attained.

Because of these potential sources of errors, CFD results should be validated by analytical solutions or
experimental data. Since in situ measurements are quite difficult and expensive to perform, scale models
are a good alternative for numerical validation. The main advantage is that the operating conditions may
be controlled with accuracy and the scale model can be equipped with more sensors in order to obtain a
more complete understanding of its thermal behaviour. Moreover, scale models are more flexible than
prototypes and allow testing of different cooling configurations.

In order to perform a CFD simulation three steps are generally required: the generation of the geometry
via a CAD tool (e.g. Catia, ProEngineer, SolidWorks etc.), the meshing of the computational domain and

51
Chapter 2: Temperature rises in steady state

the solution of the discretized set of governing equations. As illustrated in Figure 2.22, the creation of the
geometry is often the easier part of the process since with modern tools almost any body can be generated
in a 3D environment. Nevertheless, during this process it is critical to determine the level of geometrical
detail that is required for properly representing the problem to be solved. For example, in the case of a
power transformer, it may be decided to neglect the presence of the radial spacers or of the sticks since an
axisymmetric calculation is performed. Moreover, the location of the domain boundaries (i.e., inlets,
outlets, etc.) has to be chosen carefully since it may affect the accuracy of the numerical model.

Figure 2.22: Geometry of a low-voltage transformer winding generated with a CAD tool.

Once the geometry is built, the domain has to be subdivided or discretized into much smaller volumes
(cells) where the governing equations can then be solved. This process is generally referred to as “mesh
generation”. Various types of meshing techniques are available but they can be subdivided into two main
categories: structured and unstructured. The former type is the simplest grid structure and it is equivalent
to a Cartesian grid where the location of any grid point in the domain is uniquely identified by a set of
three indices (i, j and k) in three-dimensions. This type of mesh is often used for simple geometries such
as for the transformer winding domain illustrated in Figure 2.23. In general, structured mesh yields to
higher quality elements that can be oriented with the main flow direction. A disadvantage of structured
meshes is that a mesh refinement in a given region automatically results in an equivalent refinement in
other parts of the domain where a fine concentration of points is unnecessary.

52
Chapter 2: Temperature rises in steady state

Figure 2.23: Computational grid of fluid (oil ducts) and solid (discs) domains of a transformer winding.

In contrast, unstructured meshes are much more flexible and can be used for very complex geometries.
The elements may have any shape, although tetrahedral and hexahedral are the most common, and many
algorithms such as Octree, Delaunay, advancing front, etc. are available to create these types of meshes.
With unstructured grids, node locations and connectivity need to be specified and the matrix of algebraic
equations does not have a regular diagonal shape, which usually implies a slower solution time.

After generating the mesh, the system of algebraic equations needs to be solved using either direct or
iterative methods. The former approach is rarely used in CFD since it has a higher computational cost
particularly when the matrix is sparse. Examples of direct methods are Gauss elimination and LU
decomposition. Thus, iterative techniques such as Conjugate gradient, Gauss-Seidel and Multigrid are
more commonly used in the solution of CFD problems. With this approach, an initial solution is guessed
at and the algebraic equations are used to systematically improve it until convergence criteria are reached.
Consequently, the better the choice for the initial solution the faster the convergence rate will be and the
computational time may be reduced.

Since the number of grid points in the domain can reach 10s or 100s of millions, especially for 3D
problems, solver routines are parallelized and the problem is simultaneously run on multiple processors.
Nowadays, clusters equipped with thousands of CPUs are available and this has greatly helped to reduce
the computational time. Nevertheless, large problems handled with inadequate computational resources
may require weeks before a converged solution is obtained and this partly explains why simpler methods
such as THN are widely used in the industry particularly during the first steps of the design process.

One of the great advantages of CFD over other numerical techniques is that the values of the physical
quantities (i.e., pressure, velocity, temperature, etc.) that have been solved are known in fine detail

53
Chapter 2: Temperature rises in steady state

everywhere in the computational domain. This allows researchers and engineers to have a more complete
understanding of the fluid phenomena that are present. Figure 2.24 presents CFD results for an ON-cooled
winding section where the development of the thermal boundary layers along the disc walls can be
observed. Moreover, the dynamics of the hot streak is captured and its effect on the winding temperature
distribution may be determined.

Figure 2.24: Contours of temperature and velocity vectors in the cooling channels of an ON-cooled
transformer winding.

In conclusion, CFD is a numerical method that requires a higher effort from the user and a higher
computational time, but it gives access to detailed information on the flow that may not be obtained with
empirical or network techniques. Thus, CFD is a great tool to solve specific issues with a given design, to
validate simpler numerical methods or to feed the latter with appropriate coefficient values (i.e., local loss
coefficients, convection heat transfer coefficients, etc.).

2.6.3 CFD application in transformer thermal modelling

In a power transformer the windings cooling oil flows in a loop consisting of:

 the windings, where owing to winding losses (DC and eddy), the oil heats up;

 the coolers/radiators where the oil cools down,

 the tank in which the oil mainly acts as a heat carrier (although some tank cooling occurs as well,
it generally represents less than 10% of the transformers cooling capacity).

This implies that the cooling circuit consists of a combination of internal flow in narrow ducts (windings,
radiators/coolers) and open flow in the tank. One reason for this focus is that a strong reduction of detail
in analysing regions of less crucial interest is more difficult to accomplish with CFD modelling compared
to the application of TNMs. Consequently, the required efforts connected to a CFD analysis are mostly
54
Chapter 2: Temperature rises in steady state

concentrated on the transformer windings. In the thermal network models (THN) that are in use today
(explained in detail in Section 2.5), to predict temperatures and oil flows, the tank is therefore modelled in
a more simplistic way whereas the windings and radiators can be modelled in greater detail. Since most of
the losses occur in the windings and the winding hotspot temperature is known to be a decisive factor in
the ageing process of the transformer, a lot of efforts have been put into the detailed thermal modelling of
the windings, particularly for oil-guided disc type designs.

In principle, CFD modelling could be used to analyse all of the components of the transformer cooling
loop, but has been applied almost solely to the windings, often as an extension to the THN approach.
Since the main advantage of a CFD model is the possibility to resolve in a very detailed way the full set of
fluid flow and heat transfer equations, the primary applications have been the winding heat transfer
analysis and the validation of the THN models that are used for both design and analysis. Thus, the goal
is to improve (not to replace) THN methods, since the latter combine reasonable accuracy (on the
condition that the physical processes related with heat transfer and fluid flow are properly modelled) with
reduced demands on compulational resources, as required by the design process. Published CFD studies
on winding simulations [12], [108], show that CFD models have been used as an extension to the THN
approach with winding heat transfer analysis, THN winding model validation and winding hotspot
temperature determination as primary goals.

2.6.4 CFD in winding heat transfer analysis

In [108] it was shown that CFD winding models reproduce the typical oil velocity and temperature
profiles for the different cooling modes (ON: minimum velocity / maximum temperature at the top of each
pass, OF and OD: minimum velocity / maximum temperature somewhere in the middle or towards the top
of each pass) if, contrary to many published THNs, internal buoyancy effects are taken into account, i.e.
the convective heat transport in the windings is modelled using mixed convection instead of forced
convection. However, recent CFD studies [135]-[136] have shown that for OD cooling modes with high
flow velocities, a stagnant flow or even a reverse flow can be observed in the first horizontal duct of a
pass. This phenomenon is mainly due to the impingement of the oil flowing in the last duct of the
upstream pass on the oil flowing in the axial duct. Such interaction greatly affects the velocity profile in
the axial duct and prevents the oil from entering the first horizontal duct of the downstream pass.
Moreover, a low pressure region generated near the washer can suction part of the oil flowing in the
second horizontal duct and redirect it toward the first duct (instead of letting it flow upwards), thus
creating an inverse flow. This flow behaviour can cause a significant overheating of the bottom disc in the
pass which sometimes becomes the hottest disc in the winding. Other CFD analyses on a disc-type
transformer winding are described in [113], [114] and [118].

Figure 2.25 reveals that the winding-internal buoyancy differences play an important role in determining
the local pressure balance and have a large effect on the velocity and temperature distributions, including
hotspot temperature height and position in ON-cooled transformer windings. Including these upward
forces promoted by the fluid density variation with temperature (buoyancy forces) becomes increasingly
important as the overall magnitude of velocities decrease (typically for ON regimes). Furthermore, hot
streaks of oil (i.e. local temperature gradients in the oil due to the large Prandtl number of transformer
oils) and insulation of the strands/cables in the discs may have a significant effect on the cooling of
individual discs, as shown in Figure 2.26. Consequently, the location of the (local) hotspot does not
necessarily coincide with the lowest horizontal duct velocities at the same location.

55
Chapter 2: Temperature rises in steady state

Figure 2.25: The effect of including internal buoyancy in a CFD disc winding model [108] on velocity and
temperature distributions (identical oil mass flow and heat loss rates, ON cooling mode, [108]). The red curves
correspond with internal buoyancy included (mixed convection), the blue curve with internal buoyancy
excluded (forced convection).

Figure 2.26: Detailed temperature distribution in a section of a disc winding as a result of a CFD simulation.

The flow of the cooling fluid in the transformers windings may be laminar flow or it may be turbulent.
The question on whether turbulence occurs in transformer windings, as well its subsequent consequences
on the related heat transfer mechanisms and on the modelling approaches needed, has not been answered
yet.

At this moment, it is not clear whether turbulence occurs in practical designs with no official publications
available showing its relevancy.

What is known is that on one hand the viscous nature of the typical mineral oil generally limits the
Reynolds number below the critical value for turbulence to set in (the limit in the common literature for
56
Chapter 2: Temperature rises in steady state

ideal circular pipes is around 2100), and on the other hand due to the risk of streaming electrification the
average oil velocities are kept within certain bounds by design. These two known aspects together create
the conditions for a Reynolds number below the mentioned critical value, hence reinforcing a probable
laminar flow in most of the designs.

Notwithstanding this, it is deemed potentially possible that turbulence can occur for cases of OD cooled
windings with high oil flow rates and especially at high temperatures. In such cases the essentially three-
dimensional nature of the flow and its unsteady behavior, would require appropriate additional modelling
approaches, like described in [119]. Given the current state of knowledge, it can be said that the subject of
turbulent flow and turbulence related heat transfer in transformer windings is still an open area of
research.

Finally, the CFD approach can be used to study and compare the thermal performance of different
winding types [115], as for example shown in Figure 2.27.

Figure 2.27: CFD study of the thermal performance of different winding arrangements.

2.6.5 CFD and THN validation

A detailed modelling and measurement study [12] on a tested ON-cooled unit (thermal failure repair,
66 MVA, ONAN/ONAF, which was chosen in this working group as a reference unit) expands on the
work in [108] in many ways. The study investigates a comparison of the THN and CFD model results for
the 2D (uniform DC losses per disc) and 3D (non-uniform: DC + eddy losses) cases. For the 2D case

57
Chapter 2: Temperature rises in steady state

(Figure 2.28), the THN and CFD results generally agree and it is shown that the difference in calculated
temperatures can be explained by the effect of hot streaks, confirming the results in [108].

Figure 2.28: Flow and temperature distributions inside an ON-cooled winding [12] obtained with 2D THN and
2D CFD methods (uniform losses).

For the 3D case with non-uniform loss distribution, the differences between the 3D THN and 3D CFD
results are larger, both qualitatively and quantitatively, as shown in Figure 2.29. Now the effect of the
finite circumferential duct geometry due to the ribs and spacers is taken into account. As a consequence
of the 3D geometry, flow details like the hot streaks appear on a slightly different position, leading to
different temperature distribution.

Figure 2.29: Flow and temperature distributions in an ON-cooled winding [12] obtained with 3D THN and 3D
CFD methods (DC + eddy losses).

In the same study, direct winding temperature measurements using fiber-optic probes revealed that the
measured temperatures are of the same order as the hotspot temperatures in 3D models.

It should be noted that the LV windings of the test case as studied in WG A2.38 and in [12] has a very
large number of discs per pass between two oil guides, resulting in a relatively uneven, non-optimal flow
distribution in the horizontal ducts (as witnessed in Figure 2.29, particularly in passes 1 and 3 from the
left). This means that small changes in velocities (either because of the model applied, or owing to design
and manufacturing process induced variability) are supposed to have a large impact on the temperature
distribution. This may mean that in a design with a better oil guide distribution the differences between
the THN and the CFD results may appear to be smaller.

58
Chapter 2: Temperature rises in steady state

2.6.6 CFD and thermal model improvement

Since currently the application of direct full-scale CFD simulations on windings (or a complete
transformer) in the design process is too time consuming and THNs may show a good balance between
performance and accuracy, an important role of CFD may be to contribute to the improvement of the
THN-based thermal design models. This may be feasible by deriving more accurate coefficient
formulations as compared to known “textbook” formulations which depend on simplified geometries and
assumptions about temperature and flow distributions. Potential candidates for improvement are Nusselt
numbers and friction coefficients like derived in [127] and expressions for the pressure drop at junction
nodes (i.e. the areas where the vertical winding ducts connect with the horizontal ducts that separate the
winding discs) as published in [128].

2.6.7 Global winding pressure drop and heat transfer correlations

An alternative approach of using CFD in winding thermal calculations is to use a large number of CFD
simulations to derive correlations for use in the direct calculation of temperatures and oil velocities, in
case a THN is not used. In [109] such an approach is taken for the derivation of frictional pressure drop
and heat transfer correlation for use in the calculation of winding temperatures, as shown in Figure 2.30.

Figure 2.30: CFD-calculated correlations [109] between Reynolds number and friction loss (left) and Nusselt
number (right).

2.6.8 Radiator modelling

2.6.8.1 General

Fan-cooled radiators are used in the majority of power transformer installations and are complex
geometrical entities whose cooling properties can be modeled usefully using CFD. In [110] the buoyancy-
induced flow of air has been studied using a 3D model for free convection between parallel plates of finite
width. Depending on the Rayleigh number range, the flow and heat transfer are characterized by different
physical effects and thus describe different flow and heat transfer regimes.

CFD is particularly useful for optimization studies of radiator design. In [111] for example it is shown
how CFD is used to model the radiators, the fans and a large surrounding area, with Figure 2.31 below
showing the domain geometry, the air velocity, air temperature and oil temperature distributions
respectively for a specific transformer radiator. In the same publication is an optimization scenario based

59
Chapter 2: Temperature rises in steady state

on a number of CFD simulations, revealing the effect of fan size and fan coverage at the radiator air inlet
on cooling capacity. Used in this way, CFD supports the development and improvement of design
guidelines for cooling equipment.

Since the cooling efficiency of the coolers and radiators may be strongly affected by the surrounding area,
CFD may be the most appropriate approach to study the effect of enclosures, fire walls, sound screens and
similar objects on cooler efficiency.

2.6.8.2 Other components of the cooling loop

Apart from the windings and the cooling equipment where the bulk of the heat generation and removal
takes place and thus a coupled heat and mass transfer problem needs to be modelled, there may be other,
geometrically complex, parts in the cooling circuit that may profit from CFD analysis, in particular
regarding pressure drop calculation. In [112] an example of such an application (Figure 2.32) is used to
verify the pressure drop in so-called pressure elements (used for balancing the oil flow between
regulating, high- and low-voltage windings.)

Figure 2.31: CFD-based modelling results [111] of heat transfer and transport in and around two fan cooled
radiators. Upper left: impression of computational grid (horizontal crosscut at mid-height). Upper right: air
flow magnitude (m/s). Lower left: Absolute air temperature (K). Lower right: oil temperature (K) distribution
in radiator.

60
Chapter 2: Temperature rises in steady state

Figure 2.32: Geometry of a flow resistance element, and example velocity and pressure distributions [112].

2.6.8.3 Example of a detailed radiator modelling

Figure 2.33 illustrates the results of a detailed radiator modelling using CFD [125], [126]. For the
simulation, the inner oil and the surrounding ambient air are discretized. The constructed model reflects
the real surface shape of the radiator with an internal oil flow from top to bottom and a convective heat
transfer on the outer surface with the surrounding ambient air. All flows were specified in steady state and
as incompressible. These CFD results were also compared with experimental results as shown in Figure
2.34.

Figure 2.33: Detailed modelling of a radiator. Left: Computational domain. Center & Right: Radiator
temperature distribution at flow rates of 1.0x10-3 m3/s and 4.0x10-3 m3/s

61
Chapter 2: Temperature rises in steady state

Figure 2.34: Experiments carried out to assess the radiator cooling performance. Left: Test apparatus. Right:
Comparision of CFD simulation and experimental results.

2.6.9 CFD in tank wall computation

Tank heating due to stray losses is also another issue which may profit from the use of CFD. The
empirical analytical method might not be precise enough to predict the local temperature rise of tank walls
since the temperature rise of the tank walls depends on not only the loss density in the tank wall but also
the oil flow condition around them. The prediction of the local oil flow is quite complicated and may only
be well calculated using CFD-based methods.

2.6.10 CFD and the design process

As a consequence of the design and manufacturing process, small deviations from the design description
will inevitably occur (for example paper bulging, causing the effective oil duct diameter to be reduced
compared with the original design). This implies that extreme accuracy with a thermal modelling
approach like CFD will not guarantee small deviations during heat run testing. Instead, it is most
important to guarantee thermal robustness of the design with respect to small variations: here CFD can be
used to investigate the effect of these variations, particularly hotspot temperature (including those effects
that cannot be resolved by a THN or similar thermal design model, for example the consequences of hot
streaks in the oil ducts for hotspot temperature position and level). Based on these investigations, proper
design guidelines should be defined such that the thermal robustness of the design is guaranteed. In
addition to these guidelines, the CFD-based improvements of the thermal design models as discussed in
the previous sections may also be expected to improve the quality of the thermal design.

2.6.11 Conclusion

In the study of the thermal behavior of transformers, CFD may be usefully applied in various ways such as
the following.

 It provides for increased understanding of the physical processes behind heat and mass transport
and their effect on hotspot temperatures since it allows the governing equations of fluid flow and

62
Chapter 2: Temperature rises in steady state

heat transfer to be applied on a very fine numerical grid limited only by computational resources
and time.

 It can resolve local temperature gradients (like for example hot streaks in oil) and their
consequences for hotspot temperature position and strength. In this matter CFD exceeds the
capabilities of thermal network models and can be used to refine design guidelines to guarantee a
thermally robust design.

 It can be used to improve thermal models for analysis and design (both THN and non-THN)
through improved correlations formulations for pressure terms and convection heat transfer
coefficients.

 It can be used for design optimization of complex parts of the cooling circuit, fan-cooled radiators
for instance.

CFD is however not a replacement of THNs and in design because:

 It requires specialized resources, both material and human, for an appropriate implementation as
the quality of the simulation result depends significantly on the competence of the CFD tool user
on the physics behind heat and mass transfer.

 It requires considerably more computational resources than THNs. A CFD simulation requires a
calculation time that exceeds the THN calculation time by several orders of magnitude.

63
Chapter 3: Benchmark of numerical tools

Chapter 3: Benchmark of numerical tools

3.1 Introduction

This Chapter of the Brochure presents a comparison of modelling results obtained by the working group
members using their own tools and the characteristics of an existing transformer from a utility in Canada.
The simulations concern the calculation of the eddy losses and the temperature calculation. The loss
calculation results are used as an input for the thermal model. The temperature calculation is made with
uniform and non uniform losses to evaluate the impact on the thermal modelling.

It is worth mentioning that only the data of the following Sections 3.2 and 3.3 were shared with the
participants. The participants had to use their own methodology and there was no specification for the
meshing, 2D projection plane or any other boundary conditions. This approach was selected in order to
open the discussion on the general methodology for eddy losses and temperature evaluation instead of
comparing only the numerical tools.

3.2 Description of the transformer

The transformer under investigation is rated 40/53/66 MVA, ONAN/ONAF/ONAF, (225/26.4 kV, YNd1)
with an impedance of 22,72% (60 Hz). Figure 3.1 illustrates the overheating of the discs located at the top
of the LV winding that causes a turn-to-turn fault in the transformer. After the repair, this transformer was
equipped with fiber-optic probes in the windings to monitor the temperature during service. Figure 3.2
shows a schematic of the windings with the rated current flowing in the HV and LV windings.

Figure 3.1: Overheating of the conductors at the top of the LV winding of the transformer.

65
Chapter 3: Benchmark of numerical tools

H1, H2, H3 H0 X1, X2, X3


169.4 A
1443.3 A
833.3 A
84.7 A 169.4 A
833.3 A

169.4 A

84.7 A

Figure 3.2: Transformer windings schematic.

A detailed description of the winding is as follows:

 The LV winding comprises 78 discs and 232 turns.

 Each turn is made with six strands in parallel.

 The conductors measure 2.057 x 14.275 mm and are individually wrapped with 0.381 mm of
paper.

 There are 18 sticks with spacers and 18 sticks without spacers (see Figure 3.1); their width is 15.9
mm.

 The width of the cooling duct spacers is 28.6 mm.

 The disc-type winding design comprises five washers (0.99 mm), forming four passes of 19 discs
each, used to force the oil to circulate in the horizontal cooling channels between the discs (two
discs at the bottom are not included in these four passes).

Figure 3.3 illustrates the geometrical details of passes 1 and 3 (pass 4 being at the top).

66
Chapter 3: Benchmark of numerical tools

Figure 3.3: Geometrical details of the discs and cooling channels in one pass of the LV winding.

The heat-run test results indicated a winding temperature rise of 61.6°C. The ambient temperature, top oil
temperature and bottom oil temperature at the end of the heat run test (rated current, ONAF condition)
were respectively 30.2°C, 80.4°C and 46.7°C.

Table 3.1 presents other dimensions required for the eddy loss calculation.

67
Chapter 3: Benchmark of numerical tools

Table 3.1: Transformer dimensions.

Parameters Mm
Core window height 1690
Limb diameter 590
HV inner diameter 960
HV outer diameter 1122
LV inner diameter 650.2
LV outer diameter 751.7
HV/LV clearance bottom tank 760
HV/LV height 1499
Core clamp (z dim x r dim) 355 x 207
Upper clamp clearance from tank top 520
Lower clamp clearance from tank 160
bottom
Clamps clearance from core 12
Clearance HV winding to tank wall 295
Clearance top yoke – tank cover 400
Clearance bottom yoke – bottom tank 95

3.3 Modelling specifications

The following describes a list of simulation parameters specified to obtain a meaningful comparison
between the different tools used by the WG members. The first part concerns the calculation of the eddy
losses and the second part the temperature calculation. The result of the first part is used as an input for
the second part. The temperature calculation is made with uniform and non uniform losses to evaluate the
impact on the thermal modelling.

Ideally, the analysis should have been done for the whole cooling circuit of the transformer, including the
radiators. However, the complexity of the problem and size of the computational domain can be reduced
using heat-run test data. In fact, using the bottom-oil temperature rise, winding average temperature rise
and losses, it is possible to estimate the total oil flow rate in the winding. This value is then used as an
initial condition in a detailed thermal model of the winding and the oil flow rate is iterated until an average
winding temperature matching the heat-run test result is obtained. Once the exact value of the flow rate is
thus determined, the hotspot temperature and location may be readily calculated. This iteration process
was not easily applied for most of the teams performing the thermal calculations, so it was decided to
specify the inlet boundary conditions, i.e. the oil flow rate and the bottom oil temperature.

The detailed modelling specifications are:

1. Calculate the eddy losses at a temperature of 75°C. The average will be used for the temperature
calculation with non uniform losses.

68
Chapter 3: Benchmark of numerical tools

2. For temperature calculations, use material properties of mineral oil, paper and copper as follows:

Fluid properties (mineral oil):

k = 0.1509 [W·m-1·K-1] – 7.101*10-5 [W·m-1·K-2] * T

ρ = 1098.72 [kg·m-3] – 0.712 [kg·m-3·K-1] * T

μ = 0.08467 [Pa·s] – 0.0004 [Pa·s·K-1] * T + 5*10-7 [Pa·s·K-2] * T2

cp = 807.163 [J·kg-1·K-1] + 3.58 [J·kg-1·K-2] * T

Conductor (copper):

k = 401 [W·m-1·K-1]

ρ = 8933 [kg·m-3]

cp = 385 [J·kg-1·K-1]

Insulation (paper):

k = 0.19 [W·m-1·K-1]

ρ = 930 [kg·m-3]

3. Calculate the temperature using uniform losses

a. Apply 677 W per disc (no temperature compensation for the losses)

b. Inlet boundary conditions:

i. Oil mass flow rate of 0.78 kg/s

ii. Bottom oil temperature of 46.7°C (from heat-run test data)

4. Calculate the temperature using non-uniform losses

a. Set the eddy-loss set to the average of the CIGRE results

b. Temperature compensation of DC and eddy losses

c. Inlet boundary conditions:

i. Oil mass flow rate of 0.78 kg/s

ii. Bottom oil temperature of 46.7°C (from heat-run test data)

69
Chapter 3: Benchmark of numerical tools

3.4 Loss calculation

3.4.1 Results

A total of 10 teams (referred to here as A to J) made the eddy loss calculations. In addition to the
simulations made by the WG members using their own software, a simulation was made using a
commonly used package (Andersen). The numerical results are shown in Table 3.2, Table 3.3, Figure 3.4
and Figure 3.5. As expected, there is an increase in the eddy losses for the top discs due to the increased
contribution of the radial leakage flux at the ends of the windings. The variation in the absolute value of
the losses is discussed in Section 3.4.3.

Table 3.2: Calculated DC and eddy losses (total for all winding).
Std.
Ander- Ave- Std.
A B C D E F G H I J Dev. /
sen rage Dev.
(W) (W) (W) (W) (W) (W) (W) (W) (W) (W) Ave
(W) (W) (W)
(%)
DC
41724 41714 42296 42480 42588 42975 42828 42616 42822 43483 43056 42598 538 1.3
losses
Eddy
6638 7062 4308 6527 6078 8147 8835 4764 7243 6552 6944 6645 1305 19.6
losses
Total
48362 48775 46604 49008 48666 51122 51664 47380 50065 50034 50000 49244 1519 3.1
losses

Table 3.3: Calculated eddy losses (for selected discs).


Disc A B C D E F G H I J Andersen
# (W) (W) (W) (W) (W) (W) (W) (W) (W) (W) (W)
78 803 829 528 945 709 1072 1215 615 894 698 889
77 504 521 321 552 448 640 723 109 530 450 552
76 343 356 215 366 310 430 475 132 377 313 372
75 244 255 153 255 224 304 330 96 272 226 258
74 179 189 113 184 165 222 237 72 195 167 189
73 135 143 86 134 126 166 176 57 156 126 139
72 104 111 68 101 99 127 134 46 115 96 107
71 82 89 55 78 79 99 105 40 96 75 81
70 67 73 46 61 64 79 84 36 76 59 65
69 56 61 40 49 54 65 69 33 64 47 52
39 25 28 25 15 24 25 25 33 25 76 25

The Q factor was derived for each simulation results according to different formulations, as shown in
Table 3.4.

Table 3.4: Calculated Q factors.

Q-A Q-B Q-C Q-D Q-E Q-F Q-G Q-H Q-I Q-J
Lossesave(4 top discs) / 1.50
Losseave 1.63 1.64 1.42 1.71 1.55 1.77 1.86 1.30 1.66
Lossesave(3 top discs) / 1.60
Lossesave 1.75 1.76 1.50 1.86 1.66 1.93 2.04 1.38 1.79

70
Chapter 3: Benchmark of numerical tools

Lossesave(2 top discs) / 1.80


Lossesave 1.92 1.93 1.62 2.06 1.80 2.15 2.29 1.50 1.96
Losses(top disc) / 2.00
Lossesave 2.16 2.18 1.79 2.37 2.01 2.48 2.66 1.92 2.25

2000 1764
(128%)
1800
1383
1600 (100%)
1400 1070
Total losses (W)

1200 (77%)

1000
800

600
400
200
0
Q- A

Q- B

Q-C

Q-D

Q- E

Q- G
Q- F

Q-H

Q- I

Q- J

sen

e
rag
der

Ave
An
Team

Figure 3.4: Total losses simulation results for disc 78.

1400 1215
(145%)
1200

1000 836
Eddy losses (W)

(100%)

800
528
600 (63%)

400

200

0
C

I
A

J
sen

e
rag
der

Ave
An

Team

Figure 3.5: Eddy losses simulation results for disc 78.

3.4.2 Eddy losses calculations on a second transformer geometry

Another series of loss calculations were carried out with a transformer having a lower Q factor. The
electrical characteristics are listed below and the dimensions are shown in Figure 3.6.

 40/50/63 MVA, ONAN/ONAF/ONAF, 150/31.5 kV, YNd11, 50 Hz

71
Chapter 3: Benchmark of numerical tools

 ±8×1.25% On-Load Tap-Changer in HV Winding

 LV 260 turns, 90 discs, Copper Conductors dimensions: 10.6 mm ×3.75 mm

 HV 726 turns, 82 discs, Copper Conductors dimensions: 10 mm ×3.4 mm

 Uz = 12.0% @ 63 MVA and 75ºC - HV Tap 9 (150 kV)

Figure 3.6: Windings dimensions for the losses calculation on the second transformer.

The results for LV winding (rated current - tap 9) are illustrated in Figure 3.7 and Figure 3.8. The
calculated Q factors are shown in Table 3.5.

1000 858
(115%)
900
746
800 (100%)
641
Total losses top disc (W)

700 (86%)

600

500

400

300

200

100

0
A E F G H I Ave
Team

72
Chapter 3: Benchmark of numerical tools

Figure 3.7: Total losses for top disc in the LV winding of the second transformer.

500 436
(137%)
450

400
319
Eddy losses top disc (W)

350 (100%)

300
216
250 (68%)

200

150

100

50

0
A E F G H I Ave
Team

Figure 3.8: Eddy losses for top disc in the LV winding of the second transformer.

Table 3.5: Calculated Q factors for the second transformer.

Q-A Q-E Q-F Q-G Q-H Q-I


Qave(4 top discs) / Qave 1.27 1.18 1.30 1.33 1.19 1.25
Qave(3 top discs) / Qave 1.32 1.22 1.36 1.41 1.23 1.30
Qave(2 top discs) / Qave 1.39 1.27 1.44 1.50 1.26 1.38
Qave(top disc) / Qave 1.48 1.35 1.54 1.64 1.29 1.47

3.4.3 Discussion

In the present investigation, it is notable that there is a significant variation between the results obtained
by the different teams. The Q factor (for top disc) ranges from 1.79 to 2.66 (+21.9% and -18.0% from the
average) for the first transformer and from 1.29 to 1.64 for the second (12.0% and -12.1% from the
average).

The following items were identified as potential sources of divergence of the results:

 Definition of the boundary conditions. This is the most likely source of divergence and a more
detailed discussion is provided below. It is worth mentioning that the investigated transformer has
a quite significant increase of the eddy losses at the ends of the winding and it may be a case that
it is especially sensitive to the selection of boundary conditions. The combination of a conductor
axial dimension of 14.3 mm and a transformer impedance of 22.7% produce a significant radial
flux crossing the axial dimension of the conductors. A slight variation of 10% in the radial

73
Chapter 3: Benchmark of numerical tools

magnetic field calculation would result in a 20% variation of the corresponding eddy losses, as
shown in Equation (2.6) on page 14.

 Level of detail used in the geometry. The number of segments used to define the windings (one
per winding, one per disc, one per conductor, etc.) can influence the leakage flux pattern and then
the eddy loss calculation results [12].

 The analytical equation used to calculate the losses from the leakage flux. The industry generally
recognises Equation (2.6) as an adequate analytical formula to calculate the eddy losses induced in
a rectangular winding conductor by an incident (radial or axial) leakage flux. However,
manufacturers may have developed their own formula based on their experience and
investigations. These are proprietary tools and the details are not published here.

The different boundary conditions of the various 2D projection planes will influence the leakage flux
patterns especially at the ends of the windings where the losses are maximal (Figure 3.9). Additional
numerical simulations made by team Q-F has shown that the eddy losses calculated using cross section A
are 28% lower than the ones obtained using cross section C (total losses reduced by 17%). This is
explained mainly by a reduction of the radial leakage flux because of the proximity of the magnetic top
yoke to the winding ends.

The leakage magnetic flux distributions around the end of windings are different

CORE CLAMP CORE YOKE

Figure 3.9 : Leakage magnetic flux distribution for two different boundary conditions.

From this discussion, the following conclusions may be made.

 Leakage magnetic flux distributions around the end of the windings are different outside and
inside the core window.

 This change of leakage flux pattern will modify the eddy losses in the top conductors. In general,
the losses are higher outside the core window because of the increase in the the radial flux
component.

 Boundary area has a significant influence on eddy loss calculations for this model.

A comparison was previously made by CIGRE between the different magnetic field calculation programs
[11]. The set-up for calculation was quite simple and the windings were split up into about 10 axial

74
Chapter 3: Benchmark of numerical tools

segments. The calculated values for each segment were compared and put into tables. Based on different
geometries and calculation methods, the following conclusions were made, which are relevant for this
working group.

 The methods in this study will produce consistent and sufficient accurate results (within limits of
+/-10%) provided that the winding geometry, ampere-turns and core boundaries are properly
defined. In most cases the deviations are not due to the method itself but to the manner in which it
is employed.

 The effect of curvature of the winding is negligible in the cases considered.

In this investigation the segments for calculation of radial field were rather large, 10 segments for winding
height of 2400 mm. With advancing computational technology it is possible to use more elements and the
differences might be larger (for instance for the investigated transformer: 78 elements over a height of
1500 mm).

The losses obtained in the second transformer, having a lower and more usual Q factor value, are in line
with the previous CIGRE study results [11].

3.5 Temperature calculation

3.5.1 Temperature calculation using uniform losses

As stated in the modelling specifications, the temperature calculation was first performed using uniform
losses. The simulation conditions are repeated here:

 Apply 677 W per disc (no temperature compensation for the losses)

 Inlet boundary conditions:

o Oil mass flow rate of 0.78 kg/s

o Bottom oil temperature of 46.7°C (from heat-run test data)

The results are shown in Table 3.6. H, Q and S factors were calculated using the methodology described
in section 2.2. This methodology uses the mixed top-oil temperature as a reference which in this case is
80.4°C (from the heat-run test results).

75
Chapter 3: Benchmark of numerical tools

Table 3.6: Temperature calculation results using uniform losses.


Average
Top-oil Hotspot Hotspot Q S H
Team Model winding
temperature temp. location factor factor factor
temperature
S1-a THN 80.2 90.8 120.9 Disc 54 1 1.49 1.49
S1-b CFD-2D 80.2 86.9 115.0 Disc 54 1 1.48 1.48
S1-c CFD-3D 80.4 89.2 109.3 Disc 62 1 1.13 1.13
S2 THN 79.5 92.7 110.1 Disc 61 1 1.02 0.02
S4 CFD-3D 80.7 94.9 114.7 Disc 72 1 1.10 1.10
S6 THN 81.4 87.1 113.0 Disc 74 1 1.38 1.38
S7 THN 80.2 94.9 119.8 Disc 54 1 1.26 1.26
S9 THN 81.5 92.3 115.2 Disc 71 1 1.21 1.21

3.5.2 Temperature calculation using non uniform losses

Another temperature calculation was made using the average of the calculated DC and eddy losses
calculated by the WG members. Figure 3.10 summarizes the loss data used for the temperature calculation
and Table 3.7 shows the temperature calculation results.

76
Chapter 3: Benchmark of numerical tools

Disc # DC losses (W) Qave – CIGRE*


1400
78 544 839
1200 77 544 477
76 544 328
1000
Eddy loss 75 544 233
DC loss 74 544 170
Losses (W)

800
73 544 128
600 72 544 99
400 71 544 78
70 544 64
200 69 544 53
0
39 544 25
39 42 45 48 51 54 57 60 63 66 69 72 75 78 * The average of results from teams A to H (Table
Disc number 3.3) were used because results I and J were
provided later.

Figure 3.10 : DC and eddy losses used for the temperature calculation using non uniform losses.

Table 3.7: Temperature calculation results using non uniform losses.


Average
Top-oil Hotspot Hotspot Q* S H
Team Model winding
temperature temp. location factor factor factor
temperature
S1-a THN 78.9 88.5 132.0 Disc 78 2.21 0.94 2.07
S1-b CFD-2D 78.4 84.4 137.4 Disc 78 2.21 1.24 2.73
S1-c CFD-3D 79.1 87.9 134.0 Disc 78 2.21 1.00 2.20
S2 THN 80.0 91.3 128.0 Disc 78 2.21 0.78 1.72
S4 CFD-3D 78.3 91.4 127.9 Disc 78 2.21 0.77 1.71
S6 THN 78.4 83.2 128.0 Disc 78 2.21 1.10 2.42
S7 THN 78.4 91.8 133.4 Disc 78 2.21 0.85 1.88
S9 THN 79.9 88.9 136.3 Disc 78 2.21 1.00 2.21
*Q factor is estimated for a temperature of 75°C.

3.5.3 Discussion

As with the eddy loss calculation, there is a significant variation in the hotspot temperature estimation
from team members, although the location of hotspot is consistently the same – Disc 78 in the case with
non uniform losses. This is a fact because in this case, the Q-Factor is enormous in the last disc – 2.2 – and
as a result this effect overlaps the flow factor. However the values depicted in Table 3.6, when the same
exercise had been done with uniform losses, show a widespread difference in the prediction of the hotspot
location.

There are several sources of divergence, some are listed here:

 The modelling approach (THN vs. CFD): the THN approach solves the oil flow with analytical
formulas that use average velocity and temperature values; in a CFD computation, instead of
using empirical or analytical formulas, the governing differential equations are solved for the fluid
(Navier-Stokes equations) and solid (heat conduction equation) domains. Consequently, the
dynamics of the hot streak cannot be captured with the THN numerical method. Comparisons of
results obtained using THN and CFD are presented in Figure 3.11 and Figure 3.12.

77
Chapter 3: Benchmark of numerical tools

40 130
Pass 1 Pass 2 Pass 3 Pass 4
Mass flow rate fraction (%) 35 120
TNM

Max. disc temperature (°C)


TNM
CFD
30 110 CFD

25 100

20 90

15 80

10 70

5 60
Pass 1 Pass 2 Pass 3 Pass 4
0 50
0 10 20 30 40 50 60 70 80 0 10 20 30 40 50 60 70
Channel number Disc number

Figure 3.11: Flow and temperature distribution obtained with THN and 3D CFD methods (methods S1-a and
S1-c, uniform losses).

40 150
Pass 1 Pass 2 Pass 3 Pass 4 Pass 1 Pass 2 Pass 3 Pass 4
35 140
Mass flow rate fraction (%)

TNM TNM

Max. disc temperature (°C)


CFD CFD
30 130

25 120

20 110

15 100

10 90

5 80

0 70
0 10 20 30 40 50 60 70 80 0 10 20 30 40 50 60 70
Channel number Disc number

Figure 3.12: Flow and temperature distribution obtained with THN and 3D CFD methods (methods S1-a and
S1-c, non uniform losses).

 2D or 3D modelling for CFD computations: since the windings are cylindrical in shape, the
analysis can be performed in 2D (i.e., axisymmetrical) mode. However, the drawback of this
simplification is that the impact of sticks and spacers on the temperature rise cannot be fully
considered. In fact, these components reduce the exposed area of the discs to the oil and change
the bulk velocity of the fluid in the cooling channels, thus affecting the heat transfer [114].

 Even the 3D modelling with CFD computations can produce different results because of different
models or parameters intrinsic to each CFD code and user.

78
Chapter 4: Dynamic thermal modelling

Chapter 4: Dynamic thermal modelling

4.1 Introduction

As dynamic thermal models are used in several on-line monitoring system applications, their empirical
verification, statistical evaluation and fundamental development are of key importance. Therefore, the
obvious objective of this dynamic modelling chapter is to evaluate the in-service long-term accuracy of
the three commonly used International Loading Guide dynamic thermal models i.e. IEC 60354 (1991),
IEC 60076-7 (2005) and IEEE Std C57.91 (2011) Annex G. The models' exponents and constants are
reviewed and discussed here as well as uncertainties in their determination. These parameters are usually
obtained from the heat run test procedure, which is also explained. This chapter provides the reader with a
comprehensive and well-referenced introduction to state-of-the-art of this topic.

4.2 Review of the state-of-the-art

On-line monitoring of transformer temperatures is an important contributor in developing a strategy to


maximise service life. Traditionally the winding and oil temperatures in power transformers are measured
by the application of a thermal image device, as shown in Figure 4.1 [53], commonly referred to as a
winding temperature indicator.

A thermocouple sensor is immersed in the top oil in the tank, i.e. oil pocket. A proportion of load carried
by the heating element is adjusted by the matching resistance to obtain the appropriate “image” of the
hotspot temperature. The top-oil to hotspot temperature gradient, obtained during heat-run test, is added
to the top-oil temperature reading and the thermal image device copies the temperature of the hottest spot
of the winding. The control of cooling equipment is based on these temperature readings by the contact
thermometers fitted in the associated box. Over the years this principle has been unchanged as defined in
the former IEC 354 Loading Guide for Oil-Immersed Power Transformers.

Direct measurement of actual transformer winding temperatures using fibre optic probes has been
increasing since the mid-1980s [30], [53], [56], [57], [58], [59], [60], [61] and [62]. By analysing
measured results from tested power transformers it has been noticed that the hotspot temperature rise over
top-oil temperature following load changes is a function dependant on time as well as transformer load
(overshoot time dependent function) shown in Figure 4.2. Similar results were obtained for distribution
transformers with external cooling [42]. The maximum values and shapes of this function (represented in
terms of the overshoot factor, Bp) for different transformers with external cooling, different loads,
different oil circulation modes in the windings (zig-zag and axial) and different cooling modes are given
in [58].

79
Chapter 4: Dynamic thermal modelling

Figure 4.1: Thermal Image Device (WTI-winding temperature indicator) [53].

It has been shown that the dynamic winding hotspot calculation methods proposed in the loading guides,
[22] and [54], yield significantly lower hotspot temperature than actual values during transients, especially
in the case of a short-term emergency loading [59]. The same conclusion is also valid for hotspots in the
core and structural parts if the principles proposed in [22] and [54] are applied. This is a critical limitation
in the open market environment, where network planners, operators and asset managers are trying to fully
exploit the capacity of existing equipment. It has also been observed that the top-oil temperature time
constant is shorter than the time constant suggested by the former loading guide [54] especially where the
oil is guided through the windings in a zig-zag pattern for the ONAN and ONAF cooling modes [58].
These values are estimated by exponential curve fitting in order to obtain the relevant numerical
quantities, due to the fact that transformer time constants are functions of the transformer oil viscosity [63]
and further, they are dependent on the oil temperature [22].

Figure 4.2: Normalized time variation of hotspot temperature rise above top-oil temperature, f2(t), (in tank)
for a step increase in load current. Bp is overshoot factor.

A thermal test made on distribution transformers without external cooling [67], showed that the hotspot
temperature rise over top-oil temperature for the oil temperature measured in the oil pocket due to a
80
Chapter 4: Dynamic thermal modelling

change in load is an exponential function, f2(t), with the time constant equal to the winding time constant,
as shown in Figure 4.2 (the dashed line). It has also been observed that this top-oil time constant is longer
than the time constant obtained for large power transformers with zig-zag oil circulation through the
windings.

In order to overcome these thermal obstacles and to increase transformer loading capacity, different
calculation procedures for the winding hotspot temperature response to load changes have been proposed
by many authors. For example, Aubin and Pierce dealt with the overshoot phenomenon by avoiding its
direct modelling. They obtained hotspot temperature calculation methods that were based on the bottom-
oil temperature. These methods are published in scientific papers [43], [64] and [66]. Pierce’s method is
also presented as a more complex dynamic hotspot temperature calculation procedure in the IEEE
Loading Guide Annex G in [22].

Different solutions have also been suggested, such as Alegi, Blake, Declercq, Lesieutre, Pierce, Pradhan,
Radakovic, Ryder, Susa, Tang, and Van der Veken in references [67], [68], [69], [70], [71], [72], [73],
[74], [75], [76], [77], [78], [79], [80], [81], [83] and [84]. It is necessary to stress the importance of the
work published by Swift for further understanding of the state of the art in transformer thermal modelling,
[85] and [86]. A different approach named Dynamic Thermal-Hydraulic Network Model, DTHN,
suggested in references [87], [88], [89], considered hydraulic resistances of windings, heat transfer
coefficients for outer cooling depending on environmental influences or changed cooling conditions, heat
transfer coefficients for inner cooling, load for different tap positions for each winding, change of cooling
conditions.

The more recent IEC 60076-7 Loading Guide for oil immersed transformers [21] recommends the thermal
model, where the variation of the winding hottest-spot temperature rise over the top-oil temperature after a
load change is defined by exponential functions with constant parameters. The model depends only on
data received in a normal heat run test, (i.e., the top-oil in the tank of the transformer, the average
winding-to-top-oil gradient) [58], [90] and [91].

From an academic point of view, the IEC exponential top-oil model has been considered inaccurate since
the conventionally measured top-oil temperature below the tank top offers a less representative
temperature calculation reference point, compared with the oil temperature in the winding cooling ducts
[96]. However in ON- and OD- cooled transformers this inaccuracy has been considered as marginal
since the steady-state volume rate of flow of liquid through the windings is, in principle, equal to the rate
of flow through the radiators. Nevertheless, in-service observations have shown that in case of low
ambient temperatures with restricted oil flow in windings, the use of conventional top-oil temperature
models based on the constant parameters can lead to an underestimation of maximum oil and hotspot
temperatures, potentially resulting in accelerated transformer ageing [93], [94].

Finally, CIGRE WG A2.24 published a Brochure on thermal performance of power transformers [92].
The Brochure, among other relevant topics, dealt with the testing of thermal performance of power
transformer through defining the temperature rise test to verify transformers loading capability and the
special temperature rise test to verify long term emergency loading capability. Also, thermal design and
thermal modelling of power transformers were carefully discussed. Under dynamic thermal modelling, all
loading guide models, (IEC and IEEE), were compared and tested by using thermal characteristics
obtained using fibre optic temperature sensors during a typical factory heat-run test. The calculation
procedure, corresponding formulas and results for the IEC winding and oil exponents were then discussed
and published.

81
Chapter 4: Dynamic thermal modelling

In-service analysis of dynamic thermal model accuracy by use of adaptive thermal models has shown that
error deviations of the models arise also from inadequate control of input-data and lack of significant
driving variables in thermal models used [98]. The development of data quality control techniques as well
as adaptive thermal models and their proper in-service application represents a vital step in improving
accuracy and reliability of in-service transformer temperature prediction [93], [99], [100], [101]. Some of
the adaptive thermal models are represented by grey and black-box modelling techniques with application
of neuro-fuzzy and soft computing [123], [124].

4.3 Loading Guide Dynamic Thermal Models (DTMs)

4.3.1 General

The chapter compares and discusses the in-service accuracy of dynamic thermal models (DTMs),
presented in the international standard Loading guides, i.e. the IEC 60354 model (1991) [54], IEC 60076-
7 model (2005) [21], IEEE C57.91 Clause 7 (2011) [21] and IEEE C57.91 Annex G (2011) [22], and their
sensitivity to changes in ambient temperature and related phenomena.

The aim of the DTMs is to provide a simplified solution of a complex three-dimensional physical
phenomenon of heat transfer inside transformer in the form of critical transformer temperatures (i.e. top-
oil and hotspot temperatures) that can be applied for real-time monitoring, diagnostics and transformer
protection applications.

In order to quantify objectively the in-service error of DTMs and analyse its origin, specific accuracy
metrics, presented in Section 4.4 are needed. DTM errors can be categorised in short-term (transient) or
long-term (seasonal). Their origin may be identified and quantified using statistical error analysis, which
is also necessary as the driving variables of DTMs in service conditions (i.e. the load, cooling profile and
ambient conditions) are stochastic by nature.

The IEC 60354, IEC 60076-7 and IEEE C57.91 Clause 7 Loading Guide DTMs have been widely applied
primarily as the models require just a basic set of input parameters that can be obtained from a standard
transformer heat-run test report. The three models’ transient equations are derived using a simplified
transformer thermal diagram, representing a linear transformer temperature distribution inside a
transformer. In the simplified thermal diagram, the assumptions are made that the winding coil to duct-oil
temperature between the bottom and top winding parts increase linearly and that the winding coil to duct-
oil temperature gradient is constant with height. The three models also neglect the difference between the
tank top-oil and top-duct oil temperature (i.e. the multi-flow model [106]), as well as temperature-based
conductor resistance change and coolant viscosity change, resulting in limited accuracy during short-term
overload and low ambient temperature conditions. On the other hand, the IEEE C57.91 Annex G (2011)
[22] model incorporates a refined multi-flow thermal model, which incorporates change of losses with
temperature, liquid viscosity temperature compensation and parameterisation of the thermal model’s
components time constants.

A brief formulation of the DTMs is given in the following chapters with definitions and descriptions of
their parameters, input (driving) variables and output results. Temperatures are noted with Latin letters
and °C units, temperature rise values (above ambient temperature) and temperature gradients (between
two reference points) are noted with Greek letters and K units.

82
Chapter 4: Dynamic thermal modelling

4.3.2 IEC 60354 model

The top-oil temperature rise and hotspot gradient temperature models of the IEC Loading Guides are
expressed by first-order differential equations (4.1) and (4.2).

Steady state top-oil  to and hotspot temperature rise hs values (above the ambient temperature Tamb) are
nonlinear functions whose parameters depend on load factor (k) and cooling system type.

dTto
 to   tor  Tto  Tamb  (4.1)
dt

dThs
 to   tor  Ths  Tamb  (4.2)
dt

where:

x
 1  R  k2 
 to   tor    (4.3)
 1R 

 hs  Hg r  k y (4.4)

In the case that ambient temperature is taken as constant and load changes as step functions, the  to and
hs model transient response may be expressed in the form of an exponential function, as explicated in
the IEC and IEEE loading guides. In the usual case, model input data are not step functions, and a
simplified iterative solution (i.e. forward Euler method approximation) of the above equations can be
established by conversion of equations (4.1) and (4.2) into difference equations with discrete time step t
as in (4.5) and (4.6).

t  
x
 1  R  k  i 2 
Tto  i  1  Tto  i    to     Tto  i   Tamb  i  (4.5)
 to   1  R 
 
 

t 
Ths  i  1  Ths  i   Hgr  k  i   Ths  i   Tto  i 
y
 (4.6)
hs

4.3.3 IEC 60076-7 model

The IEC 60076-7 model [22] uses the same top-oil model approach as the IEC 60354 and refines the top-
oil time constant (  to ) to reflect the stagnation of the bottom part of the bulk oil in the tank with ON
transformer cooling types, using the k11 correction factor.
x
 1  R  k2  dTto
 tor    k11   to  Tto  Ta  (4.7)
 1R  dt

83
Chapter 4: Dynamic thermal modelling

The IEC 60076-7 upgrades the hotspot gradient (Hgr) transient behaviour by imposing an overshoot
transfer function to Hgr, which imitates the effect of mass inertia in natural convection based duct oil flow
and consequent temperature behaviour of oil in the windings. The overshoot transfer function is defined
by a 2nd order linear system equation that can be translated into two 1st order differential equations (4.8),
(4.9) and rejoined for final solution of Ths.

dThs1
k21  Hgr  k y  k22  hs  Ths1  Tto  (4.8)
dt

 to dThs2
 k21  1 Hgr  k y   Ths2  Tto  (4.9)
k22 dt

Ths  Ths1  Ths2 (4.10)

4.3.4 IEEE C57.91 Annex G model

In comparison with IEC models, the IEEE C57.91 Annex G DTM [23] is based on energy balance
equations, describing heat transfer between the basic transformer elements. In principle, the hotspot
temperature is calculated as the sum of ambient temperature, bottom-oil rise, bottom-oil to top-duct-oil
gradient and top-duct-oil to hotspot gradient as in (4.11).

hs  amb  bo   wo   hs (4.11)


bo wo

The steady-state values of bottom oil rise, average winding and hotspot temperature rise are nonlinearly
proportional to instantaneous losses and are dependent on the cooling mode.

A transient temperature increase of a particular transformer element (i.e. core and coils, tank and fittings
and/or active cooling liquid) after a period of time ( t ) is determined so that the differential temperature
change of that element is proportional to the ratio of balance of generated (load dependent) and dissipated
(cooling) losses (Qgenerated  Qdissipated ) and thermal capacitance MwCp of the basic transformer model
elements as in (4.12).

Qgenerated  Qdissipated
  t   (4.12)
MwCp

A temperature transient for each element is then iteratively calculated using (4.13).

Tx  i  1  Tx  i   Tx (t ) (4.13)

Such temperature calculation methods allow iterative updates of the thermal model parameters, consider
variation of ambient temperature, oil viscosity, winding resistance and related losses changes, as well as
incorporate logical temperature conditions in the calculation procedure. As the complete model
description and calculation method is rather extensive and openly available [22], [106], further model
explanation is omitted.

84
Chapter 4: Dynamic thermal modelling

4.4 In-service accuracy evaluation of Loading Guide Dynamic Thermal Models

4.4.1 Description of the transformer under investigation

In the presented case study, the top-oil and hotspot temperatures, predicted by three DTMs (IEC
60354(1991), IEC 60076-7(2005), IEEE Std C57.91(2011) Annex G), are compared with a year period of
in-service fiber-optic probe temperature measurements, installed in a transmission transformer [12].

The transformer under test is a medium size power transformer with rated data 66 MVA, 225/26.4 kV,
YNd1 winding arrangement, short-circuit impedance of 22.7% and ONAN/ONAF cooling with a zig-zag
cooling arrangement. The transformer was refurbished after a fault caused by conductors overheating at
the top of the low-voltage (LV) windings and fiber-optic probes were inserted in the radial spacers of new
LV windings in direct contact with conductor insulation at the hotspot-location. An on-line monitoring
system was installed, providing continuous measurements of the currents, cooling mode
(ONAN/ONAF1/ONAF2), ambient, top-oil and LV hotspot temperatures during transformer in-service
operation. The data were logged once per minute into a remotely accessible database.

The transformer heat-run test results indicated a rated top-oil and bottom-oil temperature of 80.4°C and
46.7°C respectively at an ambient temperature of 30.2°C. The average oil temperature rise was 33.4 K,
the average winding temperature rise 61.6 K and the winding to oil gradient was determined to be gr =
28.2 K. The hotspot to top-oil gradient Hgr of 40.0 K and resulting hotspot factor (H) of 1.4 was evaluated
at the 76th disc of the LV winding based on direct temperature measurements in summer 2009 [12]. The
hotspot gradient at the 76th disc was also determined by network thermal model (Hgr= 32.5 K … 41.8 K)
and computational fluid dynamics simulations (Hgr= 31.5 K … 41.7 K) [12].

The parameters for the three investigated models are shown in Table 4.1, Table 4.2 and Table 4.3.

Table 4.1: IEC60354 model parameters.

Load to no-load losses ratio R 8.4

Rated top-oil temp. rise tor 50.0

Rated hotspot gradient Hgr 40.0

Top-oil time constant to 150

Hotspot time constant ths 7

Top-oil exponent x 0.80

Hotspot gradient exponent y 1.6

Table 4.2: Additional IEC6076-7 model parameters.


to correction factor k11 0.5
Overshoot factor k21 2.0
Overshoot factor k22 2.0
Hotspot gradient exponent y 1.3

85
Chapter 4: Dynamic thermal modelling

Table 4.3: IEEE Annex G model parameters.

MVA Base for Loss Data 66.0 MVA


Temperature Base for Loss Data 75 °C
Winding DC Losses 331459 W
Winding Eddy Losses 0 W
Stray Losses 93031 W
Core Losses 50300 W
Cooling mode ONAF
Nameplate MVA 66 MVA
Guaranteed Avg. Wnd Rise 65 K
Rated Average Winding Rise 61.6 K
Rated Hotspot Rise 90 K
Rated Top Oil Rise 50 K
Rated Bottom Oil Rise 16.5 K
Rated Ambient Temperature 30 °C
p.u. Eddy Loss at Hotspot 0.55 p.u.
Winding Time Constant 7 min
p.u. Winding Height to Hotspot 0.97 p.u.
Weight of Core & Coils 73200 lbs
Weight of Tank & Fittings 54200 lbs
Oil volume 8960 gals
Cooling fluid Mineral oil
Duct oil rise exponent (x) 0.5
Average oil rise exponent (y) 0.9
Radiator oil rise exponent (z) 0.5

4.4.2 Comparison of the three Loading Guide Dynamic Thermal Models

4.4.2.1 General

This section presents a comparison of the three Loading Guide dynamic thermal models (DTM) with
respective direct temperature measurements during a one-year observation period. The measured model
input (Indexes used in figures: ambient temperature, load, cooling status) was used to calculate the
modelled top-oil temperature rise above ambient (iec354.T_to, iec767.T_to, ieeeAG.T_to) and hotspot gradient
(iec354.T_hgr, iec767.T_hgr, ieeeAG.T_hgr), using the equations (1)–(12) for the IEC models and calculation
procedure from [106] for the IEEE Annex G model. To evaluate the modelled results, the top-oil and
hotspot and ambient reference temperature (meas.T_to, meas.T_hgr, meas.T_amb) were measured and logged
minute by minute by a transformer on-line monitoring system. Two characteristic observation periods, 1
month of summer and 1 month of winter data were processed to analyse the effect of variation of limit
ambient temperature conditions on model accuracy.

DTM accuracy is evaluated by the two following methods. The first method is the statistical evaluation of
the DTM error, i.e. the evaluation of difference between modelled and measured temperature values. The
first method is focused on statistical evaluation of the modelling error of the top-oil rise prediction for the
three models (to_err-iec354.T_to, to_err-iec767.T_to, to_err-ieeeAG.T_to), of the hotspot gradient (to_err-
86
Chapter 4: Dynamic thermal modelling

iec354.T_hgr, to_err-iec767.T_hgr, to_err-ieeeAG.T_hgr) and their cumulative value – the hotspot temperature
rise (to_err-iec354.T_hs, to_err-iec767.T_hs, to_err-ieeeAG.T_hs). Heat-run test parameters and recommended
parameter values from the loading guide in Tables 3.1–3.4 were used for calculation of the DTMs.

Results of the first interpretation method, extracted from two periods during the yearwith extreme
temperatures, are presented in the Figures below. In the left-side figures, temperature rise curves and their
respective error curves are displayed in addition to per cent load factor, cooling status (ONAN=0%, ONAF1
= 50%, ONAF2=100% cooling power) and ambient temperature.

In the right side figures, error duration curves are presented (iec354.err_sort, iec767imp.err_sort,
ieeeAG.err_sort) indicating the per cent time sorted error duration inside the observation time-window and
an error statistics table to quantify the individual DTM’s error parameters: The average error represents a
typical temperature offset of the model against measurements, indicating how well the steady-state
parameters of the DTM are determined during the heat-run test. The slope of the error duration curve
indicates how well the DTM mimics the static nonlinearity (i.e the top-oil/hotspot exponent), whereas the
standard deviation and min/max error values evaluate the accuracy of the DTM’s dynamic behaviour.

The second DTM accuracy evaluation method presented is adaptive DTM parameterisation. The adaptive
parameterisation enables determination of rated load (i.e. nominal) DTM parameter values, based on in-
service measurements with fluctuating load and variable ambient temperatures. In the method under
study, based on IEC 60076-7 DTM, the investigated DTM parameters were “unlocked”, taken as variables
and optimised in order obtain a minimum deviation of the DTM to the reference temperature
measurements. The top-oil temperature rise, hotspot gradient and the k- factor values were fitted using a
nonlinear regression algorithm [101] and are presented in Table 4.4 and Table 4.5. Two observation time
windows (5 and 21 days) were analysed to observe seasonal DTM accuracy variation, to evaluate the
difference between the shot-term and long-term DTM behaviour and to check consistency of the
evaluation method.

Sorted error duration (%) 29.8.10 3.9.10


10%

20%

30%

40%

50%

60%

70%

80%

90%
Load, cooling (%)
0%

80 150%
100% 2
70 50% 0
60 0%
-2
temperature difference (K)

50
-4
Temperature (°C)

40 -6
30
-8
20 -10
10 -12
0 Erro -14
-10 (°C) -16
-20 -18
29.8.10

30.8.10

31.8.10

1.9.10

2.9.10

iec3 5 4 .er r _sor t iec7 6 7 .er r _sor t


ieeeA G.er r _sor t

iec354.T_to to_err-iec354.T_to A mbient av g.err st.dev max min


ieeeA G .T_to to_err-ieeeA G .T_to meas.T_to to_err-iec354.T_to -1.4 2.1 2.3 -6.1
iec767.T_to to_err-iec767.T_to C ooler status
to_err-iec767.T_to -1.5 2.4 2.3 -6.7
Load
to_err-ieeeAG.T_to -1.6 2.1 3.1 -5.2

Figure 4.3: Top-oil & top-oil error vs. time (summer). Figure 4.4: Sorted top-oil error duration.

87
Chapter 4: Dynamic thermal modelling

4.4.2.2 Top-oil model results

Comparison of the measured and modelled top-oil temperatures for the summer period shows a similar
behaviour of all three DTMs. There is a good correlation to reference measurements with an average error
of < 2 K, considering the cooling mode variations were not taken into account by any of the three DTMs.

During the winter period, top-oil temperatures are underestimated by all three DTMs, the average model
underestimation is increased by ca. 6 K, compared to the summer period. As the transformer cooling was
operating in a forced ONAF2 mode, extremely low bottom oil temperatures below 0 °C were reached
during the observed period. Consequently higher oil viscosity was causing reduced heat transfer from
transformer active parts and higher measured top-oil temperatures than expected by the three DTMs.
Underestimation of the top-oil temperatures can be explained by the fact that that viscosity correction
equations are not included in any of the three top-oil model algorithms.

Load, cooling (%) Sorted error duration (%) 29.1.10 3.2.10


80 150%
100%

10%

20%

30%

40%

50%

60%

70%

80%

90%
0%
70 50%
2
60 0%
0
50
Temperature (°C)

-2

temperature difference (K)


40
-4
30
-6
20
-8
10
-10
0 Erro -12
-10 (°C
-14
-20
-16
29.1.10

30.1.10

31.1.10

1.2.10

2.2.10

-18
iec3 5 4 .er r _sor t iec7 6 7 .er r _sor t
iec354.T_to to_err-iec354.T_to A mbient
ieeeA G.er r _sor t
ieeeA G .T_to to_err-ieeeA G .T_to meas.T_to
iec767.T_to to_err-iec767.T_to C ooler status av g.err st.dev max min
Load to_err-iec354.T_to -7.2 3.0 -1.5 -16.3
to_err-iec767.T_to -7.2 2.8 -2.1 -12.5
to_err-ieeeAG.T_to -9.7 2.9 -0.5 -16.5

Figure 4.5 : Top-oil & top-oil error vs. time (winter). Figure 4.6 : Sorted top-oil error duration.

Results of the IEC 60076-7 adaptive top-oil DTM are presented in Table 4.4 for the two extreme
temperature time periods in Aug. and Jan. 2010. The adapted values of the rated top-oil rise, shown in
Table 4.4, can be used to quantify the seasonal variation of DTM parameters independently of the load
profile. There is a ca. 7 K difference between the summer and winter DTM parameters (summer
conditions rated top-oil rise 51-52 K, winter 56…58 K). The comparison of adaptive DTM
parameterisation for the 5-day and 21-day periods shows a similar error deviation of the results ( < 2 K for
summer, < 3 for winter), indicating that the evaluation method is relatively independent of the observation
time window length.

88
Chapter 4: Dynamic thermal modelling

Table 4.4: Fitted top-oil parameters with statistical error evaluations (IEC 60076-7 model).
Time period Avg. Avg. Avg. Std. err. Max Min err. Rated k11
amb.(°C) load (%) err. (°C) dev. (°C) err. (°C) (°C) TO
rise

29.8.10-3.9.10 27 65 <0.1 0.9 2.4 -2.2 52 0.78


29.8.10-20.9.10 19 58 <0.1 1.9 6.8 -4.3 51 0.79
29.1.10-3.2.10 -15 91 <0.1 2.8 5.5 -6.6 58 0.79
29.1.10-19.2.10 -7 80 <0.1 2.6 5.6 -9.6 56 0.82
05.1.10-10.1.10 -6 78 <0.1 2.3 3.5 -5.4 56 0.75
05.1.10-26.1.10 -5 78 <0.1 2.6 4.8 -8.0 57 0.86

4.4.2.3 Hotspot gradient model results

The same comparative procedure as for modelled top-oil temperature is repeated for the hotspot gradient
(Hgr). The measured reference hotspot gradient (Hgr) is calculated as the difference between the hotspot
and top-oil temperature measurements. The DTM simulation results indicate general overestimation of
Hgr for all three models which is reflected in a positive offset of the sorted error duration curve. The
IEEE model shows the smallest deviation in terms of dynamic behaviour, reflected in the smallest
standard error deviation.
Sorted error duration (%) 29.8.10 3.9.10
10%

20%

30%

40%

50%

60%

70%

80%

90%
0%

Load, cooling (%) 10


70 150 8
100 6
60 50%
0% 4
temperature difference (K)

50 2
0
40
Temperature (°C)

-2
30 -4
-6
20 -8
10 -10
-12
0 -14
Err -16
-10 (° -18
-20 -20
iec3 5 4 .er r _sor t iec7 6 7 .er r _sor t
29.8.10

30.8.10

31.8.10

1.9.10

2.9.10

ieeeA G.er r _sor t

iec354.T_hgr err.iec354.H gr A mbient ieeeA G .T_hg av g.err st.dev max min


err.iec354.Hgr 2.2 1.1 4.4 -1.6
err.ieeeA G .H gr meas.Hgr iec767.T_hgr err.iec767.Hg
err.iec767.Hgr 4.8 2.8 10.0 -1.3
C ooler status Load
err.ieeeAG.Hgr 4.9 0.7 6.9 1.8

Figure 4.7: Hgr & Hgr error vs. time (summer). Figure 4.8: Sorted Hgr error duration (summer).
Winter period hotspot gradient comparisons in Figure 4.9 and Figure 4.10 demonstrate the impact of low
ambient temperature conditions on Hgr sub-model behaviour. An increase of oil viscosity and resulting
decrease of heat transfer from the windings causes an increase of the hotspot gradient and strong
underestimation of the two IEC Hgr DTMs. With the IEEE model, which includes the viscosity
compensation inside the Hgr sub-model, the error offset is noticeably smaller. As the dynamics of the

89
Chapter 4: Dynamic thermal modelling

typical load profile during winter is reduced, the differences in dynamic behaviour of the three DTMs are
also smaller. This is reflected in the similar slopes of the sorted error duration curves.
Sorted error duration (%) 29.1.10 3.2.10

10%

20%

30%

40%

50%

60%

70%

80%

90%
0%
Load, cooling (%)
70 150 10
100 8
60 50% 6
0%
50 4

temperature difference (K)


2
40
Temperature (°C)

0
-2
30
-4
20 -6
-8
10 -10
0 -12
Err -14
-10 (° -16
-18
-20
-20
29.1.10

30.1.10

31.1.10

1.2.10

2.2.10

iec3 5 4 .er r _sor t iec7 6 7 .er r _sor t


ieeeA G.er r _sor t
iec354.T_hgr err.iec354.H gr A mbient ieeeA G .T_hg
av g.err st.dev max min
err.ieeeA G .H gr meas.Hgr iec767.T_hgr err.iec767.Hg err.iec354.Hgr -11.8 2.4 -6.3 -19.0
C ooler status Load err.iec767.Hgr -10.8 2.9 -2.6 -17.1
err.ieeeAG.Hgr -3.4 2.0 4.8 -9.4

Figure 4.9: HGr & Hgr error vs. time (winter). Figure 4.10: Sorted Hgr error duration (winter).

Results of the IEC 60076-7 adaptive Hgr sub-model are presented in the Table 4.5. A seasonal drop of
Hgr by 13 K – 20 K can be observed due to high oil viscosity changes. The optimal fitted values of the
overshoot factors k21 and k22 remain stable and independent of ambient temperature.
The error is the smallest at high ambient temperatures, which complies with the fact that the model
parameters are established during a temperature rise test in a test environment at an ambient temperature
of about 30 °C.
Table 4.5: Fitted Hgr parameters with statistical error evaluations (IEC 60076-7 model).

Time period Avg. amb. Avg. Std. err. max/min Rated Hgr k21 k22
temp. (°C) load (%) dev. (°C) err. (°C)

29.8.10-3.9.10 27 65 0.7 1.8/-2.5 31.8 1.5 1.1


29.8.10-20.9.10 19 58 1.2 3.1/-5.4 32.7 1.6 1.0
29.1.10-3.2.10 -15 91 2.3 6.0/-6.4 52.3 1.0 1.0
29.1.10-19.2.10 -7 80 4.1 8.8/-10 47.1 1.4 1.0
05.1.10-10.1.10 -6 78 2.3 6.7/-9.1 45.0 1.5 1.0
05.1.10-26.1.10 -5 78 3.1 6.7/-9.1 45.0 1.5 1.0

4.4.2.4 Hotspot model results

Comparison of the hotspot results shows the cumulative accuracy of the top-oil and Hgr models. The
DTMs are using input parameters from Table 4.1 to Table 4.3. The purpose of the comparison below is to
quantify the DTM hotspot temperature error at extreme ambient temperatures for the three common
DTMs.
90
Chapter 4: Dynamic thermal modelling

It can be observed that the top-oil rise and hotspot gradient errors partially compensate during the summer
period. On the other hand, during extreme winter period all models underestimated the hotspot
temperature. These results are expected as the IEC hotspot temperature models are based on the constant
parameters, i.e. not considering change in the oil viscosity and loss variation with the temperature.
Although, the IEC 60076-7 estimates sufficiently well the hotspot temperature due to sharp changes in the
transformer load (i.e. the fact that it takes some time before the oil circulation has adapted its speed to
correspond to the increased load level is modeled by the overshoot function) yet under extreme sub-zero
condition the expected function peaks are much higher than considered in the model [64]. The IEEE
model calculates the hotspot temperature based on the winding top-duct to bottom-oil temperature rise
with fixed time constant equal to the winding time constant and assumes that the bottom oil rise follows
the same time constant as the top-oil temperature rise. Again, while under moderate ambient conditions
this parameter set up works very well, it does not work during sub-zero ambient, which cause changes in
the operating oil volume and consequently in the transformer effective cooling surface needed for the heat
dissipation. Overall, the IEEE Annex G and IEC 60076-7 DTMs yield the same peak hot temperatures, of
course, lower than the measured ones. The IEC 60354 DTM can work well during summer periods if the
sharp changes in the loading pattern are not present.
Load, cooling (%) Sorted error duration (%) 29.8.10 3.9.10
120 150
100

10%

20%

30%

40%

50%

60%

70%

80%

90%
110

0%
100 50%
0% 14
90 12
80 10
70 8
Temperature (°C)

6
60
temperature difference (K)

4
50 2
40 0
-2
30 -4
20 -6
10 -8
-10
0 Err -12
-10 -14
(° -16
-20 -18
-30 -20
-22
29.8.10

30.8.10

31.8.10

1.9.10

2.9.10

-24
-26
-28
iec354.T_hs err.iec354.H s A mbient ieeeA G .T_h -30
err.ieeeA G .H s meas.H s iec767.T_hs err.iec767.H iec3 5 4 .er r _sor t iec7 6 7 .er r _sor t
C ooler status Load ieeeA G.er r _sor t

av g.err st.dev max min


err.iec354.Hs 0.7 2.9 5.6 -6.7
err.iec767.Hs 3.3 4.6 9.9 -5.8
err.ieeeAG.Hs 3.3 2.3 9.7 -2.4

Figure 4.11: Hotspot temperature & hotspot error vs. Figure 4.12: Sorted hotspot error duration
time (summer). (summer).

91
Chapter 4: Dynamic thermal modelling

Load, cooling (%) Sorted error duration (%) 29.1.10 3.2.10


120 150
100

10%

20%

30%

40%

50%

60%

70%

80%

90%
110

0%
100 50%
0% 14
90 12
80 10
70 8
Temperature (°C)

6
60

temperature difference (K)


4
50 2
0
40 -2
30 -4
20 -6
-8
10 -10
0 -12
Err -14
-10 (°C -16
-20 -18
-30 -20
-22
29.1.10

30.1.10

31.1.10

1.2.10

2.2.10
-24
-26
-28
iec354.T_hs err.iec354.Hs A mbient ieeeA G .T_h -30
iec3 5 4 .er r _sor t iec7 6 7 .er r _sor t
err.ieeeA G .Hs meas.H s iec767.T_hs err.iec767.H
ieeeA G.er r _sor t
C ooler status Load
av g.err st.dev max min
err.iec354.Hs -19.0 3.7 -12.0 -30.6
err.iec767.Hs -18.0 3.8 -9.6 -27.4
err.ieeeAG.Hs -13.0 2.9 2.4 -22.2

Figure 4.13: Hotspot temperature & hotspot error vs. Figure 4.14: Sorted hotspot error duration (winter).
time (winter).

92
Chapter 4: Dynamic thermal modelling

4.4.3 Discussion and conclusion

The presented case study quantifies and statistically evaluates the in-service accuracy of the three
commonly used Loading guide models. The top-oil model evaluation shows that the IEC as well as IEEE
models underestimate the top-oil rise values in low ambient temperature conditions because of the missing
top-oil viscosity sub-model. The results of the IEC hotspot-gradient models are analogous, whereas IEEE
Annex G hotspot gradient sub-model shows improved low temperature accuracy for the cost of higher
DTM complexity.

However, all three models fail to match the measured hotspot temperature during sub-zero ambient
conditions irrespective of their complexity. The reason for this one can find is in change of the
transformer effective cooling surface resulting from decreased oil circulation affected by increased
viscosity, which results in the increased volume of stagnant bottom oil as well as lack of the oil circulation
through the attached cooler/radiator units. In turn, this directly affects the temperature rises and
corresponding models' parameters, which will be changed beyond expectation set in the models. In the
worst case, the transformer external cooling units are literally cut off and the effective tank surface is
significantly decreased resulting in the transformer heavy overloading condition, as once the cooling units
are ineffective the transformer rating is changed proportionally.

As transformer winding hotspot temperature is one of the most critical parameters when defining the
power transformer thermal conditions and overloading capability beyond the nameplate rating, further
research and development is needed to improve the existing models. Accurate and timely utilized thermal
models allow both transformer manufacturers and network operators to run different loading and ambient
scenarios and, by analysing the results, improve the transformer design (costs, size and load carrying
capacity).

4.5 Transformer Dynamic Thermal Rating evaluation

Temperature of transformer insulation system is the basic criterion which limits the transformer loading
capabilities. Loading beyond transformer nameplate rating and evaluation of cumulative loss of
transformer technical lifetime have been practised since many decades and have been described in
literature [58], [65] as well as in current IEC and IEEE Loading Guides [21], [23]. With development of
on-line transformer monitoring applications and inclusion of dynamic thermal models, real-time loading
and rating of transformers by hotspot temperature instead by nominal current is viable, also known as
transformer Dynamic Thermal Rating (DTR). Transformer DTR offers network operators to take full
advantage of the transformer thermal performance enables safe overloading and keeps the transformer
lifetime consumption under control.

Dynamic Thermal Model (DTM) enables prediction (calculation) of the transformer hotspot temperature
from the measured ambient temperature, cooling and load profiles. Transformer DTR evaluation on the
other hand, represents a reversed DTM process, i.e. determination of the load factor limit (kcDTR) that
allows transformer operation below the chosen hotspot temperature limit. DTR can be utilized for various
purposes and conditions, as shown in Table 4.6 [21], [23]. These conditions define the criteria for
calculation of the maximum load factor.

93
Chapter 4: Dynamic thermal modelling

Table 4.6: Transformer dynamic thermal rating types with criteria for the load factor limit
Dynamic thermal rating type Load factor limit criterion
Normal cyclic loading Hotspot temperature limit
(Normal life expectancy loading)
Planned loading beyond nameplate Daily Loss-of-life (LOL)
Long time emergency loading LOL, bubbling inception temperature
Short time emergency loading Bubbling inception temperature

To predict the impact of transformer loading on its lifetime expectancy, it is crucial to understand the basis
of the insulation system ageing process. Ageing of transformer is a complex, time-integrated function of
temperature, moisture, acidity and oxygen content of paper-oil insulation system [102]. Hotspot
temperature, as the only variable in this function that is typically monitored and indirectly controlled
through the cooling system settings, thus plays a central role in assessment of the transformer dynamic
thermal rating. Nevertheless the impact of other ageing factors, especially moisture and oxygen, can have
a distinct impact on insulation ageing and should thus not be neglected in the thermal rating evaluations
[103].

In this chapter, an example of DTR evaluation for normal cyclic loading is presented; input data for the
evaluation is taken from the study case in the previous chapter. The DTR result is output in form of the
load factor limit (kcDTR) at which transformer can operate in given ambient conditions without exceeding
the pre-defined hotspot temperature limits.
The DTR load factor limit fundamentally depends on three variables: ambient temperature, intensity of
transformer cooling (cooling stage) and on the chosen criterion for the maximum load factor (Table 4.6).
For normal cyclic loading, hotspot temperature is the limit defining the nominal insulation ageing rate and
thereby also the transformer’s expected lifetime. The IEC 60076-7 defines the nominal ageing rate at the
hotspot temperature of Ths= 98 °C, assuming an average yearly ambient temperature of 20 °C.
To calculate the load factor limit kcDTR , a steady state loading guide hotspot temperature model in (4.14)
can be used to express the kcDTR as function of the ambient temperature (Tamb). In the equation, thermal
parameters obtained by heat-run tests or from field measurements (use of adaptive DTM approach) can be
used.

1 ∙
∙ ∙ (4.14)
1

As (4.14) is nonlinear, a solution for kcDTR (Tamb) can be obtained implicitly. A simplified explicit solution
for the kcDTR (Tamb) can then be approximated by a linear function using data fitting.

kcDTR  a  Tamb  b (4.15)

Figure 4.15 shows three solutions for kcDTR as function of the ambient temperature, each solution based on
specific DTM, i.e. the heat-run test DTM (kcDTR_heat-run), the summer period adaptive DTM (kcDTR_summer)
and winter period adaptive DTM (kcDTR_winter). The bold colored part of the kcDTR curves in the diagram
represents the range of the measured temperatures where the adaptive DTM was parametrized, whereas
the dashed line is the linearly approximated kcDTR model for extended range of ambient temperatures to
ease the comparison of kcDTR models.

94
Chapter 4: Dynamic thermal modelling

140%

135%
kcDTR - Dynamic Thermal Rating (%)

130%

125%

120%

115%

110%

105%

100%

95%

90%
-20 -10 0 10 20 30 40 50
Ambient temperature (°C)

Figure 4.15: Dynamic Thermal Rating (kcDTR) as a function of the ambient temperature, based on in-service
(top and bottom curve) and on heat-run test DTM parameters (middle curve).

Figure 4.15 demonstrates the differences between the thermal ratings, evaluated using different DTMs, as
well as the impact of ambient temperature on transformer rating. Seasonal variations of DTM parameters
(i.e. rated temperature rise values) evidently shows their impact on transformer thermal rating. When the
summer ambient temperatures are higher compared to the heat-run test conditions, the oil viscosity
decreases and heat-transfer (cooling) improves, reflecting in a ca. 5% increased thermal rating
(kcDTR_summer) compared to the heat-run test based DTM (kcDTR_heat-run). During the extreme winter
conditions (ambient temperature below -20 °C) , the adaptive DTM based thermal rating (kcDTR_winter) is
lowered by ca. 15% compared to the heat-run test conditions, but due to the lower ambient temperatures,
the effective kcDTR values at typical winter ambient temperature conditions are still 10% higher compared
to the heat-run test conditions.

4.6 Standard exponents and constants of loading guide dynamic models

4.6.1 General

Each of given loading guide models require a set of specific exponents and constants which are empirical
values. In principle, the corresponding parameters are obtained from extended heat run tests. As an
example, the IEEE extended heat run test load profile is given in Figure 4.16 [95]. However, shorter and
less time consuming profiles are also available as defined in [92] and [96].

95
Chapter 4: Dynamic thermal modelling

 
Figure 4.16: Extended heat run test.

The extended test consists of the three successive thermal runs.

1. Regular Heat Run

This part of the test is the regular heat run, i.e. the transformer is short-circuited and supplied with full-
load losses. Once the oil temperature is stabilized and corresponding oil temperature rises recorded, the
test immediately continues with the rated-load losses for one hour. The transformer is then shut down in
order to measure the warm resistance and to estimate the corresponding average winding to average oil
gradient.

2. Under-Load Run

After the warm resistance measurements for the rated current are complete, the test continues with the
supplied losses produced by 70% of the rated current plus the additional current to simulate the rated no-
load losses. When the oil temperature is stabilized and the corresponding temperature rises are recorded
the test immediately continues with supply consistent with the losses produced by 70% of the rated current
for one hour. This is followed by transformer shut down in order to measure the warm resistance and to
estimate the corresponding average winding to average oil gradient.

3. Over-Load Run

On completion of the warm resistance measurements for the 70% of rated current, the test continues with
the supplied losses produced by 125% of the rated current plus the additional current to simulate the rated
no-load losses. When the oil temperature is stabilized and the corresponding temperature rises are
recorded the test immediately continues with supply equivalent to the losses produced by 125% of the
rated current for one hour. This is followed by transformer shut down in order to measure the warm
resistance and to estimate the corresponding average winding to average oil gradient.

4.6.2 IEC 60354 and IEC 60076-7

4.6.2.1 Winding exponents "y" and "z"

The exponent y is the slope of the line on a log-log chart that best fits the plot of the average winding
temperature rise above average oil temperature measured at the end of two or three temperature-rise tests,

96
Chapter 4: Dynamic thermal modelling

versus the loading factor. This method assumes that the hotspot rise over top oil temperature is
proportional to the average winding rise over average oil temperature.

Alternatively, if the fiber-optic sensors are available for temperature measurements, then the winding
exponent z can be estimated as the slope of the line on a log-log chart that bests fits the plot of the hotspot
temperature rise above top-oil temperature measured at the end of two or three temperature-rise tests,
versus the loading factor.

4.6.2.2 Oil exponent "x"

The exponent x is obtained as the slope of the line on a log-log chart that bests fits the plot of the top-oil
temperature rise measured at the end of two or three temperature-rise tests versus the loading factor.

4.6.2.3 k11, k21, and k22 thermal constants

These constants are specific for IEC 60076-7 model. The thermal constant k11 should be estimated for the
transient top-oil rise temperature curve obtained during the heat run test period with total losses [97].

Similar to the oil time constant correction factor, k11, estimation procedure, the thermal constants k21 and
k22 for the hotspot to top-oil thermal gradient are also obtained from the part of the heat run test with
supplied total losses [97]. In general, this approach is acceptable owing to fact that the shape of the
thermal curve, f2(t), Figure 4.2, is not greatly affected by the total supplied loss level [58] compared with
when the unit is supplied with only rated load loss. Alternatively, a good technical approach would
require additional temperature tests, which would be run from the cold start until the corresponding
steady-state conditions are observed at the rated current. This procedure requires that the transformer
winding is equipped with fiber-optic sensors.

4.6.3 IEEE Annex G

4.6.3.1 Duct oil rise exponent "x"

The exponent x is obtained as the slope of the line on a log-log chart that best fits the plot of the duct-oil
temperature rise over bottom-oil temperature measured at the end of two or three temperature-rise tests
versus the loading factor. For ONAN and ONAF modes, the rated duct oil rise is the same as the top-oil
rise over bottom-oil providing steady state conditions exist, in heating cases these two values differ.

4.6.3.2 Average oil rise exponent "y"

The exponent y is obtained as the slope of the line on a log-log chart that best fits the plot of the average-
oil temperature rise measured at the end of two or three temperature-rise tests versus the loading factor.

4.6.3.3 Radiator oil rise exponent "z"

The exponent z is obtained as the slope of the line on a log-log chart that best fits the plot of the top-oil
temperature rise over bottom-oil temperature measured at the end of two or three temperature-rise tests
versus the loading factor.

97
Chapter 4: Dynamic thermal modelling

4.7 Uncertainties in Exponents Determination

Measuring temperatures may sound straightforward, as it is a familiar concept, but for accurate
temperature measurement, even for the human body, we are confronted with inconsistencies and
inaccuracies. In the case of the human body we assume that it is isothermal, but the temperature reading
differs depending on the location of the thermometer. Not only does the location matter but the human
body changes temperature throughout the day because of a number of factors e.g. ambient temperature,
cooling condition ( = clothing ) and, sometimes, medicine. Monitoring temperature changes during the
day necessitates high accuracy to distinguish between any two temperature measurements. Presenting
temperature measurements to a doctor, for instance, will only have meaning with information about the
location where they were measured, the times during the day and the body’s condition before they were
measured. These same principles apply when determining the temperature exponents of a transformer
based on 2 or even 3 different loadings. 

The exponents x, y and z are based on the difference of two temperature rise measurements. Also each
temperature difference is based on two measured values each with its own tolerance. The cumulative
effect of these measurement tolerances may lead to a large inaccuracy in the results. 

A small numerical example for calculation of the y exponent for a difference in load of 100% and 125%
will illustrate this effect:

 gradientat125% 
ln 
gradient at 100% 
y  (4.16)
 125% 
ln  
 100% 

Average winding temperature rise at 100% : ( w ) 60 +/- 1


Average oil temperature rise at 100% : ( o ) 45 +/- 1
Average winding temperature rise at 125% : ( w ) 80 +/- 1
Average oil temperature rise at 125% : ( o ) 60 +/- 1
Neglecting the measurement tolerances, one calculates

80 60
ln
60 45 1.29
ln 1.25

as a reasonable y-exponent value. Taking into account all the measurement tolerances, the value of y will
be in the interval between

81 59
ln
59 46 2.4
ln 1.25

and

79 61
ln
61 44 0.3
ln 1.25
98
Chapter 4: Dynamic thermal modelling

which is a very large inaccuracy.

Even when one includes a temperature measurement at 70% load, the accuracy will not increase, because
the tolerances (+/- 1°C in the example) will have a very large impact on the small gradient (about 9 K) at
70% load. The result: 9 +/- 2 K which is more than +/- 20%.

By performing many temperature rise tests on “identical” transformers, the average value and standard
deviation of the oil-exponent x and winding exponent y can be statistically determined. This will create a
solid basis for determining the exponents so that transformers can be loaded according the specification.
Simulation of 50 temperature rise tests including all the tolerances explain the large standard deviation in
the x- and y-exponent as determined by the temperature rise tests [2], [13], [14]. Based on the current
practice, considering the available test equipment and the available test environment at transformer
factories, a reduction of the standard deviation of the x- and y-exponent is not expected.

The exponents x, y and z can, of course, be determined during a temperature rise test of one single
transformer. Due to the large inaccuracy, there is a possible risk that the use of these exponents will result
in unacceptably high temperatures during overload. It is therefore recommended to use only the values
according to the standard in that case.

However, there is somewhat different approach to obtain the more reliable exponent values through
application of the heat transfer theory, and corresponding similarity equations as given in [3] and [43]. As
this approach would demand several heat runs, it is only recommended for prototype units.

99
Chapter 5: Direct measurements

Chapter 5: Direct measurements

5.1 Introduction

The most commonly used temperature monitoring device to measure local temperature rises in the
windings of a transformer is a fiber-optic probe. The most commonly used device to measure temperature
rises outside the winding block like in the core and structural parts is a thermocouple.

Both fiber-optic sensors and thermocouples are essentially temperature sensors. A thermocouple is an
electronic device that contains two metal wires welded together at the ends. The material properties of the
temperature monitoring junction change in electrical properties, i.e. voltage based on a change in
temperature. The measuring junction is connected to an electronic circuit which measures these changes
in resistance and converts them to a temperature reading.

Unlike a thermocouple that uses electrical transmission, fiber-optic temperature sensors work in a similar
manner to that described above, however it uses an optical fiber as a means of delivering a light signal
from the sensing device to an electronic signal processor. The sensing device changes some parameter of
the light as a function of the temperature, and the signal processor converts this change to a temperature
reading.

5.2 Fiber-optic measurements in core-type transformers

5.2.1 General

When fiber-optic probes were taken into use for hotspot temperature measurements in transformer
windings in the 1980s, endless discussions started on how many probes are needed and where to locate the
sensors in order to obtain an accurate value for the hotspot temperature rise. No relevant, all covering
answer has been found and the discussions are still going on. The natural, and also the most popular
statement has been that the location can be defined by the calculation of local loss densities and cooling
conditions.

Unfortunately, big thermal scatter will often override the theoretical predictions made by calculation (see
[15], Figure 5.1 and Figure 5.2). Thus, a certain minimum number of probes are needed to guarantee a
good quality of the hotspot temperature measurement. By “good quality” is here meant that the measured
maximum value is close enough to the real hotspot temperature for being used as a base for safe operation
of the transformer (i.e. in practice within the interval [-3,0] ºC from the hottest spot temperature).

From the values in Figure 5.1 and Figure 5.2, it is obvious that the minimum number of sensors / winding
of 2 recommended by “IEC 60076-2: Temperature rise (2011-02)” cannot guarantee a good quality of the
measurement. The hottest spot temperature rise might be missed by a significant number of Kelvin
degrees.

101
Chapter 5: Direct measurements

Top oil temperature (in tank)


= 65,8 ºC

Load current = 1 pu

Relative loss densities 156 %,


133 %, etc.

Tap position (-)

Horizontal oil circulation


from the outside to the inside
of the winding

Figure 5.1: Temperature rises above top oil temperature of the HV-winding of a 400 MVA ONAF-cooled
transformer.

Top oil temperature (in tank) =


104,8 ºC

Load current = 1,6 pu

Relative loss densities 151 %,


129 %, etc.

Tap position (±)

Horizontal oil circulation from


the outside to the inside of the
winding

Figure 5.2: Temperature rises above top oil temperature of the HV-winding of a 400 MVA ONAF-cooled
transformer.

This Section gives recommendations for delivery test purposes concerning the number and location of
sensors in windings of approximately equal length, either with axial oil circulation or with horizontal zig-
zag oil circulation obtained by oil guides / block washers. The recommendations are based on experiences
gained at regular installations and measurements since 1986.

Particular cases with non-sinusoidal load current and two axially located windings have to be studied
specifically and are not considered in the following general recommendations.

Additional to the information given in this Section, Annex C provides some examples of fiber-optic
measurements.

Another promising technique, not discussed here, uses distributed measurement along the winding by an
optical fiber inside the winding cable embedded when the cable is manufactured. It has been
implemeneted in some traction transformers made by one manufacturer [134].

102
Chapter 5: Direct measurements

5.2.2 Number of sensors

Transformers are divided into three categories based on the total leakage flux / phase at rated current. This
is estimated by the following very approximate formula [15]:

ˆ
 (5.1)
max  1.8  Z  S

where ̂ max is the maximum unidirectional flux / phase [mVs], Z is the short-circuit impedance [%] and S
is the rated power / wound limb of the main winding pair [MVA]

For auto-connected transformers Z and S refer to the values of the frame MVA and not to the MVA
transformed (i.e. S = rating plate rated power multiplied with the co-ratio (Primary Voltage – Secondary
Voltage) / Primary Voltage).

General recommendations for transformers with sinusoidal load current are the following.

 The sensors should be concentrated to one phase (preferably the phase for which the Warm
Resistance Curve will be recorded).

 For large and strategically important transformers, e.g. ̂ max  400 mVs, the number of sensors
should be 8 / winding with full rating.

 For transformers with e.g. 150  ̂ max  400 mVs, the number of sensors should be 6 / winding
with full rating.

 For transformers with e.g. ̂ max  150 mVs, the number of sensors should be 4 / winding with full
rating.

 It is normally not of interest to install sensors in regulating windings, tertiary windings or in


circulating oil. An exception here might be a need to check that the regulating winding is properly
dimensioned.

 The sensors should be installed on one of the sides of the active part, approximately
perpendicularly to the limb.

There is a belief that OD-cooled windings show a lower thermal scatter than illustrated in Figure 5.1 and
Figure 5.2, i.e. that it would be enough to install a lower number of sensors in such windings than in ON-
and OF-cooled windings. A short study of results obtained on OD-cooled windings was made. An
example from this study is the OD-cooled Smit HV-windings of a 375 MVA generator step up
transformer where 8 installed sensors showed a range of variation of 6.5 K [2], but an uncertainty between
discrete sensors up to 9 K could be derived. Thus, it seems reasonable to recommend the same number of
sensors for all cooling modes.

103
Chapter 5: Direct measurements

5.2.3 Location of sensors

5.2.3.1 Transformers with windings without tapping discs

For an axially cooled inner winding, the sensors should be concentrated to the corner with the highest
eddy losses, and ditto for an outer winding (see Figure 5.3).

Figure 5.3: Location of fiber-optic sensors in a winding with 1…4 internal cooling ducts.

For a zig-zag cooled winding with the oil circulation directed from the main gap, the sensors should be
concentrated to the corners where a) the highest eddy current losses occur, b) the horizontally circulating
oil has accumulated maximum heat (see Figure 5.4a).

For zig-zag cooled windings with the oil circulation directed towards the main gap, the temperature
distribution of the oil and the eddy current loss distribution counteract each other. In these cases a
compromise could be to locate the sensors at the center of the disc (see Figure 5.4b).

104
Chapter 5: Direct measurements

Figure 5.4: Location of fiber-optic sensors, in a zig-zag cooled winding pair, where circulating oil is
accumulating heat towards the corner with a) maximum eddy loss, b) minimum eddy loss (the zig-zag oil
circulation pattern has been obtained by oil guides / block washers).

For large power transformers the 8 sensors / winding with full rating could be located as two sets
according to Figure 5.4 at two different locations around the circumference. For mid-size units the 6
sensors / winding with full rating could be located as one set according to Figure 5.4, and sensors below
disc nr 2 and 3 from the top at some other location around the circumference. For small units the 4 sensors
/ winding with full rating could be located as one set according to Figure 5.4.

A specific study on the radial location of the sensors has been made at Stuttgart University. They have
found that for all cooling modes, the radial position of the sensor is very important because significant
radial gradient differences have been observed [122].

5.2.3.2 Double-Concentric windings without tapping discs

For double-concentric winding arrangements with about 50 % of the LV-winding located on each side of
the HV-winding, there is a strong eddy current loss concentration in the centre of the top disc of the HV-
winding. Thus, the sensors should be located in the HV-winding according to Figure 5.4b, irrespective of
whether the oil circulation is zig-zag or axial. For the LV-winding it should be enough to install sensors in
the other one of the two LV-shells. For the LV-shells, the general principles shown in Figure 5.3 and
Figure 5.4 can be followed.

5.2.3.3 Windings with tapping discs

In a design with tapping discs in the outer winding close to the winding ends, the hottest spot will be
located in one of the top discs in the (+) position, and in the first disc above the current less tapping discs
in the (-) position. For example, in an OFAF-cooled unit with a 300 mm winding segment at the top
followed by a 200 mm tapping section (see Figure 5.5) a gradient of 28,4 K above top oil temperature (in
tank) was measured in the top disc in the (+) position. In the (-) position, the hotspot-to-top-oil gradient
was 49,7 K in the first disc above the tapping section, whereas the sensor showing 28,4 K in the top disc
in the (+) position increased to 32,0 in the (-) position. Thus, the disc above the tapping discs was 17,7 K
hotter than the top disc in the (-) position, the load current in all cases being 1,0 pu at respective tap
position.

105
Chapter 5: Direct measurements

Figure 5.5: Leakage flux pattern at the top of an outer winding with tapping discs, in the (-) position.

In a transformer with 8 sensors in the outer winding, the sensors can be distributed between the top discs
and the discs around the tapping discs according to Figure 5.6a.

In a transformer with 6 sensors in the outer winding, the sensors can be located according to Figure 5.6a at
one location around the circumference.

In a transformer with 4 sensors in the outer winding, the sensors can be located according to Figure 5.6b.

The eddy current losses in the discs around the tapping section vary with the size of the gap and the axial
height of the conductor strands. Thus, an individual calculation should always be done and sensors located
in these discs whenever their losses exceed the losses in the top disc. The calculated losses (DC + Eddy) in
the case shown in Figure 5.5 were as follows: top disc loss = 1589 W; loss in the first disc above the
tapping section = 2466 W; loss in the first disc below the tapping section = 1753 W; average disc loss of
the whole winding shell = 1118 W.

106
Chapter 5: Direct measurements

Figure 5.6: Location of sensors in an outer winding with a tapping section and with oil circulation either from
the inside or from the outside of a winding. a) location of 8 sensors: the encircled sensors can be at 2 locations
around the circumference; location of 6 sensors: the encircled sensors are not doubled, b) location of 4
sensors.

5.2.4 Installation techniques

5.2.4.1 General

A critical parameter at the installation of fiber-optic probes is the permissible bending radius. It should be
noted that each probe manufacturer specifies his own permissible bending radius. One of the leading
probe manufacturers specifies the following 3 different minimum bending radii.

107
Chapter 5: Direct measurements

 Short term bending radius: this is the minimum radius that must be maintained before the fiber
will break. It is approximately 3 mm.

 Long term bending radius: this is the minimum radius that must be maintained before the fiber
may break after many months of being in that position. This radius is approximately 10 mm.

 Minimum bending radius to avoid light attenuation: this is the minimum radius that must be
maintained to avoid light losses inside the fiber. This radius is approximately 130 to 150 mm.

It is very important that the last 10 mm of the probe containing the sensor is not bent, pressed or glued.

The probes should be identified (= numbered) during the installation using a material able to withstand the
drying process of the transformer.

Due to the difficulty of replacing a damaged probe once the transformer is finished, it is recommended to
check the probes at several stages of the manufacturing:

 After finishing the installation of the probe.

 Once the whole phase has been assembled and the final compacting force has been applied.

 After finishing the cleats and leads and the routing of the probes to the tank wall feedthrough
plate.

5.2.4.2 Windings with radial spacers

The probe is installed in a slot in the radial spacer according to Figure 5.7. A suitable dimension of the slot
is 2x2 mm. If the spacer thickness is 2 mm or less, then the slot can be made by a band saw, and crepe
paper can be glued to the other side of the spacer.

Figure 5.7: Examples of installations in spacers of different thickness, and at different locations. The sensor
(black dot) is located about 2 - 3 mm from the end of the probe.

There are several techniques to keep the sensor in position. One is to use (transformer approved) glue at
such a distance from the sensor that the last 10 mm of the probe is not pressed by glue. Another one is to
glue a crepe paper around the spacer as shown in Figure 5.8. There a very thin paper is glued over each
one of the probes, about 10 mm from the sensor.

108
Chapter 5: Direct measurements

Figure 5.8: Installation of sensors 1 and 4 in a 120 kV winding of a 400 MVA transformer using a spacer with
2.5 mm thickness.

The spacer equipped with the probes can be inserted into the winding, when the winding has been fitted on
the limb. In order to make it possible, the tabs shall be cut as shown in Figure 5.7. The spacer should be of
the same thickness as the one it will replace in the winding.

Theoretically it is obvious that the sensor should be in contact with the bare conductor metal. In reality it
is difficult to obtain this. The error due to the location in the spacer should be less than 2 K in common
designs [16] and [17]. Previously published CFD calculations have also demonstrated that the insulation
paper that is underneath the spacer is essentially at the same temperature as the copper conductors inside
the discs (Figure 5.9) [114].

Figure 5.9: Temperature contours (from 3D-CFD) for a core-type transformer winding with radial spacers.

5.2.4.3 Layer windings made of flat or CTC conductors

Firstly, the layer to be measured is selected. When the layer is ready on the winding machine, pressboard
kits with slots for the probe are inserted. Figure 5.10 shows the pressboard kit for probe no. 9 in the HV-
winding of a 3,15 MVA distribution transformer. A layer of insulation material is fitted on the top of the
winding. This layer prevents horizontal oil circulation along the top of the winding. On top of this, the top
conductor carries the highest losses. Thus, it was expected that the hottest spot would be located in this
conductor. The slot was made deep enough to permit the sensor to come down to the mid-height of the
conductor. When the winding had been fitted on the limb, the sensors were inserted into the slots.

In case of larger transformers and CTC-conductors, the same technique as for Flat Conductors can be
used. In this case the hotspot is not expected to be located in the top cable, because the top of the top cable
is exposed to free oil circulation. A separate calculation has always to be done for such cases. Obviously,

109
Chapter 5: Direct measurements

the hotspot is embedded below cable nr 2, 3 or 4 from the top of such a winding. In this case the slot has
to be extended deeply enough to reach the hotspot location.

An attempt was made to insert extra radial spacers with embedded sensors in a layer winding (Figure
5.11). The measurement results obtained yielded a hotspot-to-top-oil gradient of 12,7 K, whereas the
average winding-to-average oil gradient g was 14,1 K. This yields a hotspot factor H = 0.9  1,0, which
indicates that the extra radial spacers permitted a local oil circulation yielding wrong measurement results.

Figure 5.10: Insulation kit between the layers for the installation of probe no. 9 to measure the uppermost
conductor of the winding. The top ring with the same colour as the conductors is pressboard insulation.

Figure 5.11: Installation of fiber-optic sensors in a layer winding by means of extra radial spacers. An
example of how the sensors should not be installed!

5.2.4.4 Foil windings

The technique to install sensors in foil windings is the same as for layer windings made of flat conductors
(see Figure 5.10 and Figure 5.12).

In the same way as for layer windings made of Flat Conductors, the sensors are inserted into the slots,
when the winding has been fitted on the limb (Figure 5.13).

110
Chapter 5: Direct measurements

Figure 5.12: Insulation kit between layers for the installation of sensors no. 3 and 4 at a distance of 8 mm from
the top of the LV-winding of a 3.15 MVA distribution transformer.

Figure 5.13: The sensors have been inserted to the bottom of the slots. After that the probes are locked by
glued tape.

5.2.4.5 Windings without radial spacers but with axial spacers

In essence the same installation principles apply for the assembly of the sensor in an axial spacer as in the
case of radial spacer (see Figure 5.7 and Figure 5.14), although the thickness of the axial spacer is always
larger than 3 mm. The axial spacer is re-inserted into the winding and the sensor lead should not block the
oil flow coming out from the axial cooling duct. In that case the cooling will be reduced and the hotspot
temperature will increase as measured by the fiber-optic sensor [18].

111
Chapter 5: Direct measurements

a) b)

c) d)

Figure 5.14: a) fiber-optic sensor in an axial spacer, b) inserting the spacer into the selected cooling duct, c)
spacers fixed in cooling ducts, d) leads are taken out between two parts of the electrostatic shield to prevent
the sensor leads from blocking the oil flow.

5.3 Fiber-optic measurements in shell-type transformers

5.3.1 General

This Section describes the location of fiber-optic sensors to measure the hotspot temperature rises in shell
type transformers and the methodology to install these sensors depending on their location [19]. This
section will not deal with the detail handling of the probes, because it was dealt with in Section 5.2.4
above.

5.3.2 Number of sensors

For shell type transformers the recommendation given by IEC 60076-2, Temperature rise, Edition 3, 2011-
02, is followed. It means that sensors are installed in each of the windings with full rating (usually LV-
and HV-winding).

112
Chapter 5: Direct measurements

5.3.3 Location of sensors

5.3.3.1 General

In shell type transformers there are two different locations for hotspots in the windings: coil edges and
connections between coils. Fiber-optic probes should be installed at least in the location with the higher
calculated hotspot temperature rise.

5.3.3.2 Coil edges

Due to the winding construction of shell type transformers, the leakage flux through the coils that
contributes to load losses is higher at the internal and external edges (first and last turns) of the coils and
especially in those adjacent to high-to-low voltage spaces. The accuracy of the location of the hotspot will
depend on the accuracy when calculating the leakage flux density distribution over these domains (Figure
5.15).

Figure 5.15: Winding section of the coils showing the magnetic flux vectors and the high-to-low space.

The magnetic field together with the current through the coil makes it possible to calculate load losses
(DC, eddy and circulating current) for each one of the layers of the coil. In a shell type transformer the
turns can be made of one or more layers of conductors and the losses should be calculated for each layer
(Figure 5.16).

113
Chapter 5: Direct measurements

Figure 5.16: Simplified representation of the outer edge of a coil with three layers per turn.

Load losses create a different temperature gradient in each layer at the edge of the coil. This gradient will
be the same along its whole perimeter for each layer. Due to the specific construction of shell type
transformers the oil goes through the coil increasing its temperature from the bottom to the top, i.e. the
hottest domain of the coil will be the top side, usually around the vertical axis of symmetry (see Figure
5.17).

The final location of the hotspot, and therefore of the sensor, will be the outer edge or the inner edge
layers depending on the copper-to-oil gradient of each layer and the oil temperature distribution in this top
domain. The distribution depends on several factors like the oil speed, the spacers’ pattern and the
insulation structure at that domain.

114
Chapter 5: Direct measurements

Figure 5.17: Schematic representation of the oil flow through a shell type coil.

5.3.3.3 Connections between coils

The connections between coils are usually performed by a brazing at the inside side or the outside part of
the coil. This means that the braze will receive the magnetic field of the coil edge and because this brazing
is a higher resistance point, the temperature gradient of this point will be high so it is also a candidate for
containing the hotspot of the transformer.

As explained in the previous Section, the top domain of the coil is the one with higher oil temperatures.
Therefore the maximum absolute temperature values for connections will be most probably the ones with
brazings performed at the top.

5.3.4 Installation techniques

5.3.4.1 Coil edges

It is recommended to install the probe between the two layers with higher calculated temperatures at the
vertical symmetry axis of the coil (Figure 5.18).

115
Chapter 5: Direct measurements

For that purpose, modified pressboard filler is used with the same height as the coil and thickness slightly
higher than the probe (Figure 5.19).

The probe is located in the slot of the filler insulated with crepe paper making sure that the glue does not
contact the tip of the probe. The purpose of this paper wrapping is to simulate the real conditions of the
copper in the layer that has also its own insulation and to prevent oil from circulating inside the slot which
could yield a too low value. Also some extra crepe paper should be applied and glued around the filler to
prevent the probe from sliding out from the slot.

Figure 5.18: Representation of the layers of the coil edges with a hypothetical location of the sensor.

a)

116
Chapter 5: Direct measurements

b)

Figure 5.19: a) special filler to install the probe between layers, b) installation of the sensor inside the filler.

Finally the two selected layers are carefully separated and the filler is inserted between them making sure
that the height does not exceed the one of the coil (Figure 5.20). As mentioned before, it is important to
verify that none of the spacers of the washer over the coil will press the fiber-optic probe in order to avoid
breaking.

Figure 5.20: Insertion of the filler between layers at the coil edge.

5.3.4.2 Connections between coils

In order to measure the real temperature of the hotspot it is recommended to install the fiber-optic sensor
in direct contact with the brazing. First, one or two layers of crepe paper are applied and glued to keep the
sensor in position without covering the sensor (Figure 5.21).

After that both the cable and the probe are insulated with the required number of layers of paper for that
particular connection between coils (Figure 5.22).

117
Chapter 5: Direct measurements

Figure 5.21: Fiber-optic sensor in direct contact Figure 5.22: Final appearance after insulating
with the brazing. the connection.

5.4 Measurement of the duct oil temperature next to the hotspot

In order to measure the local hotspot gradient (copper hotspot absolute temperature minus absolute
temperature of the oil surrounding the hotspot) it is recommended to install a probe to measure the local
oil temperature. For that purpose a modified spacer can be used (Figure 5.23).

The spacer will protect the probe from the pressure. In order to fit tightly the probe in the spacer some
crepe paper can be applied taking care that the tip of the probe is free from insulation and glue so that the
probe will be in direct contact with the oil.

The location of the modified spacer will be a coil washer or an insulating channel depending on whether
the location of the copper hotspot probe is in a coil edge or in a connection between coils (Figure 5.24).

Figure 5.23: Installation of the probe in a modified spacer, and final location in the coil washer.

118
Chapter 5: Direct measurements

Figure 5.24: Finished installation of probes in the coil edge and in a spacer next to it.

5.5 Routing of the fiber-optic probes

5.5.1 Through the phase insulation

After installing the fiber-optic probe it is necessary to take out the probe through the insulation structure of
the transformer. For that purpose some adapted spacers are recommended. Also it is very important to
avoid cuts or modifications in the channels and angles that could jeopardize the dielectric strength of the
structure (Figure 5.25).

Figure 5.25: Exit of the probes through the insulation.

5.5.2 Through the cleats and leads

Once the tank and the cover have been assembled it is necessary to take the fiber-optic probes up to the
tank wall feedthrough plate. This plate is usually located in one of the side walls of the tank, so the probes
located at different points of the phase have to be routed properly across the transformer cleats and leads.
For this reason the probes should be protected to avoid any accidental damage while working inside the
transformer. Some routing using pressboard tubes is recommended, always respecting the minimum
bending radius mentioned in Section 5.2.4.1. The extra length of a probe must be adequately supported
(Figure 5.26).

119
Chapter 5: Direct measurements

Finally and despite the fact that the fiber-optic probes do not have any metallic part inside it is
recommended to avoid electrically stressed domains like the high voltage exits and also to follow
equipotential lines to prevent from creepage along the probe.

Figure 5.26: Pressboard tubes to protect the probes and connection to tank wall feedthrough plate.

5.6 Measurement of local temperatures in core and structural parts by


thermocouples

5.6.1 General

External transformer tank temperatures are fairly well known thanks to the extensive use of
thermocameras. A more tricky case is the temperature rises of internal, not visible parts like the core, the
yoke clamps and the flitch-plates. Especially, in auto-connected transformers with high short-circuit
impedance these temperature rises can – if not well controlled – cause extra gassings and shorten the life
length of the transformer. Although permissible temperature limits are not specified in the leading
transformer product standards (IEC and IEEE), it is important that the manufacturers of transformers with
high leakage flux make their own measurements and quality checks. Gas-in-oil analysis is a helpful tool,
but not a complete guarantee that the transformer is free from internal hotspots.

This Section gives hints for the installation of thermocouples to measure local temperature rises in the
core and structural parts of selected prototype and project transformers. The recommendations are based
on experiences gained at installations since the 1970s.

It is assumed that this type of measurements occurs rarely, and that it is replaced by thermal modelling and
calculations for the main stream of transformers manufactured.

5.6.2 Number and location of thermocouples

The number of thermocouples is not as critical as the number of fiber-optic probes, because
thermocouples are cheap and robust to install. There are mainly two aspects that will limit the number of
thermocouples: 1) the effort to install them and to arrange routings to the thermocouple bushings in such a
way that they do not injure the voltage strength, 2) the number of thermocouple bushings and available
bushing leads.

It is very difficult to give general recommendations for the location of thermocouples, since local
temperatures are highly dependent on local loss densities, local oil circulations, leakage flux control
120
Chapter 5: Direct measurements

methods etc. These characteristics vary much more between different design concepts than e.g. the
winding designs, for which some recommendations for the installation of fiber-optic sensors could be
given. Thus, the hints given in this Section are given by some selected examples from real practice.

5.6.3 Location of Thermocouples in Yoke Clamps

The thermal behaviour of a yoke clamp is illustrated by the temperature rise distribution of a flat,
unshielded magnetic steel clamp without extending magnetic details. The transformer in question is a 3-
phase 175 MVA, 330 kV auto-connected transformer with a 3-limb core.

Figure 5.27 shows the installation of thermocouples opposite to phase B of the 175 MVA transformer
[29]. Eleven thermocouples were installed on the top of the clamp, 6 on the outside of the clamp opposite
to the centre-line of phase B, and 7 on the inside of the clamp at the end of the core (not shown in Figure
5.27). The winding support pipes were made of non-magnetic steel and earlier experiences indicated that
there was no reason to install thermocouples in them.

The reason for the installation of thermocouples in this transformer was that the order comprised 3
identical units and that some experiments to guide leakage flux were made on two of the units. Thus, the
unit shown here acted as the comparison base representing a “normal” design.

Figure 5.27: Permanent installation of thermocouples opposite to phase B.

An overload temperature rise test at 1,40 pu was made. The measured temperature rises above the bottom
oil temperature are shown in Figure 5.28.

121
Chapter 5: Direct measurements

Figure 5.28: Local temperature rises above bottom oil temperature at the load current 1.4 pu.

Some general hints for the location of thermocouples based on this test and on similar tests on other
designs could be as follows.

 Extending parts made of magnetic steel will attract leakage flux and cause high loss
concentrations at locations where flux gets its best chance to “jump” over to the core. Typical
extending parts are winding support feet and supports for cleats and leads made of magnetic steel.

 The temperature rise distribution in the clamp in Figure 5.28 can be considered as typical for flat
magnetic steel clamps without extending magnetic details and located in a transformer with
wound outer limbs. In this case the yoke clamp extensions at the ends of the core will act as an
“extending magnetic part” attracting flux into the core. Thus, a local temperature maximum (34,6
K) occurs at the location where flux enters the core.

 Box clamps made of magnetic steel are considerably hotter than flat clamps, because the
horizontal surface of the clamp towards the winding block will represent an “extending part”
attracting flux into the core.

5.6.4 Location of thermocouples in flitch-plates and outer core packets

The thermal behaviour of flitch-plates and outer core packets is illustrated by flat flitch-plates made of
magnetic steel, and by sub-divided outer core packets. The transformer is a 3-phase 400/400/125 MVA,
410/120/21 kV transformer equipped with fiber-optic probes (see Figure 5.1 and Figure 5.2 above) and
thermocouples in the core and structural parts [30], [31]. Figure 5.29 shows the installation of
thermocouples in flitch-plates and outer core packets of phase B of the unit.

122
Chapter 5: Direct measurements

Figure 5.29: Permanent installation of thermocouples in flitch-plates and outer core packets opposite to the
top of the winding block of phase B.

Several temperature rise tests were made on the transformer. One of the tests was a 15 h test at the
constant load current 1.6 pu. The local temperature rises in steady state are shown in Figure 5.30.

Figure 5.30: Temperature rises above adjacent oil temperature 0.75 x TO + 0.25 x BO in steady state at the
load current 1.6 pu.

The relevant adjacent oil temperature at natural oil circulation appeared to be the average of the top oil
temperature and the average oil temperature in the tank, i.e. 0.75 x TO + 0.25 x BO, where TO = top oil
temperature (in tank) and BO = bottom oil temperature (in tank). The low difference between this oil
temperature and the two “Reference” thermocouples attached to core packets without stray losses
(showing -0.3 K and -1.2 K) shows that this oil is more relevant as adjacent oil than the top oil.

The steady state oil temperatures at the end of the 15 h test were TO = 104.8 ºC and BO = 63.4 ºC (the
cooling mode was ONAF). It means that the maximum temperature rises above top oil temperature due to
the radial leakage flux were 18.1 K for the flitch-plates and 27.4 K for the outer core packets.

Especially, in the case of outer core packets also the contribution from the main flux has to be taken into
account. An 8 h thermal no-load test was made at the supply voltage 110 %. The temperature rises are
shown in Figure 5.31.
123
Chapter 5: Direct measurements

Figure 5.31: Temperature rises above the adjacent oil temperature 0.75 x TO + 0.25 x BO in steady state at
the supply voltage 110 %.

From the values in Figure 5.31 it is obvious that also the flitch-plates show a certain temperature rise due
to the main flux, i.e. they act as a parallel path to the magnetic circuit.

The steady state oil temperatures at the end of the 8 h no-load test were TO = 57.8 ºC and BO = 43.0 ºC
(only two radiators were kept open). It means that the maximum temperature rises above top oil
temperature due to the main flux were 5.6 K for the flitch-plates, 11.3 K for the hottest outer core packet,
and 10.1 K for the outer core packet with the maximum temperature rise due to leakage flux.

Thus, it is expected that the temperature rises above top oil temperature in service are at 1.6 pu load
current and 110 % supply voltage:

 Flitch-plates   = 18.1 + 5.6 = 23.7 K.

 Outer core packets   = 27.4 + 10.1 = 37.5 K.

The general recommendations based on this test and on similar tests on other designs could be the
following.

 The thermocouples should be located opposite to the top of the winding block.

 The thermocouples in the sub-packets should be located close to the edge(s) of a sub-packet. The
thermocouples in the sub-packets in Figure 5.29 to Figure 5.31 are inserted between the core steel
laminations no. 2 and 3, seen from the outside of the core. They were located either 10 mm from
the edges of the sub-packet or in the centre of the sub-packet.

 From the wide range of variation of temperature rises in Figure 5.30 it is obvious that the
installation of 1 or 2 thermocouples might not give the right picture of the temperature rises.
Preferably, a minimum number of 4 thermocouples should be installed both in the flitch-plates
and in the core packets, i.e. totally 8 thermocouples if both parts are to be measured.

124
Chapter 5: Direct measurements

5.6.5 Location of thermocouples in the top yoke

The thermal behaviour of the top yoke is illustrated by the top yoke of the 400 MVA transformer
presented above. Figure 5.32 shows a cross-section of the top yoke with 3 internal cooling ducts and the
local temperature rises both inside one of the yoke sections and cooling ducts [32].

Figure 5.32: Temperature rises above top oil temperature at the end of an 8 h thermal no-load test at 110 %
supply voltage.

The three left hand thermocouples were installed in the centre of a core section between two adjacent
cooling ducts, and the three right hand thermocouples in the cooling duct, on the surface of the core
section (for investigation purposes). The temperature rises measured in the centre of the core section
showed a variation from 23.8 K to 31.8 K, and the question remains: “Did the measurement yield the
hottest spot of the core?” Previous measurements made on similar units indicated that the value 31.8 K
could be considered to represent the hotspot temperature rise of the core.

The general recommendations based on this test and on similar tests on other designs could be:

 In order to measure the internal hotspot temperature rise in a core, the thermocouples should be
installed at the top of the centre phase core steel laminations in a 3-phase transformer, and ditto
for a wound limb in a single-phase transformer.

 The thermocouples should be installed at the joint between the yoke and limb laminations.

 The thermocouples should be installed as “deeply” as possible in the core section to be measured,
i.e. as far as possible from circulating oil.

 Minimum 4 thermocouples should be installed at the expected hotspot location, if previous


measurement results obtained on a similar unit are not available.

5.6.6 Core temperature

A certain number of customers’ technical specifications include the following requirement:

 “Thermal sensors shall be installed to measure the temperature rise of the magnetic circuit.”
125
Chapter 5: Direct measurements

It does not exist a unique hotspot of the magnetic circuit to be measured. The Sections above have pointed
out two different hotspots in the core:

 Core surface hotspot opposite to the top of the winding block. The temperature rise of this hotspot
depends on:

o The leakage flux density towards the limb, i.e. the short-circuit impedance of the loading
case, the loading current and the geometrical configuration of the active part.

o The number of slots in the outer core packets.

o The oil circulation around the flitch-plates and outer core packets.

o The rated flux density of the core.

o The core type (e.g. 3-limb or 5-limb core).

 Internal core hotspot in the top yoke. The temperature rise of this hotspot depends mainly on:

o The rated flux density of the core.

o The number of cooling ducts in the core.

o The core type (e.g. 3-limb or 5-limb core).

The 400 MVA example above shows that the hotspot of the magnetic circuit at the load current 1.6 pu was
located in the outer core packets (37.5 K above the top oil in the tank). At a lower load current, and at the
same supply voltage (110 %), the hotspot would be located in the top yoke (31.8 K above the top oil in the
tank).

5.6.7 Installation technique of thermocouples

Figure 5.33 shows a thermocouple installed in a steel part.

126
Chapter 5: Direct measurements

Figure 5.33: A hole with the diameter of 2 mm is drilled in the steel part to a depth of about 7 mm. A wooden
piece is pushed into the hole together with glue to keep the sensing element in place and to keep the
thermocouple earthed during the lifetime of the transformer.

The measuring point of the sensing element is the point where the two conductors are galvanically
connected together for the first time, seen from the insulated thermocouple leads. This point is usually
located at the steel surface. If deeper parts of the steel body are hotter than the surface, the Cu/Konstantan
leads will by conduction transfer this heat to the measuring point.

The thermocouples in the core have usually been installed between the core steel laminations at the core
stacking process. The thermocouples on the core surface e.g. in Figure 5.31 were installed as follows:

 The thermocouples at the edges of the sub-packets showing the values 14.7 K; 14.8 K; 14.8 K;
13.8 K were inserted between sheet no. 2 and 3, about 10 mm from the edge.

 The thermocouples in the centre of the sub-packets were inserted at the centre line of the sub-
packets between sheet no. 2 and 3.

 The Reference thermocouple showing 16.5 K was inserted about 10 mm from the edge, between
packet no. 7 and 8.

 The Reference thermocouple showing 14.7 K was glued on the surface at the centre of step no. 7.

Figure 5.34 shows an example of the installation of thermocouples in the top yoke of a 3-phase
68.5 MVA, 220 kV transformer. In this case thermocouple no. 7, just at the top of phase B, showed the
highest value.

127
Chapter 5: Direct measurements

Figure 5.34: Thermocouples installed in the “deepness” of the core section on the left hand side of the cooling
duct, at stacking of the top yoke. The elements 7, 9 and 10 were located along the centre line of phase B. The
elements 2, 4 and 5 were located at the centre line between the phases, for comparison.

The thermocouples inside a transformer are often routed through a significant leakage flux. Thus, a
thermocouple pair including Fe should be avoided. A pair proven to perform well in high leakage fluxes is
Cu/Konstantan. Moreover, the thermocouple leads shall always be kept tightly together and twisted in
order to avoid extra induced voltages between the thermocouple partners. The thermocouple leads
Cu/Konstantan are fairly stiff, but good twisting results have been obtained by a hole drilling machine
tuned to low revolving speed.

The leads are routed to the bushing leads in such a way that they do not injure the dielectric strength of the
transformer. Figure 5.29 shows how thermocouples from the bottom yoke clamp and the flitch-plates are
collected in a pressboard pipe on the other side of the flitch-plates to the top of the transformer. Also core
step no. 1 is used, as well as step no. 7 for the Reference thermocouples. All these leads remain inside the
earthed cylinder around the limb.

It is important that the sensing element remains in galvanic contact with the steel part or the core, i.e. that
it remains earthed and not becomes a body with floating potential during the lifetime of the transformer.

The measurement device shall be selected by assuming that the sensing element of the thermocouple is
earthed. Some thermocouple reading devices for multiple thermocouples have internally or externally one
side of the thermocouples connected. In this way closed loops are created and voltages are being built up
between different thermocouple pairs. Such devices shall be avoided and replaced by complete differential
measurement devices.

128
Chapter 5: Direct measurements

Standard current transformer bushings can be used as thermocouple bushings. A current transformer
bushing comprises 9 leads, i.e. one bushing can take 4 thermocouples, two bushings 9 etc. The Cu-bushing
leads will not mix up the measurement results although connected to Konstantan in the top oil temperature
below the cover. This has been verified by separate tests and calibrations.

5.6.8 An experience with fiber-optic installation in the yoke clamps

5.6.8.1 Introduction

In some cases fiber-optic probes are used also for temperature measurements in the core and structural
parts. Some of the probe manufacturers have even developed insulation kits for these applications. This
Section describes and compares two types of fiber-optic sensor mounting on a yoke clamp: surface
mounted and hole mounted. The difference between these installations is the placement on the surface of a
metal part versus inside a hole drilled in the metal part [33].

Three transformers (commercial designs) of identical electrical characteristics were chosen for this
project. The aim was to install probes in one of the bottom yoke clamps at the centre limb so as to measure
temperature rises. Sixteen probes were installed in each one of the three transformers.

5.6.8.2 Location of the sensors

The 16 probes were installed in the bottom yoke clamp on the LV-side of the active part in the 3 units.
The probe installations were divided equally between 8 surface and 8 hole mounted probes. The location
of the 16 sensors are shown in Figure 5.35. Probes nr 1 to 5 (and nr 12 to 16) were located along the
length and nr 8 to 5 (and nr 9 to 12) over the height of the clamp. The distances from the centre line of
Phase B and the probes 1; 2; 3; 4; 5 (and 16; 15; 14; 13; 12) were 466; 386; 306; 226; 100 mm,
respectively. The distances from the probes 8; 7; 6; 5 (and 9; 10; 11; 12) to the top of the clamp were 360;
260; 140; 20 mm, respectively.

Figure 5.35: Installation of 16 fiber-optic probes in the LV bottom yoke clamp.

5.6.8.3 Installation technique

In the installation of the probes, the important part was to create a robust supporting mechanism that could
prevent movement of the probes during the manufacturing process of the transformer. A special spacer of
3 mm thickness was designed with a 3 mm slot at the center for both hole and surface installation. This
slot was not cut right through, to fit the probe tightly. On the surface mounting, the tip of the probe was
inserted into the slot to prevent exposure to oil circulation. The hole installation was slightly different, the
129
Chapter 5: Direct measurements

probe was taped with insulation paper to ensure that there was no oil exposure. The probe ran through a
spacer and was inserted into the clamp hole with tight insulation closer to the tip. Figure 5.36 and Figure
5.37 summarize the procedure explained. The figures also expose the sensitivity of the supporting
mechanism.

Figure 5.36: Surface installation of fiber-optic probes.

Figure 5.37: Hole installation of fiber-optic probes.

5.6.8.4 Results and conclusion

To determine the preferred method of installation, the temperature rise results obtained during the heat run
test are shown in Figure 5.38 below. The temperatures were recorded during full heat run test, however
presented in Figure 5.38 are the results measured during the last hour, i.e. they are steady state values
corresponding to rated current.

130
Chapter 5: Direct measurements

Figure 5.38: Steady state temperature rise above bottom oil temperature (in tank) at rated current.

The distribution of temperatures along the length and over the height is not uniform. On average the
temperature difference between hole and surface measured results show that surface values and hole
values differ by only 3.0 %; 2.4 % and 0.7 % for unit 1, unit 2 and unit 3, respectively. This means that
any of the methods can be safely chosen with low error.

5.7 Measurement of top-oil temperature with magnetically mounted sensor

Measurement of top oil temperature is an essential step in the determination of the winding hottest spot
temperature. For new transformers, submitted to the temperature rise test, the conventional method is to
provide a resistive thermal detector (RTD) immersed in oil. However, it is recognized that the oil
temperature under the tank cover is not uniform and any top oil temperature measurement is intended to
be only “representative of top-liquid in the cooling flow stream” (IEC 60076-2). In fact, IEC recommends
2 separate oil pockets for transformers above 20 MVA and 3 pockets for transformers above 100 MVA. If
more than one pocket is used, the reading of the sensors should be averaged in order to obtain a single
value representative of the oil temperature at the top of the tank. It is also recognized by the industry that
measurement of temperature on the header leading from the tanks to the radiators (or coolers) also
provides a “representative” value of the top oil temperature.

With the deployment of transformer on-line monitoring, it is becoming frequent to measure the top oil
temperature with a temperature sensor magnetically held on the upper part of the tank wall (mag-mount
sensor) [34]. This is a common practice for transformer monitoring as it is not usually possible to have
access to the RTD that was used as a reference during the temperature rise test and that is still used in
service to feed the top oil temperature indicator and also the transformer protection system. It is
recognized that the value measured with the mag-mount sensor can differ from the one measured with
sensors immersed in the oil well. The temperature difference between the two locations should be
acknowledged and some correction needs to be applied.

131
Chapter 5: Direct measurements

In the application of magnetically held temperature sensor it is essential to select a location where the oil
flow is adequate. The proximity of the header or the header itself is the best practice. Corners of the tank
should be avoided as the oil flow is often reduced in these locations.

The difference between a magnetically held temperature sensor and RTD used as a reference during the
heat run test should be determined experimentally. Field experience indicates that the temperature
difference tends to increase at reduced load or lower ambient temperature. The value of most interest is
the one observed at full load as this is the critical condition for winding hotspot temperature assessment.

The best positioning to minimize the temperature difference is to locate the magnetically held temperature
sensor close to the RTD immersed in the oil pocket whether it is on the tank wall or on the top of the tank.
In this case the difference is likely to be less than 1.5 K.

If the RTD sensor is on the top of the tank and the magnetically held temperature sensor has to be
mounted on the tank wall, the south side should be avoided as the effect of sunshine can lead to error of up
to +6 or +7 K in the afternoon.

The temperature difference between magnetically held temperature sensor and RTD in oil pocket should
be characterized under high loading conditions. Measurements should be carried out over several days as
the effect of sunshine, wind and rain can induce temporary excursion that are not representative of the
average value.

The average temperature difference should be used to correct the measured value in order to simulate
properly the reference top oil temperature that was determined during the temperature rise test. The best
practice is to use a fixed correction value, independently of oil temperature.

132
Chapter 6: Shell-type transformer thermal modelling

Chapter 6: Shell-type transformer thermal modelling

6.1 Introduction

Shell-type is widely spread in the US and it is predominant in other countries: all the nuclear fleet in
Belgium, half of the nuclear fleet in France, more than 85% 400 kV network transformers in Spain, are a
couple of examples. It shows the significant place of shell-type on the high power transformer market and
obviously this technology has to be depicted in this CIGRE document.

This Section of the Brochure shows an overview of the shell-type transformer thermal modelling
techniques while highlighting other important thermal-related aspects.

The steady-state temperature field inside shell-type is built up from an interaction of the electromagnetic
field and the oil velocity field. As a result, the electromagnetic field is generally assumed temperature
independent reducing the temperature prediction complexity. Afterwards the focus is shifted to: the
electromagnetic losses generated in the coils (at a reference temperature) and the adjacent oil flow
velocity. The influence of the latter variables is more clear in the newer version of IEC 60076-2-2011 [1]
by introducing two factors Q and S, both yielding the final hotspot factor, H.

In the first section some design differences between core-type and shell-type are depicted in order to put
both in perspective [120], then the content is focused in the windings by providing a state of the art of the
different approaches used by manufacturers, but also by users to predict the thermal performance of shell-
type units, with a special emphasis in: oil-flow velocity calculation, electromagnetic losses calculation,
temperature calculation and direct temperature measurements both during heat run tests and in-service
conditions.

Ultimately, some evidences of accumulated thermal stress (Hotspots) in a scrapped shell-type transformer
are reported, where several layers of insulating paper had been analysed [38].

CFD analysis indicates probable hotspots in the upper part of coil and close to the magnetic circuit.
Although even if that region seems to be less cooled, having the lowest oil velocities, some CFD findings
suggest that the maximum and average temperatures are weakly dependant on the oil flow [37], being not
conclusive the usefulness of the new Hotspot factor approach for shell-type, proposed in the newest IEC
standards [1]. Further efforts must be allocated to its comprehension and clarification.

The U and L shaped insulating structures seem to be key structures as they reduce the effective heat
transfer area in the outer/inner parts of the coils where the losses are expectably higher, so these pieces
should be taken in to account in daily thermal calculations, not only CFD [121].

6.2 Design aspects

The mainstream difference comparing to core-type, is that in shell-type transformers, the magnetic circuit
is around the windings as shown in Figure 6.1 (top). This concept opposes the core-type arrangement
where the windings are wounded concentrically around a magnetic circuit limb. As a result the magnetic
circuit in shell designs acts as a structural member which reduces the amount of external clamping and
enhances the mechanical capability of withstanding large forces. Besides this, the tank is ‘form-fit’
reducing the weight of oil and increasing the relative amount of cellulosic insulating weight.

133
Chapter 6: Shell-type transformer thermal modelling

Figure 6.1: Shell-type transformer typical geometric arrangement (on the top), a possible cooling circuit (on
the left) and a detail of a pressboard washer with its adjacent coil (on the bottom right). Images from [35].

In the classical arrangement named 2 High-Low (2 H-L), the Low Voltage Winding (LV) is composed of
two groups with the High Voltage Winding (HV) between them, as shown on the top of Figure 6.1. Each
winding or, more precisely, each part of the winding (LV or HV), is composed of coils. The coils
arrangement in such alternating groups is referred as interleaved. This allows the short-circuit forces of the
HV and LV windings to act in opposite directions, as depicted in Figure 6.2, thus partially cancelling each
other and further improving the capability of withstanding large forces.

134
Chapter 6: Shell-type transformer thermal modelling

Figure 6.2: Short-circuit forces into shell-type transformers. Acting forces (F) into a window between HV and
LV gap due to leakage flux B [120].

For a shell-type transformer, as the capacity increases, the size of each coil is similar, while the ampere-
turns are reduced by introducing additional coils. This maintains the forces magnitude size-independent as
well as the hydraulic characteristics of the coil ducts.

The coils are typically stranded with a rectangular cross section and multiple strands can be winded in
parallel, being in many cases necessary to transpose them in certain locations or even to use CTC.

A crucial component of each coil are the insulating pieces folded around the outer and/or inner edges of
the coil for dielectrical reasons and that have typically a U or L shape as shown in Figure 6.3. These are
important structures in the subsequent thermal models as they represent an effective reduction of the coil
heat transfer efficiency from the coil edge to the cooling oil so it should be considered in everyday design
Figure 6.3 (b).

135
Chapter 6: Shell-type transformer thermal modelling

Figure 6.3: (a) Zoom of the insulating pieces applied in the inner and outer edges of the coils in a shell-type
transformer. Images from [38]. (b) Influence of those insulating pieces on the temperature.

Different to core-type transformers, coils are rectangular-shaped, as shown in Figure 6.4. Usually in both
sides of the coil (generally made of copper) there is a washer made of pressboard with spacers distributed
along it. These spacers might have different shapes depending on the manufacturer as shown in bottom
right of Figure 6.1 or in Figure 6.5. The volume and cross section area between the pressboard washer and
the copper coil defines the duct through which oil flows - Figure 6.4 (a) – while the spacers’ height define
the oil duct size (typically between 4 and 6 mm). The ducts sizes might vary from coil to coil, or group to
group, depending on design considerations, as the case herein reported in Figure 6.6, Figure 6.7 and
Figure 6.8.

136
Chapter 6: Shell-type transformer thermal modelling

(a) (b)
Figure 6.4: (a) Coil over a pressboard washer with spacers. Zoomed view of a cooling duct. (b) Stacked coils
with the LV groups without L or U insulating pieces [120].

Afterwards each coil is stacked-up according to Figure 6.4 (b) and the spacer’s location must be
coincident from bottom to top of the pile in order to transmit forces homogeneously guaranteeing effective
mechanical stability.

Moreover the location and number of these structures (Figure 6.3) must be balanced in terms of
mechanical withstanding capability and heat transfer area covered. For example, the lower the L1 distance
shown in Figure 6.5 the better mechanical resistance but at the same time lower heat transfer area. A
typical L1 would be 70-90 mm for a heat transfer area covered of 25-40% in trapezoidal-shaped spacer
configuration as in Figure 6.5.

137
Chapter 6: Shell-type transformer thermal modelling

Figure 6.5: Pressboard washer with trapezoidal-shaped spacers and a highlight how the non-supported
copper distance can be evaluated (L1 in this case). Images modified from [37].

A comprehensive study over the influence of the mentioned spacer’s location in the temperature
distribution along a copper coil and its insulating structures can be found in [37]. In this study it is
concluded that, even with better oil flow distribution the temperature is weakly affected, revealing another
fundamental difference from core-type, which is the relative position between the main oil velocity
direction and the mean turn direction. In shell-type the oil flow follows a zig-zag pattern in most of its
path aligned with the mean turn direction, while in core-type the oil flows perpendicularly to the mean
turn direction. This promotes a faster uniformization of temperatures in shell-type.

Because of these structures and mainly due to the form-fit tank, the cooling regime in shell-type
transformers is intrinsically oil directed (forced or not) while in core type it depends on the design
considerations.

6.3 CFD models of shell-type units

6.3.1 General

First of all it is relevant to remark that the rectangular shaped coils do not exhibit circumferential
symmetry as in core type, which typically increases the needed mesh size. Furthermore, models need to be
studied in 3D because in 2D the friction forces are not sufficiently representative.

As previously said, the oil is confined to flow in channels (typically 4-6 mm height) and on its path it
suffers several contractions and expansions. Therefore, several flow structures arise, namely, stagnation
and high velocity zones, along with vortex formation influencing the effectiveness of heat transfer as well
as the overall pressure drop [36].

Computational Fluid Dynamics is the most suitable tool available at this moment to model these flow
characteristics as it comprises strong numerical techniques that solve the Navier-Stokes set of differential
equations by dividing the domain under study into finite elements. Herein a work where the global phase
is modelled [35] is presented as well as other additional findings that concern exclusively with the oil flow
inside the coil with its insulation pieces as well as its experimental validation [35], [36].

Finally it must be concluded that CFD is used up to now in a case-by-case basis and, to the best
knowledge of the authors, there is no analytical engineering tool, such as the THNs for core-type units,
138
Chapter 6: Shell-type transformer thermal modelling

being applied for shell-type units. Some authors tried to include CFD findings in an analytical model to
predict the temperature rise inside a shell coil. However the approach was uni-dimensional and targeted
for a rough indirect estimation of the viscometric degree of polymerization (DPv) of the insulating paper
[38].

6.3.2 CFD simulations of the global phase

A typical question in shell-type transformers is whether the pressboard structure around the inner and
outer edges of the coils introduce additional pressure drop, reducing the local oil flow rate and inducing
non-homogeneous distribution of oil at the entrance of each coil duct.

In order to answer this question, a CFD simulation of the global phase is herein reported. The domain
chosen comprises the windings and as well as the upper and lower tank volumes. Due to the different
scales involved (millimetres inside the coils and meters in the tank) the winding oil ducts have been
modelled through a porous medium, avoiding the detailed geometry representation inside oil ducts (e.g.
pressboard spacers). The U and L-shaped insulation pieces have been considered.

Therefore Figure 6.6 shows a single-phase unit where a quarter of the chosen transformer has been
modelled with two symmetry planes. The oil inlets have been placed on the bottom tank and the outlets at
the top of the tank. The boundary conditions consisted in a constant mass flow rate in the inlet and a
relative pressure -0 Pa- at the outlet, as shown in Figure 6.6, resembling a unit operating under an OD
cooling regime.

Figure 6.6: Global phase model [35].

Figure 6.7 shows a global overview of the oil velocity in the bottom of the tank highlighting that when the
entering in the tank the oil expands and homogenises, inducing what is typically called a ‘pool’ effect. In
practice, other inlet positions - perpendicular to the surface of the coil- are common. Although this had
been also studied in this work and no influence had been reported [35]. The total amount of oil in the tank
seems sufficient to create a homogeneous pressure at the entrance of each oil duct.

139
Chapter 6: Shell-type transformer thermal modelling

(a)
(b)
Figure 6.7: (a) Oil x-velocity contours (b) Velocity vectors coloured by oil velocity magnitude.

Figure 6.8 (a) presents the pressure drop along each oil duct (in both sides of each coil) – from the bottom
inlets to the top outlets. It can be seen that the pressure drop is similar in all oil ducts – 37.5 kPa – as
expected as all the oil ducts consist in hydraulic parallel resistances. The differences in the oil mass flow
rate plotted in Figure 6.8 (b) arise because in the studied unit there were different oil duct sizes
(corresponding to different spacer’s height and different coil dimensions).

(a) (b)

Figure 6.8: (a) Relative pressure drop of each oil duct. (b) The respective oil mass flow rate.

140
Chapter 6: Shell-type transformer thermal modelling

Figure 6.9: Oil velocity magnitude in each oil duct.

Figure 6.9 shows that the average oil velocity in each oil duct is quite similar. These findings simplify
further thermal models, because reveals that the oil flow distribution is area proportional being easier to
predict it when compared with core-type where the geometric differences between windings impose
additional unbalances, that are in some designs ‘tuned’ with additional elements to control the oil flow to
each winding.

6.3.3 CFD simulations inside coils

Once the distribution of the total oil flow rate among the oil ducts can be estimated the ducts can be
modelled as independent hydraulic elements.

In order to model these elements, authors [35], [36], reported to have used 3D domains representing half
of an oil duct as it is typically symmetrical and shown in Figure 6.10. The mesh-elements are polyhedral
and its total number is in this particular case 1 000 000.

The oil flow which is one of the boundary conditions for the detailed model, as shown in Figure 6.10,
must be either defined by imposing a mass flow inlet or a pressure inlet. Both seem to yield reasonable
results [35], [36].

(a) (b)

Figure 6.10: (a) General boundary conditions for a CFD model of an oil duct. (b)Typical CFD mesh [36].

141
Chapter 6: Shell-type transformer thermal modelling

In Figure 6.11 a typical inlet zone is shown on the top image – ranging from 45º left to 45ºC right of the
bottom curved zones. Typically the top outlets would have the same configuration. On the bottom image
of Figure 6.11 oil flow rate distribution among the above mentioned inlet zone is shown. The distribution
is not uniform and higher flowrate in the curved entrances are observed – this occurs due to a lower
hydraulic resistance of that particular oil path.

Figure 6.11: Typical inlet zone of a shell coil (top). Oil flow profile at the inlet (bottom) – shown as a
percentage of the total inlet oil flow rate.

142
Chapter 6: Shell-type transformer thermal modelling

These models allow identifying the zones of low and high oil velocities as shown in Figure 6.12.

Figure 6.12: Oil velocity contours along the oil path highlighting a low-velocity area (top) and high-velocity
area (bottom). Images modified from [35].

As shown in Figure 6.12 the oil mixing from inner zones and outer zones is not uniform. Generally the
lowest velocity regions are in lower and the upper inside edge of the coil, close to magnetic circuit.
Additionally, several different inlet conditions have been studied, from 11.8 kPa to 0.118 kPa with no
relative differences in oil distribution (Figure 6.13).

A typical quantity in the washers shown in Figure 6.13 is that the maximum oil velocity is around 5 times
higher than average oil velocity and it is localized in the inner curved zones.

143
Chapter 6: Shell-type transformer thermal modelling

Figure 6.13: Contours of oil velocity magnitude (m/s) in a middle plane for different pressure inlets: 11.8 kPa;
1.18 kPa; 0.118kPa [36].

This physical evidence gives at a first glance a sound basis for the need of introducing the S factor [1], as
for the core type. This non-uniformity in the oil velocity distribution suggests a thermal influence in some
extent, as a better oil mixing would expectably minimize temperature gradients. Although, the mentioned
dependence seems weak as reported in [37], whereas a better spacer/flow distribution contributed to
similar maximum and average temperatures, being at this moment not clear if an S-Factor is relevant or
not.

6.3.4 CFD flow field validation with PIV

As a validation step and in order to verify the oil velocity distributions, as previously described, there are
reported comparisons [35], [36], of the results of CFD simulations with PIV (Particle Image Velocimetry)
measurements on acrylic models as the one presented in Figure 6.14. The reported experiences had been
conducted in isothermal acrylic models, with no heated elements inside.

144
Chapter 6: Shell-type transformer thermal modelling

Figure 6.14: Model and PIV measurements.

PIV is a technique of measuring the velocity of particles seeded in a fluid. The particles are imaged at two
instants (t1 and t2) within a given short period of time. These two images are then processed by computer
to obtain their displacement and the corresponding velocity vectors, as shown in Figure 6.15.

(a) (b)

Figure 6.15: (a) Velocity magnitude contours inside a oil duct of shell-type transformer. (b) PIV
measurements and CFD simulations [36].

As seen in Figure 6.15, the experimental and the simulated oil flow patterns are identical. In [36] this
similarity has been evaluated for different oil flow rates as well as for different spacer distributions.

Figure 6.16: Velocity magnitude profiles along a horizontal line crossing spacers at 0.2m3/hr oil flow rate.
Different oil ducts (left and right) [36].

In Figure 6.16 the oil velocity magnitude is plotted along horizontal lines crossing spacers as depicted in
the top left of each image. It can be seen that the correlation is reasonable, endorsing CFD as a good tool

145
Chapter 6: Shell-type transformer thermal modelling

to model the laminar isothermal oil flow inside shell-type power transformers and consequently also give
a valuable insight into the estimation of the S-factor [1].

6.4 Leakage flux and electromagnetic losses calculation

6.4.1 General

In order to achieve a good thermal model of the transformer as important as good oil flow modelling (S
factor) is a good losses calculation (Q factor). The combination of both to obtain the hotspot will be
explained in Section 6.5.

In the current Section it is described the state of the art calculation of the losses in the active part due to
the current circulating through the windings. First are presented the different losses in the coils that
compose each winding and second the losses created in other elements of the active part due to the stray
flux.

6.4.2 In coils

In a shell-type coil the same load losses separation than in a core type can be done consisting in three
components:

 Joule losses

 Eddy losses

 Circulating current

The first component comes from the resistivity of the copper and can be reliably calculated applying the
Joule formula (RI2). For the eddy losses and the circulating current losses it is needed to estimate first the
magnetic field induced by the alternating current through the coils. According to the construction of the
shell-type transformers higher losses persist in the coils of the windings next to the high to low voltage
space - Figure 6.17 (a) - and especially in the coil edges, where the magnetic flux turns more
perpendicular to the conductors - Figure 6.17 (b).

146
Chapter 6: Shell-type transformer thermal modelling

(a) (b)

Figure 6.17: (a) Circulating Current Loss in different coils across H-L space - 2 H-L arrangement [120];(b)
Magnetic field representation in a H-L space – bottom represents the coil inner part.

In the past, the magnetic field had been calculated in 2-D (the two main directions in the surface of the
coil). Because of the geometry of the shell-type coils this could not be enough especially in the radius area
of the coil, as it does not have a cylindrical symmetry like core-type coils.

Nowadays ‘magnetic charges’ meshes can be established in the boundary surfaces to adequate the field to
the boundary conditions. These 3-D magnetic field calculations allows reducing the losses shown in
Figure 6.17 (a) - while evaluating the result of modifying the location of the transpositions and
connections between coils.

Depending on their position with respect to the core and the tank, the shell-type coils can be magnetically
divided in four different regions, exhibiting distinct flux density values - Figure 6.18. At the middle of the

147
Chapter 6: Shell-type transformer thermal modelling

limb region (LR), the radial flux densities in the inside and outside edges are practically the same and
account for 50% of the maximum value [120].

Figure 6.18: Magnetic shell-type coil regions: Top End Region (TR); Bottom End Region (BR); Limb Region
(LR); and Corner Region (CR) [120].

The significant variation on the magnitude of magnetic flux densities is located in the corner region (CR)
where the maximum leakage flux density is reported - Figure 6.19. The maximum flux occurs axially in
the inner corner (CR) region.

148
Chapter 6: Shell-type transformer thermal modelling

(a) (b)

Figure 6.19: Leakage flux density distribution. (a) Along the LV Coil at CR. (b) Axial and radial distribution
over the maximum value [120].

Besides this, from 3D magnetic field approaches the circulating currents can be estimated up to the
individual strand level which can then be included in a thermal model resulting in a temperature
calculation for each strand.

As an example, Figure 6.20 presents the circulating current magnitude in percentage of the nominal
current in one of the coils of a particular transformer wherein circulating currents are higher (coil located
in the high voltage group close to the neutral end, having each turn 6 layers of cables, with eight insulated
copper strands each). For each layer some of the strands have higher current than nominal and other lower
current than nominal. This unbalance must be considered when designing the transformer.

149
Chapter 6: Shell-type transformer thermal modelling

Figure 6.20: Circulating currents in percentage of the nominal current in the coil with higher magnitude of
circulating currents.

6.4.3 In other parts (tank, T-beams, shunts…)

The 3D calculations presented in the previous sub-chapter for the coils makes also possible to estimate the
magnetic field outside the windings: stray field. In a shell-type transformer there are three metallic
components surrounding the windings within the active part (apart from the magnetic circuit):

 The T-beams

 The tank

 The magnetic shunts


150
Chapter 6: Shell-type transformer thermal modelling

The T beams contribute to keep the winding package in vertical right position and withstand the core and
winding weights. They are usually two T-beam supports, one placed in the upper face of the core and the
bottom face the core - Figure 6.21 (a). These beams are located in the inside window of the phase and
receive the stray field. The tank as explained in the introduction is fitted to the active part so it will receive
also part of the flux from the phase. Finally, and to avoid localization of Hotspots in T-Beams and Tank
walls from the mentioned (stray) flux, magnetic shunts are placed. Both the amount of required shielding
and its optimum configuration imply 3D field calculations as the flux density is not uniform driving a
large part of the shielding into saturation - Figure 6.21 (b).

(a) (b)

Figure 6.21: (a) Upper and bottom T-beams; (b) Calculated flux density distribution at T-Beam shunts [120].

The magnetic field value is used to estimate the height of the magnetic shunts enough to limit the
induction below saturation and protecting the T beams (as depicted in Figure 6.22) and the tank from
reaching high temperatures.

Figure 6.22: Schematic representation of a T-beam with a magnetic shielding structure (magnetic shunt).

As an example, Figure 6.23 presents the calculated magnetic field in the top inner part of the windings
where the upper T-beam is located. The field along the phase window increases and decreases as the
diameter of the circles shows.

151
Chapter 6: Shell-type transformer thermal modelling

Figure 6.23: Magnetic field from the winding along the T-beam location, from the first coil on the right to the
last coil on the left.

The purple circles represent the magnitude of the field so that the higher the field the higher diameter of
the circles. In this case it is represented a 4 high to low spaces transformer so the higher magnetic field can
be noticed in these four high to low spaces [39].

6.5 Hotspot calculation (coupled CFD simulations)

The temperature rise in the windings is a consequence of the dynamic of oil coupled with the
electromagnetic losses generated in the coil. The temperature calculation has been historically pursuing
empirical approaches that many manufacturers adapt to their particular experimental experience. Besides
the heat source dependence, the hotspot calculation expression(s) reflect a joint contribution both from the
insulating paper wrapped around the coils (diffusive term) as well as from the oil flow (convective term).
The latter influence is modelled through oil convective coefficient which in turn depends upon the oil
velocity among other physical properties. The heat generated in the coils is then all transferred to oil, in
steady-state regime, through the above mentioned ‘barriers’: oil and paper (typically).

Three dimensional (3D) CFD modelling allows solving simultaneously mass, momentum and energy
equations (coupled CFD simulations). The technique is based on the discretization of differential
equations and their primary unknowns are properties as: mass, velocity and temperature in each grid point.
Such approach requires additional assumptions, namely about how each property vary between the grid
points. However the relative importance of these assumptions decreases when the number of grid points
increase. Depending on the particular models used, it enables to simulate finer local heat transfer and
accurately estimate hotspot location. The velocity differences in the oil duct as shown in Figure 6.9 are
taken into account and finally different electromagnetic losses hypothesis could be used as heat source
terms. For example, uniform losses repartition in all the coil, non-uniform losses in each turn/layer and
ultimately in each copper conductor.

Figure 6.24 shows a coupled steady-state CFD simulation where the estimation and location of the hotspot
was possible in one coil where variable electromagnetic losses were introduced in each turn. The flow was
modelled as laminar with a flow rate of 0.88 kg/s.

152
Chapter 6: Shell-type transformer thermal modelling

Figure 6.24: CFD thermal-hydraulic coupled simulation typical output, (left) Contours of temperature on the
coil, (right) contours of temperature on the duct side.

In Figure 6.24 a hotspot of 94.8ºC was calculated in the upper part of the coils near the inner edge (close
to the magnetic circuit), with an inlet oil temperature of 68.0°C. Average coil temperature was 80.3°C
while oil average temperature was 72.2°C (integrated value on the entire oil domain). Oil at the channel
output (average value on the output face of the model) was calculated at 77.2 °C. It conducts in that
specific case to a Hotspot factor higher than 2.

This matches the velocity contours shown in Figure 6.12 in that location revealing oil velocities close to
zero near to the magnetic circuit where the heat transfer is mainly due to natural convection. Additionally
in this area has U-Shaped pressboard folded around the coil edge which enhances the over-heating
potential.

Regarding meshes for such model, this study shows the need to work with 3D model. Figure 6.25
introduces an example of the sensitivity of the Hotspot calculation with variable number of uniform mesh
elements in the thickness (smallest dimension) of the oil duct.

153
Chapter 6: Shell-type transformer thermal modelling

Figure 6.25: CFD thermal-hydraulic coupled simulation - sensitivity to mesh refinement.

The flow was considered as laminar in this model and the oil duct was 5 mm thickness.

Table 6.1: Number of mesh elements and the respective maximum temperature in the coil.

As depicted in Table 6.1, above 7 elements the differences are negligible as they are lower than the
accuracy of any measurement (-0.2ºC). Additionally, the finest mesh with 10 elements in the thickness
leads to a final number of 3D elements of more than 10 000 000 elements. So, 7 represent in this case the
optimum mesh.

However the effect of the refinement might assessed in a case-by-case basis as it can be more or less
significant depending on the models used and on the flow regime (laminar, turbulent model), the
sensitivity of the simulation tools, and the hypothesis of calculation. Thus the mesh profile has to be
adjusted and refined each time it has to be done with the target of the best ratio accuracy of
results/computation time.

6.6 Direct measurements (FO location)

The use of fibre optic (FO) is really valuable for qualification of the transformer during factory acceptance
test. CFD and electromagnetic modelling of the transformer can bring recommendation for the
implementation of those FO inside the active part by using a coupled thermo-hydraulic simulation.
However in addition to possible uncertainties linked to the quality of the assembly of the FO, the location
shall of course be representative of what happens at the real hotspot. Due to manufacturing constraints but
also design considerations (dielectric, cooling channel), it is not always possible to put the FO at the
estimated or calculated hotspot location.

For example the position calculated for the design and model introduced in Section 6.5 is not really
suitable (Figure 6.26). In that case customer and manufacturer could come to an agreement for a potential

154
Chapter 6: Shell-type transformer thermal modelling

corrective factor to apply at the measured value on the basis of their calculation model as introduced
before.

Figure 6.26: Contours of temperature in a coil underlining different locations of optical fibre (FO) and the
calculated Hotspot.

More details on FO installation in shell-type transformer according to manufacturing constraints are


presented in Chapter 5 of the Brochure.

After the transformer is commissioned and the measured values by FO validated in factory, the FO can
provide valuable information for monitoring hotspot and thermal behaviour of transformers on site. The
first year monitoring, for example, enables to compare with the performance measured in the laboratory
and to follow the evolution of thermal characteristics related to real operating conditions.

Hereafter is introduced a comparison (from [40]) between the measurements made during factory
acceptance test and the analysis of the measurements during one year in operation for a step-up single
phase unit installed in a nuclear power plant. In Table 6.2 is a summary of the oil temperature rises and of
the hotspot factors calculated from the FO expressed respectively as per unit of the average oil
temperature-rise and as the average gradients of HV and LV windings, deduced from heat-run test.

155
Chapter 6: Shell-type transformer thermal modelling

Table 6.2: Performance in operation (On Site) comparing with Heat Run Test results, both in an OD regime
and at rated load (no-load losses + load losses).

On Site
Heat Run Test
Oil temperature-rise
unity 1pu = ∆θ mo-r of Unit in factory
Δθ bo-r(1) 0.9  1.0
 1.1
Δθ mo-r(2) 1.0
 1.2
Δθ to-r(3) 1.1
Hotspot factors
unity 1pu = gr LV(4) or gr HV(4) of Unit in factory
 1.5
Hgr_LV1(5) 1.3
 1.3
Hgr_LV2 1.0
 1.9
Hgr_HV1(6) 1.8
 1.5
Hgr_HV2 1.5

(1) ∆θ bo-r: Bottom oil (in tank) temperature rise


(2) ∆θ mo-r: Average oil (in tank) temperature rise
(3) ∆θ to-r: Top-oil (in tank) temperature rise in steady state
(4) gr LV or gr HV: Average winding to average oil temperature gradient at rated current
(5) Hgr_LV1 or Hgr_LV2: Hotspot to top-oil gradient for 2 hotspots in LV winding
(6) Hgr_HV1 or Hgr_HV2: Hotspot to top-oil gradient for 2 hotspots in HV winding

 the top-oil temperature-rise is essentially constant and slightly increased by    0 .1

 HV hotspot gradients evolve on average slightly, but nevertheless about to H   0 .1 for the
hottest hotspot

 LV hotspot gradients increase more significantly in the order of: H   0 .2 to 0 .3 .

This kind of analysis could raise a discussion between the customer and the manufacturer about
representativeness of the heat run test in factory related to real on site conditions in order to reach an
optimal tuning of the regulated refrigeration system of the transformer.

6.7 Hotspot evidences

The experimental analysis of scrapped/deactivated transformers can provide useful evidences over the
thermal performance and distribution of every transformer (both core-type and shell-type). In this case,
some authors reported paper analysis in a 30-year old shell-type transformer from the Portuguese Grid –
Pracana substation. The mentioned transformer had been deactivated due to a network rearrangement [38].

156
Chapter 6: Shell-type transformer thermal modelling

An extensive program of collecting paper samples had been accomplished in 8 coils from different groups
(HV,MV and LV) - Figure 6.27.

(a) (b)

Figure 6.27: (a) Image of the works during the scrapping of a shell-type coil. (b) Image of the two outer layers
after being cuted.

Each coil had been sampled in the same seven locations. The sampled locations are shown in Figure
6.28 (a). Afterwards the viscometric degree of polymerization had been measured for each region sampled
and the average values are shown in Figure 6.28 (b).

(a) (b)

Figure 6.28: (a) Sampled locations in each coil. (b) Average DPv distribution for each point and each coil [38].

In this analysis it had been concluded that:

 DPv point 1 > DPv point 2 > DPv point 3 > DPv point 4 ≈ DPv point 5;

 in half of the discs coils the lowest DPv occurs in point 7.


157
Chapter 6: Shell-type transformer thermal modelling

The location identified as point 7 represent an identical location of the hotspot of the coil simulated
previously in this report (Section 6.5), sustaining the importance of that region as one of the main
candidates to probably containing the hotspot.

6.8 Summary remarks

CFD is being used to assess the thermal performance of shell-type transformers. Its intrinsic mathematical
models are expected to be suitable to analyse both the temperature and velocity fields, as its windings are
designed to operate in laminar regimes. Anyhow some experiments with acrylic prototypes have been
reported with the purpose of validating experimentally CFD predictions.

CFD can be a valid tool either for design or operation purposes and its scope has not been limited to the
windings. Some work has also been conducted by analysing the oil distribution in the bottom of the tank
and no significant pressure differences have been depicted reinforcing at this moment the ‘natural’ theory
of a homogeneous pool of oil.

In the daily calculations, and for reasons that are depicted in other Chapters of this Brochure, CFD is not
used to design every shell-type transformer. Its applicability is not envisaged for this purpose in the next
5-10 years, but can be used in the future if some of its limitations are withdrawn or become more mature.
For daily calculations, typical empirical correlations are still used confidently.

The reasons behind this fact need to be followed closely but detailed thermal analysis with CFD seems to
indicate a performance less dependent on oil flow velocities while comparing with core-type. The
‘natural’ hotspot location is expected to be under the U and L-shaped pieces folded around the edges of
each coil and not occasionally they embody the most convenient location to place optical fibre probes.

Equally, electromagnetic phenomena have been depicted, since losses are the heat source used as input for
CFD model. Their profile is depending of design with significant differences compared to core-type
(structure, T-beam...). Losses are computed in daily calculations for identification of “high losses” area.
They determine the thermal behaviour and hotspot location, as a result of a combined effect with the oil
flow profile. A universal ‘law’ seems difficult to define for the location of optical fibre probes as these
pieces might be differently designed from manufacturer to manufacturer.

The added-value of direct temperature measurement in operation is also highlighted through an example
of a shell-type GSU where different hotspot factors in a LV winding have been found when comparing the
heat run test against on-site performance.

At the end, a scrapped shell-type transformer is analysed through its paper samples basically to assess
hotspot evidences, but the ageing laws are far more complex and determined by other variables rather than
just temperature. The ultimate ambition of reaching its comprehension should be pursued but out of the
scope of this working group.

158
REFERENCES

[1] IEC 60076-2 Ed 3.0 (2011-02) Power transformers – Part 2: Temperature rise for liquid-immersed
transformers.
[2] J. Jaspar and C.J.G. Spoorenberg, “Practical aspects of determining the hotspot factor in large power
transformers,” THJ2b conference, November 2011, Warwick UK.
[3] K. Karsai D. Kerenyi L. Kiss , “Large Power Transformers,” ISBN 0-444-99511-0 (Vol. 25)
[4] A.K. Bose, C. Kroon, J. Wildeboer, “The loading of the magnetic circuit,” CIGRE report 12.09,
1978.
[5] P. Price, “Geomagnetically induced current effects on transformers,” IEEE Transactions on Power
Delivery, Vol. 17, No. 4, October 2002.
[6] S.V. Kulkarni, S.A. Khaparde, “Transformer Engineering Design & Practice,” ISBN 0-824-75653-3,
Marcel Dekker Inc., New York, 2004.
[7] L. Lin, C. Xiang, “Loss calculation in Transformer Tie Plate Using the Finite Element Method,”
IEEE Transactions on Magnetics, Vol. 34, No. 5, September 1998.
[8] D. Pavlik, D.C. Johnson and R.S. Girgis, “Calculation and reduction of stray and eddy losses in core-
form transformers using a highly accurate finite element modeling technique,” IEEE Transactions on
Power Delivery, Vol. 8, No. 1, January 1993.
[9] D.A. Koppikar et al., “Evaluation of flitch plate losses in power transformers,” IEEE Transactions on
Power Delivery, Vol. 14, No. 3, July 1999.
[10] A. J. Oliver, “Estimation of transformer winding temperatures and coolant flows using a general
network method,” Proc. Inst. Elect. Eng., pt. C, Vol. 127, 395–405, 1980.
[11] “Calculation of short-circuit forces in transformers,” Electra No 67, page 29 to 75, Final report
prepared by Working Group 12-04, 1979.
[12] P. Picher, F. Torriano, M. Chaaban, S. Gravel, C. Rajotte, B. Girard, “Optimization of transformer
overload using advanced thermal modelling,” CIGRE 2010, Paris, A2-305.
[13] C. Spoorenberg, The statistical analysis of temperature rise test for increased accuracy on x- and y-
exponents, TJH2b conference Chicago USA 2012.
[14] C. Spoorenberg, Statistical analysis of temperature rise test for increased accuracy on x- and y-
exponents, Cigre conference Dubrovnik 2012.
[15] H. Nordman, O. Takala, “Transformer Loadability Based on Directly Measured Hotspot Temperature
and Loss and Load Current Correction Exponents”, CIGRE report No. A2_307_2010.
[16] W. Lampe, L. Pettersson, C. Ovren, B. Wahlstrom, “Hotspot measurements in power transformers”,
CIGRE report No. 12-02, 1984.
[17] P. Heinzig, “Smart spacers for hotspot measurement accuracy and dynamic behaviour,” Tutorial
Session IEEE Transformers Committee Meeting, Munich 21 March 2013.
[18] K. Spoorenberg, “Instruction on the installation of fiber optic probes in Smit-windings,” Smit
Transformers (Nijmegen 2011, unpublished).
159
[19] M. Cuesto, “Instruction on the installation of fiber optic probes in Shell Type transformers,” ABB
Trafo, (Cordoba 2011, unpublished).
[20] V. Davydov, J. Wijaya, “Effect of Oil Flow Rate on Temperature Distribution in Disc Type
Windings,” Prepared Contribution at CIGRE, SC A2, PS3, Question no. 3-8, Paris 2010.
[21] IEC 60076-7, Loading guide for oil-immersed power transformers, 2005-12.
[22] IEEE Std C57.91-1995, IEEE Guide for Loading Mineral-Oil-Immersed Transformers.
[23] IEEE Std C57.91-2011, IEEE Guide for Loading Mineral-Oil-Immersed Transformers.
[24] H. Nordman, “Temperature rises in oil ducts at ON- and OF-cooling,” ABB internal research report
1ZFI 13-023 (Vaasa 2013-03-06, unpublished).
[25] H. Nordman, Prepared contribution at Cigre, SC 12, PS1, Question no. 6, Paris 1984.
[26] E. Hiironniemi, H. Nordman, J. Elovaara, T. Ojanen, R. Andersson, H. Eronen, K. Heinonen,
”Experiences of ON- and OFF-line condition monitoring of power transformers in service,” Cigre
report 12-102, 1992.
[27] L-A. Lundin, “Fiberoptisk temperaturmatning på transformatorer till Kimstad,” ASEA R ZK 86-133
(Ludvika 1986-09-02, unpublished).
[28] M. Cuesto & al, “Fast deployable HV network transformer with hybrid insulation: a solution against
major events,” CIGRE SC A2/C4 Joint Colloquium, Zurich, Sept. 2013.
[29] H. Nordman, “Temperature rises of unshielded and shielded yoke clamps of 175 MVA auto-
connected transformer”, ABB internal research report nr 1ZFI 95-98 (Vaasa 1995-12-05,
unpublished).
[30] H. Nordman, M. Lahtinen, “Thermal Overload Tests on a 400 MVA Power Transformer With a
Special 2.5 pu Short Time Loading Capability”, IEEE Trans. Power Delivery, vol. 18, pp. 107-112,
January 2003.
[31] H. Nordman, Prepared Contribution at CIGRE, SC A2, PS3, Question no. 3-7, Paris 2010.
[32] H. Nordman, Prepared Contribution at CIGRE, SC A2, PS3, Question no. 3-6, Paris 2012.
[33] N. Günter, “Comparison of surface mounting and hole mounting of fiber optic sensors on yoke
clamps,” Powertech Transformers (Pretoria 2011, unpublished).
[34] J. Aubin, “Application of Magnetically Mounted Sensor for Top Oil Temperature Measurement,”
(Montreal 2011, unpublished).
[35] F. Berthereau JST, « Shell Type Transformer Thermal Modelling » CIGRE 2010 presentation.
[36] Paulo J. Gomes, Renato G. Sousa, Madalena M. Dias, José C.B. Lopes, Manuel Silvestre, Duarte
Couto, Paulo Mesquita, “Large Power Transformer Cooling – Flow Simulation and PIV analysis in
an Experimental Prototype”, ARWtr2007 Advanced Research Workshop on Transformers, Baiona,
Spain, 2007.
[37] Paulo J. Gomes, Renato G. Sousa, Joana I. Cardoso, Madalena M. Dias, José C.B. Lopes, Manuel
Silvestre, Duarte Couto, J. Ramos, Paulo Mesquita, Mário Maia, “Studies in a Large Power
Transformer – Heat Transfer and Flow Optimization using CFD”, ARWtr2007 Advanced Research
Workshop on Transformers, Baiona, Spain, 2007.

160
[38] M. Augusta Martins, Mónica Fialho, Jorge Martins, Mário Soares, Maria Cristina, R. Castro Lopes,
Hugo M. R. Campelo, “Power Transformer End-Of-Life Assessment – Pracana Case Study,” IEEE
Electrical Insulation Magazine (Accepted. To be published. November 2011).
[39] A. Jalinat, P. Hurlet, A. Fernández, M. Oliva, J. Porrero, A. Prieto, M. Cuesto “Large generator step
up transformers with low temperature hot spot for EDF Nuclear Power Plants,” CIGRE Paper
A2_303, 2010.
[40] A. Tanguy, M. Ryadi, J. Channet (EDF R&D), P. Hurlet (EDF CNEPE), “In service experience of
hot spot behaviour of a GSU power transformer compared to temperature rise test results,” CIGRE
Paper PS3-O-10, A2&D1 Joint Colloquium 2011, Kyoto-Japan.
[41] Z. Radakovic, M. Sorgic, “Wirtschaftliche Betrachtung der thermischen Auslegung von ölgekühlten
Leistungstransformatoren,” Elektrizitätswirtschaft, Jg. 107, Heft 15, 32-38, 2008.
[42] Z. Radakovic, K. Feser, “A new Method for the calculation of the hotspot temperature in power
transformers with ONAN cooling,” IEEE Trans. on Power Delivery, Vol. 18, No. 4, 1284-1292,
2003.
[43] L. W. Pierce, “An investigation of the thermal performance of an oil filled transformer winding,”
IEEE Transactions on Power Delivery, Vol. 7 , Iss. 3 , July 1992, pp.1347-1358.
[44] J. Zhang and X. Li, “Oil cooling for disk-type transformer windings-Part I: Theory and model
development,” IEEE Trans. Power Delivery, Vol. 21, 1318 – 1325, 2006.
[45] J. Zhang, and X. Li, “Oil cooling for disk-type transformer windings-Part II: Parametric studies of
design parameters,” IEEE Trans. Power Delivery, Vol. 21, 1326 – 1332, 2006.
[46] M. Yamaguchi, T. Kumasaka, Y. Inui, S. Ono: “The flow rate in a self-cooled transformer,” IEEE
Transactions on Power Apparatus and Systems, Vol. PAS-100, pp. 956-963, 1981.
[47] J. Declercq, W. Van der Veken, E. Van den Bulck: „Accurate hotspot modelling in a power
transformer using a general network model, Proc. of the IEE conference Cyprus ’98.
[48] I. E. Idelchik: “Handbook of Hydraulic Resistances,” CRC Press, Boca Raton, Ann Arbor, London,
Tokyo, 1994.
[49] VDI-Gesellschaft Verfahrenstechnik und Chemieingenieurwesen (GVC): „VDI Wärmeatlas,
Berechnungsblätter für den Wärmeübertragung,“ 8. Auflage, Springer Verlag, 1997.
[50] F. P. Incropera, D. P. DeWitt: „Heat and Mass Transfer,“ Fifth Edition, John Wiley & Sons, 2002.
[51] Z. Radakovic, M. Sorgic: “Basics of Detailed Thermal-Hydraulic Model for Thermal Design of Oil
Power Transformers,” IEEE Trans. Power Del., Vol. 25, no. 2, pp. 790-802, April 2010.
[52] M. Sorgic, Z. Radakovic: “Oil-Forced Versus Oil-Directed Cooling of Power Transformers,” IEEE
Trans. Power Del., Vol. 25, no. 4, pp. 2590-2598, October 2010.
[53] AB KIHLSTROMS MANOMETERFABRIK, Instruments for Power and Distribution Transformers
& Instruments for SF6 Gas Insulated Switchgear.
[54] "IEC 60354: Loading Guide for Oil-immersed Power Transformers", 1993
[55] "Direct Measurements of the Hotspot Temperature of Transformers", CIGRÉ WG 12–09, vol. 129,
1990.

161
[56] W. Lampe, L. Pettersson, C. Ovren, and B. Wahlström, “Hotspot Measurements in Power
Transformers,” Cigre, Rep. 12-02, International Conference on Large High Voltage Electric Systems,
1984 Session, 29th August-6th September.
[57] H. Nordman, E. Hironniemi, and A.J. Pesonen, “Determination of hotspot temperature rise at rated
load and at overload,” CIGRE Paper 12-103, 1990.
[58] H. Nordman, N. Rafsback, and D. Susa, “Temperature responses to step changes in the load current
of power transformers,” IEEE Transactions on Power Delivery, Vol. 18 , Iss. 4 , Oct. 2003, pp.1110
– 1117.
[59] N.T. Räfsbäck, “Short-time emergency overloading of power transformers, Bachelor thesis,” Power
transformer company, ABB, Vaasa, 2001.
[60] M.P. Saravolac, “The use of optic fibres for temperature monitoring in power transformers,” IEE
Colloquium on Condition Monitoring and Remanent Life Assessment in Power Transformers, 1994.
[61] E.A. Simonson, and J.A. Lapworth J.A., “Thermal capability assessment for transformers,” Second
International Conference on the Reliability of Transmission and Distribution Equipment 1995., 29-31
Mar 1995, pp.103 – 108.
[62] M.V. Thaden, S.P. Mehta, S.C. Tuli, and R.L. Grubb, “Temperature rise tests on a forced-oil-air
cooled (FOA) (OFAF) core-form transformer, including loading beyond nameplate,” IEEE
Transactions on Power Delivery, Vol. 10, Iss. 2, April 1995, pp. 913 – 923.
[63] D.J. Tylavsky, Qing He, Jennie Si, G.A. McCulla, J.R. Hunt, “Transformer top-oil temperature
modeling and simulation,” IEEE Transactions on Industry Applications,Vol.36, Iss.5, Sept.-Oct.2000,
pp. 1219 – 1225.
[64] J. Aubin, R. Bergeron, and R. Morin, “Distribution Transformer Overloading Capability Under Cold-
Load Pickup Conditions,” IEEE Transactions on Power Delivery, Vol. 5, Iss. 4, October 1990, pp.
1883-1891.
[65] Montsinger VM, Clem JE. Temperature rise for short time overloads. AIEE Transactions 1946;
65:966.
[66] L.W. Pierce, “Predicting liquid filled transformer loading capability,” IEEE Transactions on Industry
Applications, Vol. 30 , Iss.1, Jan.-Feb. 1994, pp.170-178.
[67] D. Susa, M. Lehtonen, and H. Nordman, “Dynamic Thermal Modelling of Distribution
Transformers,” IEEE Transactions on Power Delivery, Vol. 20, Iss. 4, October 2005, pp. 1919 –
1929.
[68] D. Susa, J. Palola, M. Lehtonen, and M. Hyvarinen, “Temperature Rises in an OFAF Transformer at
OFAN Cooling Mode in Service,” IEEE Transactions on Power Delivery, Vol. 18, Iss. 4, October
2003, pp. 1110 – 1117.
[69] D. Susa, M. Lehtonen, and H. Nordman, “Dynamic Thermal Modelling of Power Transformers,”
IEEE Transactions on Power Delivery, Vol. 20, Iss. 1, January 2005, pp. 197 – 204.
[70] D. Susa, M. Lehtonen, “Dynamic Thermal Modelling of Power Transformers-Further Development:
Part I,” IEEE Transactions on Power Delivery, Vol. 21, NO. 4, October 2006, pp. 1961 – 1970.
[71] D. Susa, M. Lehtonen, “Dynamic Thermal Modelling of Power Transformers-Further Development:
Part II,” IEEE Transactions on Power Delivery, Vol. 21, NO. 4, October 2006, pp. 1971 – 1980.

162
[72] D. Susa, H. Nordman, “A Simple Model for Calculating Transformer Hotspot Temperature,” IEEE
Transactions on Power Delivery, Volume 24, No. 3, July 2009, pp. 1257 - 1265.
[73] G.L. Alegi, and W.Z. Black, “Real-time thermal model for an oil-immersed, forced-air cooled
transformer,” IEEE Transactions on Power Delivery, Vol. 5, Iss. 2, April 1990, pp.991-999.
[74] H.J. Blake, and J.E. Kelly, “Oil-Immersed Power Transformer Overload Calculation by Computer,”
IEEE Power Transactions on Power Apparatus and Systems, Vol. Pas-88, August 1969, pp. 1205.
[75] J. Declercq, and W. Van der Veken, “Accurate hot spot modeling in a power transformer leading to
improved design and performance,” Transmission and Distribution Conference, 1999 IEEE ,Vol. 2 ,
11-16 April 1999, pp. 920-924.
[76] B.C. Lesieutre, W.H. Hagman, J.L.Jr. Kirtley, “An improved transformer top oil temperature model
for use in an on-line monitoring and diagnostic system,” IEEE Transactions on Power Delivery,
Volume: 12 , Issue: 1 , January 1997, pp. 249 – 256.
[77] L.W. Pierce, and T. Holifield, “A thermal model for optimized distribution and small power
transformer design,” Transmission and Distribution Conference, 1999 IEEE ,Vol. 2 , 11-16 April
1999 , pp. 925 - 929.
[78] M.K. Pradhan, and T.S. Ramu, “Prediction of hottest spot temperature (HST) in power and station
transformers,” IEEE Transactions on Power Delivery, Vol. 18 , Iss. 4, October 2003, pp.1275 – 1283.
[79] Z. Radakovic, D. Kalic, “Results of a novel algorithm for the calculation of the characteristic
temperatures in power oil transformers,” Electrical Engineering 80, 1997, pp. 205–214.
[80] Z. Radakovic, “Numerical determination of characteristic temperatures in directly loaded power oil
transformer,” European Transaction on Electrical Power (ETEP), vol.13, no.1, 2003, pp.47-54.
[81] S.A. Ryder, “A simple method for calculating winding temperature gradient in power transformers,”
IEEE Transactions on Power Delivery, Vol. 17 , Iss. 4, October 2002, pp. 977 – 982.
[82] W.H. Tang, Q.H. Wu, and Z.J. Richardson, “Equivalent heat circuit based power transformer thermal
model,” Electric Power Applications, IEE Proceedings, Vol. 149, Iss. 2, March 2002, pp. 87 – 92.
[83] W.H. Tang, Q.H. Wu, and Z.J. Richardson, “A simplified transformer thermal model based on
thermal-electric analogy,” IEEE Transactions on Power Delivery, Vol. 19, Iss. 3 , July 2004, pp. 1112
– 1119.
[84] W. Van der Veken, J. Declercq, M. Baelmans, and S. Van Mileghem, “New perspectives to
overloading with accurate modeling of thermal transients in oil-immersed power transformers,”
Transmission and Distribution Conference and Exposition, 2001 IEEE/PES ,Vol. 1 , 28 Oct.-2 Nov.
2001, pp. 147 – 152.
[85] G. Swift, T.S. Molinski, and W. Lehn, “A fundamental approach to transformer thermal modelling- I.
Theory and equivalent circuit,” IEEE Transactions on Power Delivery, Vol. 16, Iss. 2, April 2001, pp.
171 – 175.
[86] G. Swift, T.S. Molinski, R. Bray, and R. Menzies, “A fundamental approach to transformer thermal
modelling-II. Field verification,” IEEE Transactions on Power Delivery, Vol. 16, Iss. 2, April 2001,
pp. 176 – 180.
[87] W.Seitlinger, “A Thermo-Hydraulic Transformer Model”, CEPSI Seminar, Manila, Philippines,
October 2000.

163
[88] W.Seitlinger, E.Knoll, G.Buchgraber, “Modeling Loadability of an Operating Transformer” , EPRI
Substation Equipment Diagnostic Conference IX, 2001
[89] M.Scala, G.Buchgraber, W.Seitlinger, “Transformer Overloading, utilizing an on-line thermo-
hydraulic Transformer Model”, EPRI Substation Equipment Diagnostics Conference XI, 2003.
[90] D. Susa, M. Lehtonen, and H. Nordman, “Transformer Temperature Calculation During Transient
States,” Modern Electric Power Systems MEPS'02, Wroclaw University of Technology, Wroclaw,
Poland, September 11-13, 2002.
[91] D. Susa and M. Lehtonen, “New aspects on the dynamic loading of power transformers,” Fifth
Nordic Distribution and Asset Management Conference, NORDAC 2002, Copenhagen, Denmark,
November 7-8 2002.
[92] CIGRE Brochure 393: "Thermal Performance of Power Transformers", CIGRE WG A2.24, 2009.
[93] T. Gradnik, D.Susa, and P.Picher, “In-service accuracy evaluation of transformer Loading guide
models,” 2012 CIGRÉ Canada Conference, Hilton Montréal Bonaventure, Montréal, Québec,
September 24-26, 2012.
[94] T. Gradnik, M. Konèan-Gradnik, “A rectifier transformer ageing study using optical fibre
temperature sensors,” Transformer thermal modelling and EIMV physical-chemical diagnostics,
Proceedings CIGRE Joint Colloquium A2/D1, Bruges, 2007.
[95] IEEE Std C57.119-2001, “IEEE recommended practice for performing temperature rise tests on oil-
immersed power transformers at loads beyond nameplate ratings”.
[96] “IEC 72-2 Power Transformers; Part 2: Temperature rise", 1994-04 second edition".
[97] D. Susa, H. Nordman, “IEC 60076-7 Loading Guide Thermal Model Constants Estimation,”
European Transactions on Electrical Power, 23 FEB 2012 DOI: 10.1002/etep.1631.
[98] D. J. Tylavsky, Qing He, Gary A. McCulla, James R. Hunt, “Sources of error in substation
distribution transformer dynamic thermal modelling”, IEEE Transactions on Power Delivery
2000;15(1):178-185.
[99] D. J. Tylavsky, “Transformer top-oil temperature modeling and simulation,” IEEE Transactions on
Industry Applications 2000;36(5):1219-1225.
[100] D. J. Tylavsky, Xiaolin Mao, Gary A. McCulla, “Transformer thermal modeling: Improving
reliability using data quality control,” IEEE Transactions on Power Delivery 2006;21(3):1357-1366.
[101] Matlab manual: Nonlinear Least Squares (Curve Fitting), available online:
http://www.mathworks.com/help/optim/ug/lsqcurvefit.html.
[102] L.E. Lundgaard, W. Hansen, D. Linhjell, T.J. Painter, "Aging of oil-impregnated paper in power
transformers," Power Delivery, IEEE Transactions on , vol.19, no.1, pp.230,239, Jan. 2004.
[103] N. Lelekakis, Wenyu Guo, D. Martin, J. Wijaya, D. Susa, "A field study of aging in paper-oil
insulation systems," Electrical Insulation Magazine, IEEE , vol.28, no.1, pp.12,19, January-February
2012.
[104] N. Schmidt, S. Tenbohlen, R. Skrzypek ,B. Dolata, “Assessment of Overload Capabilities of Power
Transformers by Thermal Modelling,” CIGRE SC A2 & D1 Joint Colloquium 2011, Kyoto Japan.

164
[105] Carruthers, M. G. and Norris, E. T., “Thermal ratings of transformers introduction of multiflow
principle,” Proc. IEE, vol. 116, no. 9, pp. 1564–1570, 1969.
[106] T. Raymond. “Sample implementation of the Annex G model calculation method”, available online:
http://goo.gl/ymuki.
[107] A. Oluwaseun, D.J. Tylavsky, G. McCulla, W. Knuth, “Evaluation of Hottest-Spot Temperature
Models using Field Measured Transformer Data,” International Journal of Emerging Electric Power
Systems. Volume 12, Issue 5.
[108] E.J. Kranenborg, C. O. Olsson, B. R. Samuelsson, L-Å. Lundin, R. M. Missing, “Numerical study
on mixed convection and thermal streaking in power transformer windings,” 5th European Thermal-
Sciences Conference (EUROTHERM 2008), The Netherlands, 2008.
[109] J. Y. Lee, S. W. Lee, J. H. Woo and I. S. Hwang, “CFD analyses and experiments of a winding with
zig-zag cooling duct for a power transformer,” CIGRE Biannual meeting 2010, Paris, France, 2010.
[110] C.O.Olsson, “Buoyancy driven flow between vertical parallel plates of finite width,” ASME-JSME
Thermal Engineering Summer Heat Transfer Conference, Vancouver (CA), 2007.
[111] R. Bel Fdhila et al, “Thermal modeling of power transformer radiators using a porous medium based
CFD approach,” ThermaComp 2011.
[112] S. Tenbohlen, A. Weinläder, R. Wittmaack, “Prediction of the Oil Flow and Temperature
Distribution in Power Transformers by CFD,” CIGRE Biannual meeting 2010, Paris, France, 2010.
[113] F. Torriano, M. Chaaban, P. Picher, “Numerical study of parameters affecting the temperature
distribution in a disc-type transformer winding,” Applied Thermal Engineering 30 (2010) 2034-2044.
[114] F. Torriano, P. Picher, M. Chaaban, “Numerical investigation of 3D flow and thermal effects in a
disc-type transformer winding,” Applied Thermal Engineering 40 (2012) 121-131.
[115] S. Chen, F. Devaux, “CFD investigation on thermal performance of different types of windings for a
300 MVA transformer,” CIGRE Canada 2012.
[116] J. Gastelurrutia, J. C. Ramos, G. S. Larraona, A. Rivas, J. Izagirre, L. del Rio, “Numerical
modelling of natural convection of oil inside a distribution transformer,” Applied Thermal
Engineering 31 (2011) 493-505.
[117] J. Smolka, “CFD-based 3-D optimisation of the mutual oil configuration for the effective cooling of
an electrical transformer,” Applied Thermal Engineering 50 (2013) 124-133.
[118] N. Schmidt, S. Tenbohlen, S. Chen and C. Breuer, “Numerical and experimental investigation of the
temperature distribution inside oil-cooled transformer windings,” 18th International Symposium on
High Voltage Engineering (ISH), Hannover, Germany, 2013.
[119] N. Schmidt, S. Tenbohlen, R. Skrzypek ,B. Dolata, “Assessment of Overload Capabilities of Power
Transformers by Thermal Modelling,” CIGRE SC A2 & D1 Joint Colloquium 2011, Kyoto Japan.
[120] X.M.López-Fernández, H.B. Ertan, J. Turowski,”Transformers: Analysis, Design and
Measurement”, CRC press, 2012.
[121] M. Cuesto, “Thermal modeling of Shell-Form transformers,” CIGRE Seminar on Sidney TechCon
Asia-Pacific, 2011.

165
[122] N. Schmidt, “Studies on radial temperature variation in winding discs,” (Universitat Stuttgart 2012,
unpublished).
[123] Qing He, Jennie Si, D.J. Tylavsky, “Prediction of top-oil temperature for transformers using neural
networks,”, IEEE Trans. on Power Delivery, vol.15, no.4, pp.1205,1211, Oct. 2000.
[124] O. Roizman, V. Davydov, W. Guo, A. Petersen, and P. Cole, “Comparison of various approaches to
transformer thermal modelling with direct temperature measurements,” CIGRE A2-304, 2010.
[125] Min-gu Kim, Sang Moon Cho, Joong-Kyoung Kim, “Prediction and evaluation of the cooling
performance of radiators used in oil-filled power transformer applications with non-direct and direct-
oil-forced flow,” Exp. Therm. Fluid Sci. 44 (2013) 392–397.
[126] Sang Moon Cho, Chul Suk Kimo, Joong-Kyoung Kim, “CFD Analysis and classical approach to
calculate the efficiency of a radiator in power transformer, International Symposium of High Voltage
Engineering 17th,” Hannover, (2011), A-013.
[127] W. Wu, Z.D. Wang, A. Revell and P. Jarman. “CFD Calibration for Network Modelling of
Transformer Cooling Flows – Part II Pressure Loss at Junction Nodes”, IET Electric Power
Applications, 2012.
[128] W. Wu, Z.D. Wang, A. Revell, H. Iacovides and P. Jarman. “CFD Calibration for Network
Modelling of Transformer Cooling Flows – Part I Heat Transfer in Oil Ducts”, IET Electric Power
Applications, 2012.
[129] Z. Radakovic, U. Radoman and P. Kostic: “Decomposition of the Hot-Spot Factor”, IEEE Trans. on
Power Delivery, Vol. 30, No. 1, 403-411, 2015.
[130] Z. Radakovic, M. Sorgic, W. Van der Veken and G. Claessens. “Ratings of Oil Power Transformer
in different Cooling Modes”, IEEE Trans. on Power Delivery, Vol. 27, No. 2, 618-625, 2012.
[131] H. M. R. Campelo, C. M. Fonte, R.G. Sousa, J. C. B. Lopes, R. C. Lopes, J. Ramos, D. Couto, M.
M. Dias, ‘Detailed CFD Analysis of ODAF Power Transformer’, International Colloquium
Transformer Research and Asset Management, Cigré HRO, 2009.
[132] A. Weinlader, W. Wu, S. Tenbohlen, and Z. Wang, “Prediction of the oil flow distribution in oil-
immersed transformer windings by network modelling and computational fluid dynamics,” IET
Electr. Power Appl., vol. 6, no. 2, pp. 82–90, 2012.
[133] H. M. R. Campelo, C. M. Fonte, R.G. Sousa, J. C. B. Lopes, R. C. Lopes, J. Ramos, D. Couto, M.
M. Dias, “Network modelling applied to CORE power transformers and validation with CFD
simulations”, International Colloquium Transformer Research and Asset Management, Cigré HRO,
2012.
[134] P. Boss, H. Brandle, “Measurement of the temperature profile in transformer windings with optical
distributed sensors,” EPRI Conf. Substation Equipment Diagnostics, 1997.
[135] A. Weinlader, S. Tenbohlen, “Thermal-hydraulic investigation of transformer windings by CFD-
modelling and measurements”, Proceedings of the 16th International Symposium on High Voltage
Engineering, Johannesburg, 2009.
[136] H.M.R. Campelo, M.A. Quintela, F. Torriano, P. Labbé and P. Picher, “Numerical thermofluid
analysis of a power transformer disc-type winding”, IEEE Electrical Insulation Conference,
Montreal, 2016.

166
Annex A: Installation altitude influence on transformer thermal modelling

Annex A
Installation altitude influence on transformer thermal
modelling
A.1 Why altitude influence thermal behavior?

Atmospheric pressure p decrease with altitude H and according to the "International Standard
Atmosphere” (ISO 2533:1975) these variations can be represented with the following formula valid for
altitudes between 0 to 10000 meters above sea-level (MASL):


p Pa   101325 1  2.2558  10 5 H m   5.25588 (A.1)

The "International Standard Atmosphere” supposed 15ºC (288.15 K) for ambient temperature and
1013.25 mbar (101325 Pa) for atmospheric pressure at sea-level but is possible to use other base values.

This atmospheric pressure variation with altitude will affect the value of the air density which is
fundamental in heat transfer by natural and forced convection and this is the reason of the altitude
influence in the thermal behavior of transformers.

From the ideal gases equation we obtain air density in function of ambient temperature and atmospheric
pressure:

 kg  1 pPa 
 3 
 (A.2)
 m  287.053 T K 

Replacing (A.1) in (A.2) we have air density in function of altitude and ambient temperature:

 H  
352.984
T K 

1  2.2558  10 5 H m   5.25588
(A.3)

In the next paragraphs we will use equation (A.3) to estimate the influence of installation altitude in the
convection heat transfer in the transformer coolers.

A.2 Installation altitude influence in the Standards

The installation altitude influence in thermal behavior of transformers creates the need for a correction
when we make heat-run tests in laboratories with a altitude different to the installation altitude of the
transformer to verify his thermal ratings.

The IEC 60076-2:2011 standard have rules to take into account the installation altitude influence. It says:

" If the installation site is more than 1000 m above sea-level but the factory is not, then the allowable
temperature rises during the test in the factory shall be reduced as follows:

 for a naturally cooled transformer (....AN), the limit of top-liquid, average and hot-spot winding
temperature rises shall be reduced by 1 K for every interval of 400 m by which the installation's
altitude exceeds 1000 m;

 for a forced-cooled transformer (…. AF), the reduction shall be 1 K for every 250 m exceeding
1000 m.
167
Annex A: Installation altitude influence on transformer thermal modelling

A corresponding reverse correction may be applied in cases where altitude of the factory is above 1000 m
and the altitude of the installation site is below 1000 m.

Any altitude correction shall be rounded to the nearest whole number of degrees kelvin."

ANSI Standards altitude corrections are little different in his formulations but the results obtained are in
line with the IEC corrections.

This standard altitude correction is not clearly explained in the standards. For instance, why AF is worse
than AN ?

For practical reasons, the standard correction is applied only for installation's altitudes that exceed 1000 m.
As presented in this annex, the natural phenomena are continuous and the altitude influence begins at 1 m
above sea-level.

The altitude influence in transformer temperatures is due to the atmospheric pressure reduction with
altitude. Air density is depending of atmospheric pressure, and this affect the cooling efficiency of
radiators or heat exchangers for any altitude, not only for altitudes higher than 1000 m above sea-level.

A.3 Installation Altitude Influence Fundamentals

A.3.1 Dimensionless Numbers used in Convection Phenomena

hD   cP g     2    D 3   D V
Nu  Pr  Gr  Re 
k k  2

 2  cP    g 3
Ra  Pr Gr   D     
k 

Nu = Nusselt Number Pr = Prandtl Number Gr = Grashof Number

Re = Reynolds Number Ra = Rayleigh Number

 
h W / m 2 º C = convection heat transfer coefficient

 º C  = convection temperature drop

D m  = characteristic geometric dimension

V m / s  = fluid velocity

g  9.81 m / s 2 = acceleration of gravity

 kg / m3  = fluid density

 kg / m s  = fluid dynamic viscosity

 m / s 2  = fluid kinematic viscosity

c P W s / kg º C  = fluid specific heat

168
Annex A: Installation altitude influence on transformer thermal modelling

 1 /º C  = fluid expansion coefficient

k W / m º C  = fluid conductivity

A.3.2 Correlations for the Heat Transfer Coefficient in Natural Convection

For natural convection Nusselt is a function of Prandtl and Grashof: Nu  f Gr , Pr 

Nu  a  Pr Gr 
n
Typical correlations for natural convection in a vertical plane are:

For laminar air flow ( 5  102  Ra  2  107 )

1
1
  2  cP    g 3 4
Nu  0.54  Pr Gr 
1
4  0.54  Ra  0.54  
4
 D   
 k  
1 1 1

hD   2  cP    g 4  k 3   2  c P    g  4  4
 0.54    D 3     h  0.54     1

k  k      D4

For turbulent air flow ( 2  107  Ra  1  1013 )


1
1
  2  c P   g 3
Nu  0.135  Pr Gr 
1
3  0.135  Ra  0.135  
3
 D 3   
 k  
1 1

hD   2  cP    g 3  k 2   2  cP    g  3 1

 0.135    D 3     h  0.135      3

k  k     

We see that in both cases (natural or turbulent flow) the dependence of the heat transfer coefficient by
natural convection with density and altitude is of the form:
K 5.25588 K
h1  1   1  2.2558  10 5 H 1 
h  cte   K
      
h2   2   1  2 .2558  10 5
H 2 

The exponent K is between 1/2 y 2/3 depending of the air flow type (laminar or turbulent).

A.3.3 Correlations for the heat transfer coefficient in forced convection

For forced convection Nusselt is a function of Prandtl and Reynolds: Nu  f (Re, Pr)

Typical correlations for forced convection in channels are: Nu  a  Re n  Pr m

One of the most used correlation is Colburn Formula (Dittus–Boelter):

Nu  0.023  Re0.8  Pr 0.4  

169
Annex A: Installation altitude influence on transformer thermal modelling

L/DH 1 2 5 10 15 20 30 40 50
 1.65 1.50 1.34 1.23 1.12 1.13 1.07 1.03 1.00

where L is channel height and DH is the hydraulic diameter of the channel

 L  L
       60    1.00
 DH  DH

   D H V     cP 
0 .8
h  DH
0 .4

 0.023      
k     k 

k 0.6   0.8  cP 
0 .4
h  0.023    V 0. 8
 0. 4
DH
0. 2

We see that the dependence of the heat transfer coefficient by forced convection with density and altitude
is of the form:
0.8 4.204704
h1  1   1  2.2558  10 5 H 1 
h  cte   K
     
h2   2  5
 1  2.2558  10 H 2 

It is noted that the dependency of the heat transfer coefficient with density is greater in forced convection
(exponent K=0.80) than in natural convection (exponent K between 0.50 and 0.66).

That's why the altitude affects more forced convection (1°C increase every 250 meters increase in altitude
for AF) that natural convection (1°C increase every 400 meters increase in altitude for AN).

A.3.4 Convection Heat Transfer Coefficient and Temperature Variation with Altitude

Applying convection Newton's law:

Q  h  S    Q  h1  S  1  h2  S   2

h1  1  h2   2

K 5.25588 K
 2 h1  1   1  2.2558  10 5 H1 
      
1 h2   2   1  2.2558  10 5 H 
 2 

where Q is the heat transferred by convection, S is the heat transfer area in contact with the air and  is
the average oil temperature rise (disregarding the oil convection temperature drop in the coolers).
Conditions 1 and 2 correspond to different installation altitudes. Exponent K is between 0.50 to 0.66 for
AN cooling and 0.80 for AF cooling.

 For Natural Convection with Laminar Flow we have:


0 .5 2.62794
 2   1   1  2.2558  10 5 H 
H1  H and H 2  H  400      
 1   2   1  2.2558  10 H  400  
5

170
Annex A: Installation altitude influence on transformer thermal modelling

It can be proved, with good approximation, for any H value,


that  1 @ H  2 @ H  400
30.0 30.8
 2
 1.025  40.0 41.0
1
50.0 51.3

Then in this case, for natural convection and laminar flow, with constant ambient temperature, the
transformer temperatures increase between 0.8 ºC and 1.3 ºC for each 400 m increase in altitude (1.0ºC for
a typical average oil temperature rise value of 40ºC).

 For Natural Convection with Turbulent Flow we have:


2/3 3.50392
 2   1   1  2.2558  10 5 H 
H1  H and H 2  H  400      
1   2   1  2.2558  10 H  400  
5

It can be proven, with good approximation, for any H value, that:

 2  1 @ H  2 @ H  400
 1.034 
1
30.0 31.0
40.0 41.4
50.0 51.7

Then in this case, for natural convection and turbulent flow, with constant ambient temperature, the
transformer temperatures increase between 1.0 ºC and 1.7 ºC for each 400 m increase in altitude.

 For Forced Convection we have:


0 .8 4.204704
 2   1   1  2.2558  10 5 H 
H1  H and H 2  H  250      
1   2   1  2.2558  10 H  250  
5

It can be proven, with good approximation, for any H value, that:

 1 @ H  2 @ H  250
 2
 1.025  30.0 30.8
1
40.0 41.0
50.0 51.3

Then in this case, for forced convection, with constant ambient temperature, the transformer temperatures
increase between 0.8 ºC and 1.3 ºC for each 250 m increase in altitude (1.0ºC for a typical average oil
temperature rise value of 40ºC).

171
Annex A: Installation altitude influence on transformer thermal modelling

A.4 Examples

We analyze altitude correction examples with different average oil temperature rises and altitude
installation. In all cases we suppose that the transformers was tested at sea-level and will be installed at
altitudes higher than sea-level (2000, 2500 and 3000 MASL).

The results for AN corrections are summarized in Table 1 and for AF corrections are summarized in Table
2. For each example we apply three different correction methods:

o IEC1: correction according to IEC 60076-2:2011 (see page 2)


o Linear:
 for a naturally cooled transformer (....AN), the limit of top-liquid, average and hot-spot
winding temperature rises is reduced by 1 K for every interval of 400 m by which the
installation's altitude exceeds sea-level;
 for a forced-cooled transformer (…. AF), the reduction is 1 K for every 250 m exceeding sea-
level.
o HTF: correction according to formula (4) based on heat transfer fundamentals (HTF)

To clarify, the calculations for 2500 MASL AF correction for Example 3, with an average oil temperature
rise of 35ºC, are shown:

2500  1000
o IEC:  6º C
250
2500
o Linear:  10º C
250
4.204704
 2  1  2.2558 10 5 H1 
o HTF:  
1  1  2.2558 10 5 H 2 

 2 
4.204704
1 
H 1  0 , H 2  2500     1.276
1  1  2.2558  10 5  2500 
 2
 1.276   2  1.276  1  1.276  35  44.7º C
1
 2  1  44.7  35  9.7 º C  Correction  10º C

A.5 Conclusions

 The actual standard altitude corrections (IEC) show some differences compared with the method
based on heat transfer fundamentals (HTF) principally for higher average oil temperature rises.

 With the Linear approach, the altitude corrections are very similar compared with the corrections
obtained using the method based on heat transfer fundamentals (HTF). The agreement is better for
higher average oil temperature rises (practically identical) and the results are similar for the lower
average oil temperature rises.

 The altitude corrections are not independent of the average oil temperature rise. The altitude
correction according to formula (4) is a factor not a constant adding term. Using this factor we obtain,
for each average oil temperature rise, the adding term that will be applied to all the temperatures of the
transformer.

172
Annex A: Installation altitude influence on transformer thermal modelling

Table A.1: AN correction examples.


Example 1 Example 2 Example 3
AN IEC Lin HTF IEC Lin HTF IEC Lin HTF
 AVG-OIL 31 35 40
 TOP-OIL
@ 0 MASL

39 45 52
G 25 20 15
Hxg 35 28 21
 AVG-WDG 56 55 55
 HS-WDG 74 73 73
 AVG-OIL 34 36 35 38 40 40 43 45 45
@ 2000 MASL

 ATOP-OIL 42 44 43 48 50 50 55 57 57
G 25 25 25 20 20 20 15 15 15
Hxg 35 35 35 28 28 28 21 21 21
 AVG-WDG 59 61 60 58 60 60 58 60 60
 HS-WDG 77 79 78 76 78 78 76 78 78

Example 1 Example 2 Example 3


AN IEC Lin HTF IEC Lin HTF IEC Lin HTF
 AVG-OIL 30 34 39
 ATOP-OIL
@ 0 MASL

38 44 51
G 25 20 15
Hxg 35 28 21
 AVG-WDG 55 54 54
 HS-WDG 73 72 72
 AVG-OIL 34 36 35 38 40 40 43 45 45
@ 2500 MASL

 ATOP-OIL 42 44 43 48 50 50 55 57 57
G 25 25 25 20 20 20 15 15 15
Hxg 35 35 35 28 28 28 21 21 21
 AVG-WDG 59 61 60 58 60 60 58 60 60
 HS-WDG 77 79 78 76 78 78 76 78 78
Example 1 Example 2 Example 3
AN IEC Lin HTF IEC Lin HTF IEC Lin HTF
 AVG-OIL 29 33 37
 ATOP-OIL
@ 0 MASL

37 43 49
G 25 20 15
Hxg 35 28 21
 AVG-WDG 54 53 52
 HS-WDG 72 71 70
 AVG-OIL 34 37 35 38 41 40 42 45 45
@ 3000 MASL

 ATOP-OIL 42 45 43 48 51 50 54 57 57
G 25 25 25 20 20 20 15 15 15
Hxg 35 35 35 28 28 28 21 21 21
 AVG-WDG 59 62 60 58 61 60 57 60 60
 HS-WDG 77 80 78 76 79 78 75 78 78

173
Annex A: Installation altitude influence on transformer thermal modelling

Table A.2: AF correction examples.

Example 1 Example 2 Example 3


AF IEC Lin HTF IEC Lin HTF IEC Lin HTF
 AVG-OIL 29 33 37
 TOP-OIL
@ 0 MASL

37 43 49
G 25 20 15
Hxg 35 28 21
 AVG-WDG 54 53 52
 HS-WDG 72 71 70
 AVG-OIL 33 37 35 37 41 40 41 45 45
@ 2000 MASL

 ATOP-OIL 41 45 43 47 51 50 53 57 57
G 25 25 25 20 20 20 15 15 15
Hxg 35 35 35 28 28 28 21 21 21
 AVG-WDG 58 62 60 57 61 60 56 60 60
 HS-WDG 76 80 78 75 79 78 74 78 78
Example 1 Example 2 Example 3
AF IEC Lin HTF IEC Lin HTF IEC Lin HTF
 AVG-OIL 27 31 35
 ATOP-OIL
@ 0 MASL

35 41 47
G 25 20 15
Hxg 35 28 21
 AVG-WDG 52 51 50
 HS-WDG 70 69 68
 AVG-OIL 33 37 35 37 41 40 41 45 45
@ 2500 MASL

 ATOP-OIL 41 45 43 47 51 50 53 57 57
G 25 25 25 20 20 20 15 15 15
Hxg 35 35 35 28 28 28 21 21 21
 AVG-WDG 58 62 60 57 61 60 56 60 60
 HS-WDG 76 80 78 75 79 78 74 78 78
Example 1 Example 2 Example 3
AF IEC Lin HTF IEC Lin HTF IEC Lin HTF
 AVG-OIL 26 30 34
 ATOP-OIL
@ 0 MASL

34 40 46
G 25 20 15
Hxg 35 28 21
 AVG-WDG 51 50 49
 HS-WDG 69 68 67
 AVG-OIL 34 38 35 38 42 40 42 46 45
@ 3000 MASL

 ATOP-OIL 42 46 43 48 52 50 54 58 57
G 25 25 25 20 20 20 15 15 15
Hxg 35 35 35 28 28 28 21 21 21
 AVG-WDG 59 63 60 58 62 60 57 61 60
 HS-WDG 77 81 78 76 80 78 75 79 78

174
Annex B: Guidelines to CFD simulations

Annex B
Guidelines to CFD simulations
B.1 Introduction

As it was mentioned in section 2.6.1, CFD is a very powerful numerical tool for thermal and flow analysis
but some care should be taken in order to ensure the quality and accuracy of the solution. Here below
some examples are given in order to instruct future CFD users about the effect that some parameters such
as mesh discretization, boundary conditions and 2D modelling can have on the flow and temperature
distribution in the winding.

B.2 Mesh discretization

After creating the geometry of e.g. a transformer winding via a CAD tool and prior to the numerical
analysis of that geometry via a CFD tool, the respective computational domain has to be discretized. For
that purpose, an appropriate mesh density should be chosen. While numerical errors result from a mesh
density chosen too coarse, a discretization scheme chosen too fine will increase the associated calculation
time and costs without offering additional benefits as an increased numerical accuracy. Consequently, a
mesh sensitivity analysis should be carried out when a high numerical accuracy is desired and the
numerical efforts want to be kept within reasonable boundaries. To give an insight into such an analysis,
Figure B.1 shows a conductor pair including the surrounding oil channels of an OD-cooled transformer
winding. Characteristic dimensions of interest for discretization are marked and enlargements of three
created meshes are displayed. Table B.1 gives further details regarding these characteristic dimensions and
their discretization in Mesh 1, 2 and 3.

Figure B.1: Marked dimensions of interest for discretization and enlargement of the investigated meshes with
clockwise increasing mesh density [119].

175
Annex B: Guidelines to CFD simulations

Table B.1 : Details about the chosen mesh densities for the dimensions of interest [119]

e f g h

Physical dimension [mm] 6.0 4.0 39.2 11.2

Nodes in Mesh 1 28 28 60 20

Nodes in Mesh 2 36 36 120 40

Nodes in Mesh 3 44 44 180 60

After creating meshes with increasing density of the computational domain of interest, a representative
operating state should be chosen to compare the results from CFD analysis gathered with each mesh. After
performing the respective calculations, the results concerning values of key interest should be compared.
As an example, Figure B.2 shows the temperature gradients between the single conductors of two passes
in upward direction (oil guides are placed before conductor number 13 and 25) and the oil surrounding the
respective conductor. Additional remarks about the investigated OD-cooled winding are given in [119]
and it is partially shown in Figure B.1. It can be noted that the deviations of the temperature gradients
between Mesh 1 and Mesh 2 are up to 1 K while the results of Mesh 2 and Mesh 3 are almost identical.
Consequently, the discretization scheme Mesh 2 already yields reliable results with regard to numerical
accuracy at the investigated operating state while the finer discretization scheme Mesh 3 gives no
additional benefit in terms of accuracy. Therefore, the appropriate mesh size in that case is given by the
discretization scheme Mesh 2.
8

7
Temperature Gradient / [K]

4
Mesh 1
Mesh 2
Mesh 3
3
14 16 18 20 22 24 26 28 30 32 34 36
Conductor Number / [1]

Figure B.2: Comparison of the temperature gradients between oil and conductors inside two passes of a
numerical winding model for three different discretization schemes [119]

In addition to the global density of the discretization, especially the discretization of the so-called
boundary layers is of importance. Taking a close look at the displayed meshes in Figure B.1 reveals an
uneven distribution of the single cells especially from the mid of a channel towards its next solid
boundary. This increase of the mesh density is necessary to capture gradients of variables being part of the
governing equations (Navier-Stokes), particularly with regard to the velocity and temperature gradients. A
noteworthy feature of the commonly applied oil products within power transformers is their high ratio of
momentum to thermal diffusivity often described with the Prandtl number. This results in steeper
temperature gradients in comparison to the velocity gradients. Consequently, an accurate coverage of
176
Annex B: Guidelines to CFD simulations

these gradients requires a respectively fine chosen discretization of the fluid domain close to solid
boundaries. Beside the global mesh density and the appropriate discretization of the boundary layers other
characteristics of the discretization also influence the numerical accuracy. Beside a preferably orthogonal
orientation of the cell faces respective to the local flow direction, parameters like skewness, aspect ratio or
volume change can be consulted to check the quality of the discretization.

B.3 Boundary conditions

To investigate the importance of selecting appropriate inlet boundary conditions for the CFD simulation,
the third pass of the 66 MVA LV winding presented in Chapter 3 of the Brochure is chosen as an example.
The main objective of this study is to determine whether the “real” temperature and velocity profile should
be specified (whenever possible) at the inlet of a pass or if constant values are sufficient to determine flow
and temperature distributions in the winding. For this study uniform losses of 677 W/disc are specified
and the simulations are 2D.

Figure B.3: Uniform and non-uniform temperature profiles specified at the inlet of the third pass.

Figure B.3 shows the velocity and temperature profiles that were obtained from a CFD calculation of two
upstream passes (i.e., passes 1 and 2) and that represent the “real” profiles present at the inlet of the third
pass. On the same image are also illustrated uniform profiles that correspond to the average value of the
non-uniform profiles. In order to determine the effect of the inlet boundary condition, three different
simulations are performed: one with the “real” velocity and temperature profiles, one with the “real”
velocity profile but a uniform temperature profile and the last one with uniform velocity and temperature
profiles.

The flow and temperature distributions in the winding obtained for the three different cases are shown in
Figure B.4 and the results are greatly affected by the type of profile that is specified at the entrance of the
pass. This seems particularly true for the temperature profile, since the flow distribution becomes more

177
Annex B: Guidelines to CFD simulations

asymmetrical with a uniform profile and the location of the hot spot shift toward the upper section of the
pass.

Figure B.4: Flow and temperature distribution as a fonction of inlet boundary condition.

As it can be seen in Figure B.5, a cold streak is present at the entrance of the pass and its dynamic
considerably affects the temperature of the lower discs in the pass. If a uniform temperature profile is set,
this effect is obviously not present and errors can be induced in the solution.

178
Annex B: Guidelines to CFD simulations

Figure B.5: Temperature contours as a function of inlet boundary condition.

In conclusion, the user should be quite careful in choosing the appropriate boundary conditions for the
numerical model and whenever possible “realistic” inlet profiles should be specified based on
experimental measurements or simulations. If not possible, a sensitivity study should then be performed in
order to quantify the effect of the boundary condition on the temperature distribution in the winding.

B.4 Numerical model (2D vs. 3D)

Two-dimensional modelling is widely used in transformer simulations in order to significantly reduce the
computational cost of such calculations. This section provides some information regarding the effect that
this kind of model simplification can have on the CFD results.

As illustrated in Figure B.6, the geometry of a winding includes a series of sticks, intersticks and duct
spacers but those elements are neglected in a 2D (i.e., axisymmetrical) simulation. To determine if 3D
effects are important, the third pass of the 66 MVA LV winding presented in Chapter 3 is once again
chosen as an example and uniform losses of 677 W/disc are specified.

179
Annex B: Guidelines to CFD simulations

Figure B.6: Low voltage winding geometry.

The profiles of Figure B.7 show that the flow and temperature distributions are quite different between the
two models and in the 2D case a more asymmetrical flow repartition is observed and it causes larger
temperature gradients in the winding. This discrepancy is in part due to the restriction effect that the sticks
and duct spacers have on the oil flow since they reduce the cross-sectional area of the cooling channels
compared to the axisymmetric geometry. Thus, for an equal mass flow rate, the bulk velocity in the
cooling duct will be lower in the 2D case and consequently the adimensional flow parameters such as the
Reynolds and Grashof numbers differ between the 2D and 3D case (i.e., the flow regimes are not
identical). This variation in flow regime also has an impact on the hot streak dynamics (see Figure B.8)
which directly affects the temperature distribution in the winding.

Moreover, in the 2D case the entire surface of the disc is in contact with the oil whereas in reality a
portion of this surface in not wetted since it is covered by the sticks and duct spacers, thus reducing the
effective heat transfer surface.

The magnitude of 3D effects can vary depending on the winding configuration and the operating cooling
mode, but the user should always be aware of the impact that geometrical simplifications can have on the
numerical solutions. Since three-dimensional CFD computations are often not feasible from a practical
point of view, a possible solution to improve the predictions of a 2D model is to adjust the flow rate in
order to better match the flow regime Reynolds and Grashof numbers of the 3D case and to then apply
correction coefficients to account for the difference in oil temperature and wetted disc surface. This
methodology is more thoroughly explained in [99].

180
Annex B: Guidelines to CFD simulations

Figure B.7: Flow and temperature distribution for 2D and 3D models.

Figure B.8: Temperature contours for 2D and 3D models.

181
Annex C: Examples of fiber-optic measurements

Annex C
Examples of fiber-optic measurements
C.1 Increased knowledge on OFAF-cooling from fiber-optic measurements

C.1.1 Unsymmetry of the oil circulation in OF-cooled transformers

An experimental study has been conducted using a 3-phase 0.5/0.9 MVA, 22/4.5/0.4 kV
ONAN/ONAF/OFAF-cooled mineral oil Special Test Transformer installed at the Centre for Power
Transformer Monitoring, Diagnostics and Life Management of Monash University, Australia [20]. Figure
C.1 shows the transformer a) at the factory during assembly, b) installed at Monash.

Figure C.1: Monash Special Test Transformer. Oil inlet from cooler header into the transformer tank is
located at the bottom of Phase C.

The on-line data gathering system included the following monitors: current and voltage transducers;
power quality analyser; temperature - 16 optic sensors, 9 RTDs and 20 thermocouples; moisture in oil – 5
relative saturation sensors; oil pressure – 3 sensors; oil flow rate – 1 ultrasonic flow meter. The pump was
equipped with a variable speed drive that allowed controlling the oil flow rate from the rated value of
1200 litres/min to a lower rate of down to 250 litres/min.

Two tests, Test A and Test B, were conducted in the OFAF-cooling mode overnight at approximately
similar load profiles and ambient temperatures, but for different flow rates. The applied current and
ambient temperature in Test A were slightly higher than in Test B. Thus, the temperature rises above
ambient temperature recorded at a well defined steady state were selected and extrapolated to the same
load current. Such a steady state had been obtained at 12 hours, when the ambient temperature was 26.5
ºC in Test A and 23.0 ºC in Test B. The corresponding load currents were 1.75 pu and 1.65 pu.

183
Annex C: Examples of fiber-optic measurements

The measured gradients above ambient temperature at 12 hours of Test B are extrapolated from 1.65 pu to
1.75 pu as follows:

 Layer winding according to the current exponent 2.0 (at OF-cooling the oil exponent is 2.0, and
the winding-to-oil gradient exponent is assumed to be 2.0 for layer windings).

 Disc winding according to the current exponent 1.6 (an approximation of the oil exponent 2.0 at
OF-cooling and of 1,3 for the winding-to-oil gradient of disc windings [21].

The values obtained are collected in Table C.1 and Table C.2.

Table C.1: Temperature rises above ambient temperature in steady state of 4,5 kV layer winding.
Test A Test B, 250 l/min
1200 l/min, Values observed at 1,65 Values upgraded to 1,75
FO Sensor No.
Values observed at 1,75 pu pu
pu / K K K
FO12 (Top) 86.9 76.5 86.0

FO13 (Mid-height) 72.8 62.5 70.3

FO14 (Bottom) 53.5 45.4 51.1

A comparison of the left hand and the right hand (gradient) columns shows that the oil circulation speed
has a very marginal influence on the local gradient to the ambient temperature.

Table C.2: Hotspot temperature rises above ambient temperature in steady state of 22 kV disc winding.
Test A Test B, 250 l/min
1200 l/min, Values observed at 1,65 Values upgraded to 1,75
FO Sensor No.
Values observed at 1,75 pu pu
pu / K K K
FO15 (Phase A) 69.0 61.5 67.6

FO06 (Phase B) 55.8 51.5 56.6

FO16 (Phase C) 49.0 57.0 62.6

The fiber-optic sensors were located in the same way in all of the 3 phases (below disc no, 2 seen from the
top at the same spot of the circumference). Thus, the big differences and the decreasing gradients in Test
A must be related to the location of the oil inlet next to Phase C as per Figure C.1.

For the disc-type winding, at low oil circulation speed the oil flow pattern inside the transformer obviously
changes, since the hotspot of Phase C starts to approach the hotspot of Phase A in Test B in spite of its
vicinity to the oil inlet.

This experimental finding demonstrates the need of taking into account the value of the oil flow rate as a
parameter of a thermal model of oil-immersed transformers. Currently, the IEC Loading Guide [21] and
the IEEE Loading Guide [23] specify an OFAF cooling mode as a generic one without specifying the oil
flow rate value.

184
Annex C: Examples of fiber-optic measurements

C.1.2 Another OFAF example

The findings from the Monash experimental transformer have been confirmed by measurements on a real
OFAF-cooled transformer fitted with point sensors to measure hotspot temperatures. The transformer was
tested at 22.5 MVA at ONAF and at OFAF at two oil flows: rated (100 %) and lower flow (50 %). Also
the oil temperatures at the top of the windings (exiting from the cooling ducts) were measured in the
ONAF and OFAF (100 %) tests. The results for the low voltage winding of phase C are summarized in
Table C.3 [24].

Table C.3: Temperature rises above ambient temperature.

Test Tank bottom Tank top oil LV-wdg top oil LV-wdg temp. LV-wdg
oil rise hotspot

ONAF 13.7 52.8 58.4 57.3 90.7

OFAF 100 % 29.8 37.1 56.9 61.8 85.7

OFAF 50 % 30.9 45.1 - 62.6 84.8

The two OFAF tests clearly show that increasing the total OFAF oil flow has very little effect on winding
temperatures: in fact hotspot rises are slightly increased when the oil flow is increased (a similar result
was obtained for the HV-winding, too).

Because the total OFAF oil flow has so little influence on the amount of oil passing through the windings
it is sometimes said that under OFAF cooling the winding oil flows are determined by buoyancy.
However, even at only 50 % of total flow the winding oil flows are likely to be substantially greater than
ON flowrates, so presumably winding oil flows are driven primarily by a pressure head introduced by the
external pump, albeit only the small fraction of the total pump head across the ‘shorted’ windings.

Lastly, the large amount of oil bypassing the windings for OFAF cooling will mean that oil temperatures
measured at the top of the tank will not be representative of oil temperature exiting the windings, so
winding gradients determined in the usual manner from the difference between mean winding and top oil
rises could be substantially in error. In this case the calculated winding gradient based on the top oil in the
tank is 28.4 K and on the top oil in the winding 18.5 K. The calculated Q-factor of the winding is 1.8.
Because the winding was axially cooled it can be assumed that the S-factor is 1.0, i.e. the calculated
hotspot factor H is 1.8. This factor yields the hotspot rises of 88.2 K and 90.2 K when based on the top oil
in the tank and in the windings, respectively. It means that the overestimated gradient is overcompensated
by the underestimated oil temperature rise. Hence the recommendation in Annex A of IEC 60076-2 that
the measured tank top oil should be used to calculate hotspot temperatures for OFAF, even though this
cannot be a good estimate of the temperature of the oil leaving the windings. Nevertheless, if it were
possible to measure the temperature of the oil exiting the windings, this would improve the thermal
modelling substantially: much more so for OF than ON or OD cooling.

C.1.3 Behaviour of OFAF cooled transformers at pumps stop

Two OFAF cooled 3-phase transformers equipped with hotspot sensors in the top of the windings were
run to steady state at the load current 1.0 pu. At this stage all cooling equipment was switched off and the
load current was kept constant at 1.0 pu. The hotspot temperature started to increase fairly linearly as a
function of time. For the first unit (22.5 MVA) it took about 170 min for the hotspot temperature to
185
Annex C: Examples of fiber-optic measurements

increase from 98 °C to 140 °C, i.e. the increase was 0.25 K/min [25]. For the second unit (31.5 MVA) the
increase from 90 °C to 130 °C took about 100 min., i.e. the increase was 0.40 K/min [26]. Obviously, the
hotspot increase is much faster for large units, since the only cooling surface after coolers switch-off is the
tank surface and the loss dissipation per m2 of tank surface is much higher. A corresponding test made on
a 3-phase OFAF-cooled 750 MVA auto-connected unit yielded a hotspot increase rate of 1.2 K/min [27].

C.2 Influence of ambient temperature on hotspot temperature rise

Three single-phase shell form generator step up transformers equipped with fiber-optic probes were heat
run tested in different ambient temperatures. The units were identical and of the design 570 MVA,
405±2.47%/20 kV, ODAF-cooled. The results are shown in Figure C.2 [28].

Figure C.2: Temperature rise test measurements comparison between three identical transformers.

It can be assumed that the individual differences were marginal, i.e. that the different hotspot rises were
due to the different ambient temperatures.

It is natural that the hotspot to top oil gradient remained constant (40 K), because the units were OD-
cooled and the heat dissipation from the windings was not ruled by buoyancy. The lower hotspot rise in
higher ambient temperature was completely defined by the lower top oil rise due to lower viscosity of the
oil.

C.3 Hotspot response to step changes in the load current and cooling capacity

A single-phase 250/250/30 MVA, 400/230-138-110/33-26.4-24 kV ODAF cooled auto-connected


transformer was tested in such a way that load current and cooling capacity were stepwise changed. The
hotspot response to those changes is shown in Figure C.3 [28].

186
Annex C: Examples of fiber-optic measurements

Figure C.3: Winding hotspot behaviour at different load ratings and cooling conditions.

During the linear part of the curve after cooling equipment switch-off at 9:30, the hotspot increase is about
6 K/min. when the load current is 1.0 pu. This information can be used also by transformer users to define
the response time in case of loss of auxiliary power supply.

C.4 Different hotspot factors at different load levels of a transformer

As a last example of the application of fiber optic sensors for research purposes is given hotspot factors
derived from thermal tests at different load levels applied on the same transformer without changing the
cooling capacity. The different factor, derived from data in Table 1 and Table 2 of [15], are summarized in
Table C.4.

187
Annex C: Examples of fiber-optic measurements

Table C.4: Hotspot factors at different load currents.

Oil flow in Winding Hotspot to top Hotspot factor


Transformer Load current %
windings gradient / K oil gradient / K H
70 15.1; LV 19.1; LV 1.26
25 MVA
Axial 100 19.6; LV 27.8; LV 1.42
ONAF
120 23.0 LV 32.6; LV 1.42
400 MVA 100 19.7; LV 20.7; LV 1.05
Axial
ONAF 139 28.5; LV 30.5; LV 1.07
400 MVA 100 18.7; LV 20.1; LV 1.07
Axial
ONAF 140 29.2; LV 26.5; LV 0.91
14.5; LV 18.0; LV 1.24
24.6; LV 35.1; LV 1.43
250 MVA 100
Zig-Zag
ONAF 149
12.5; HV 11.7; HV 0.94
22.6; HV 23.1; HV 1.02
15.6; LV 21.2; LV 1.36
18.6; LV 23.6; LV 1.27
100 23.4; LV 28.2; LV 1.21
400 MVA
Zig-Zag 129
ONAF
160 15.5; HV 18.2; HV 1.17
21.2; HV 25.2; HV 1.19
29.8; HV 35.8; HV 1.20
1000 MVA 100 19.6; HV 38.3; HV 1.95
Axial
OFAF 120 23.9; HV 42.6; HV 1.78
70 11.9; HV 16.2; HV 1.36
63 MVA
Zig-Zag 100 20.8; HV 27.7; HV 1.33
OFAF
140 36.9; HV 49.3; HV 1.34
9.7; LV 11.7; LV 1.21
14.5; LV 16.5; LV 1.14
70 25.1; LV 20.5; LV 0.82
230 MVA
Zig-Zag 100
OFAF
130 9.4; HV 12.9; HV 1.37
15.0; HV 20.3; HV 1.35
21.0; HV 28.1; HV 1.34
11.7; LV 21.3; LV 1.82
15.2; LV 34.3; LV 2.26
605 MVA 100
Zig-Zag
OFAF 130
23.0; HV 30.1; HV 1.31
35.2; HV 51.2; HV 1.45

By definition, the Q factor is unchanged at different load levels. From the values in Table C.4 it is obvious
that the hotspot factor H changes significantly for certain windings for all cooling modes. This is a good
example of the thermal scatter and the contribution of the S factor to the global H factor.

188

You might also like