Download as pdf or txt
Download as pdf or txt
You are on page 1of 131

D1

Materials and emerging


test techniques

Field grading in electrical


insulation systems

Reference: 794

March 2020
Field grading in electrical
insulation systems
WG D1.56

Members

V. HINRICHSEN, Convenor DE D. BACHELLERIE FR


J. DAS US L. DONZEL CH
M. HADDAD UK M. HAGEMEISTER CH
N. HAYAKAWA JP S. JOSEFSSON NO
I. JOVANOVIC US L. KEHL DE
M. KOCH DE M. KOZAKO JP
J. LAMBRECHT DE F. PERROT UK
C. STAUBACH DE J. WEIDNER DE
N. ZEBOUCHI FR M. H. ZINK DE

Young members
R. HUSSAIN DE M. SECKLEHNER AT

Corresponding Member
D. TABAKOVIC US

Copyright © 2020
“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”.

WG XX.XXpany 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
ISBN :warranties and conditions are excluded to the maximum extent
978-2-85873-496-2
permitted by law”.
TB 794 - Field grading in electrical insulation systems

Executive summary
High-voltage equipment like bushings and cable accessories require field grading strategies to control
and redistribute the electric field more uniformly in the material and at the interfaces of the insulation.
Different approaches provide geometrical grading (conductive layers in the insulation material),
refractive grading (high permittivity layers in the insulation), resistive grading (semiconductive layers
on the surface of insulation) or combinations of them. Semiconductive materials with linear or nonlinear
conductance, like carbon black or silicon carbide (SiC), are currently used as fillers for the purpose of
field grading. Alternating three-phase system voltage levels and design requirements are emerging.
Direct voltage applications with mixed capacitive and resistive field stress require special considerations.
New materials like microvaristors open new possibilities for the design of bushings, cable accessories
etc., but also for electrical insulation systems in other equipment, such as long rod insulators and
rotating electrical machines. The design of a field grading system is a rather complex task, requiring
knowledge about the fillers, the matrix materials and their interaction, which often cannot be predicted
from the properties of the individual components.
The purpose of this Technical Brochure is
 to inform about the fundamentals of field grading,
 to examine today’s different practices in detail and to give an outlook on possible future
developments,
 to inform about material characteristics (fillers, matrix materials), and
 to disseminate a common understanding of how to optimize electric field distribution and grading
of insulation systems and characterize the individual components.
The present Technical brochure summarizes the work carried out by the Working Group D1.56 and is
reported in seven chapters.
Following an “Introduction” in Chapter 1, Chapter 2 “Basics of electric field grading” gives the
fundamentals of electric field distribution in media. In particular, it introduces the concept of refraction
at interfaces between materials with different permittivities and/or conductivities. A classification of the
different field grading concepts is given and each concept explained.
Chapter 3 “Material systems for field grading systems” starts with a description of the physical
phenomena “conductivity” and “permittivity”. A majority of materials used in field grading are composite
materials. In this chapter, first, a description of typical matrix materials is presented, followed by a
comprehensive list of fillers known from products and literature. Then, the topic of measurements and
characterization methods for field grading materials is addressed. Especially for DC systems, design
rules and a design example are finally given.
Chapter 4 “Simulation of field grading systems” presents the different methods that can be used to
design and optimize field grading systems. Historically, transient network analysis has been used and is
still helpful in special cases. However, nowadays numerical field simulations with the finite element
method (FEM) or the finite difference method (FDM) are more widely used. Experimental possibilities
to validate the simulation results are presented.
The intention of Chapter 5 “Applications” is to give a detailed description of the main electrical devices
relying on field grading to function. First, cable accessories for both medium- and high-voltage systems
are addressed. Secondly, possible designs for terminations, joints and separable connectors (plug-
in systems) are presented. The third section deals with the special requirements of transformers. The
different types of bushings are discussed in Section 5.4. Section 5.5 is dedicated to the end corona
protection of rotating electrical machines that are either connected directly to the line or converter fed.
Finally, the subject of insulators (line insulators, hollow core insulators, insulators for gas insulated
systems) is treated.
After Chapter 6 “Summary and Conclusions”, which summarizes the content of the brochure and gives
an outlook to future needs and developments, Chapter 7 “Terms and Definitions” lists all symbols, units
and abbreviations that are in use or have been used in this Technical Brochure.

3
TB 794 - Field grading in electrical insulation systems

The Technical Brochure ends with Chapter 8 “References”, a list of the literature that was used to
prepare this Technical Brochure. It comprises some 200 publications and might be helpful for all who
need more comprehensive information on particular field grading issues.

4
TB 794 - Field grading in electrical insulation systems

Contents
Executive summary ............................................................................................................. 3

1. Introduction................................................................................................................ 7

2. Basics of electric field grading ................................................................................. 9


2.1 Capacitive and resistive electric field distributions ............................................................................... 9
2.1.1 Basics .................................................................................................................................................. 9
2.1.2 Fields at interfaces of layered materials ............................................................................................ 10
2.2 Electric field grading concepts .............................................................................................................. 11
2.2.1 General .............................................................................................................................................. 11
2.2.2 Bulk field grading ............................................................................................................................... 12
2.2.3 Surface field grading .......................................................................................................................... 16
2.2.4 Field grading in coaxial cylinder configurations by layered dielectrics ............................................... 21

3. Material systems for field grading systems ........................................................... 27


3.1 General ..................................................................................................................................................... 27
3.1.1 Conductivity ....................................................................................................................................... 27
3.1.2 Permittivity ......................................................................................................................................... 28
3.2 Matrix materials ....................................................................................................................................... 31
3.2.1 Glass (resistive glazing) ..................................................................................................................... 31
3.2.2 Thermoplastic polymers..................................................................................................................... 32
3.2.3 Thermoset polymers .................................................................................................................................. 32
3.2.4 Elastomers ......................................................................................................................................... 33
3.3 Fillers ........................................................................................................................................................ 34
3.3.1 Goal ................................................................................................................................................... 34
3.3.2 Carbon black and carbon nanotubes ................................................................................................. 34
3.3.3 Nonlinear materials ............................................................................................................................ 37
3.3.4 Further approaches under development ............................................................................................ 42
3.4 Measurement metrics of field grading materials and systems ............................................................ 47
3.4.1 Introduction ........................................................................................................................................ 47
3.4.2 Characterization of the polymer matrix .............................................................................................. 48
3.4.3 Characterization of the particulate filler.............................................................................................. 48
3.4.4 Characterization of the composite material ........................................................................................ 51
3.4.5 Electrical characterization of nonlinear field grading composite materials ......................................... 52
3.4.6 Operating stresses at component/equipment levels .......................................................................... 61

4. Simulation of field grading systems ....................................................................... 63


4.1 Purpose and requirements ..................................................................................................................... 63
4.2 Transient Network Analysis approach................................................................................................... 63
4.3 Numerical field simulation approaches ................................................................................................. 64
4.4 Verification of simulation results ........................................................................................................... 66
4.4.1 Measurement of potential and temperature distributions ................................................................... 66
4.4.2 Measurement of partial discharge (PD) inception voltages ................................................................ 69

5. Applications ............................................................................................................. 71
5.1 Cable accessories ................................................................................................................................... 71
5.1.1 General .............................................................................................................................................. 71
5.1.2 Accessories for MVAC and HVAC extruded cables ........................................................................... 75
5.1.3 Accessories for HVDC extruded cables ............................................................................................. 79
5.2 Plug-in systems / separable connectors ............................................................................................... 83
5.2.1 General .............................................................................................................................................. 83
5.2.2 Outer and inner cone systems ........................................................................................................... 84
5.2.3 Further examples of application ......................................................................................................... 86

5
TB 794 - Field grading in electrical insulation systems

5.3 Transformers ........................................................................................................................................... 86


5.3.1 General .............................................................................................................................................. 87
5.3.2 Design criteria for AC and impulse voltages ...................................................................................... 89
5.3.3 Design criteria for DC voltages .......................................................................................................... 90
5.3.4 Special field grading applications in transformers .............................................................................. 92
5.4 Bushings .................................................................................................................................................. 93
5.4.1 Basics of field grading within bushings .............................................................................................. 94
5.5 Rotating electrical machines .................................................................................................................. 97
5.5.1 General .............................................................................................................................................. 97
5.5.2 Direct line connection ........................................................................................................................ 98
5.5.3 Converter fed drives ........................................................................................................................ 100
5.6 Insulators ............................................................................................................................................... 101
5.6.1 Line insulators ................................................................................................................................. 101
5.6.2 Hollow core insulators...................................................................................................................... 104
5.6.3 GIS insulators .................................................................................................................................. 105

6. Summary and conclusions ....................................................................................109

7. Terms and definitions.............................................................................................114


7.1 Symbols and units ................................................................................................................................. 114
7.2 Acronyms ............................................................................................................................................... 116

8. References ..............................................................................................................120

6
TB 794 - Field grading in electrical insulation systems

1. Introduction
Electric field grading – sometimes also denoted as stress grading – refers to the active control of the
distribution of the electric field within or around a device. Especially apparatus for the transport and
distribution of electrical energy are facing huge potential differences between their terminals. This
cannot be avoided, but the electric field strengths must be kept below a level where partial discharges
and breakdowns within the insulating space would occur. Without a well-considered design, the voltage
and the resulting electric field is likely to not be distributed evenly across the insulating space or surface.
Since for high voltage applications one single weak or overstressed spot may decide over the
functionality and reliability of the whole given device, this has to be prevented.
Various different methods of electric field control can be implemented in the design process. Some are
seen as “good practice” or even basic knowledge for high voltage engineers, while others are subject
of ongoing research. They are all bundled together under the term “electric field grading” and will be
explained in this technical brochure.
Electric field control includes all measures that serve to lower local electric field strengths to such an
extent that the dielectric strength of the insulating materials and related interfaces are not exceeded
anywhere in the insulation system. In principle, a distinction is made between field control measures in
the volume of an insulating material and along interfaces. As the interfaces typically represent the
weakest points in an insulation system the latter is more important in most cases. Basic differences also
exist between time-varying field stress at alternating or surge voltages, constant field stress at steady-
state DC voltages or at superimposed transient stress in DC systems.
Regarding electric field control in insulating volumes, various fundamental measures for reducing electric
field stress are known:
 application of large-radii curvatures for electrodes, e.g. by shielding electrodes, large-diameter
conductors or chamfered electrode edges;
 optimization of radii ratios, e.g. in coaxial power cables, gas insulated systems or spherical capacitor
configurations;
 layering of insulating materials of different permittivities (for time-varying stress) or different
conductivities (for steady-state DC stress), e.g. by embedding a strongly curved electrode in a
high-permittivity material at alternating voltage or by the displacement of a DC field into a barrier
of low conductivity and high dielectric strength.
These measures are in most of the cases easy to implement, in contrast to the electric field control
along interfaces, which is typically a more challenging task. As long as interfaces are stressed by an
electric field in normal direction, i.e. perpendicular to the interface, field control can easily be achieved
by just controlling the maximum field strengths within the different layers. This is simply done by
appropriate choices of their permittivities and/or conductivities. When interfaces are also stressed by
the tangential electric field, i.e. parallel to the interface, the often lower electrical strength along the
interface is detrimental. The dielectric strength of an interface is typically lower than it would be
expected from the two adjacent media due to microscopic and macroscopic field displacement effects,
which can lead to local increases of the electrical stress. The formation of electron avalanches is
facilitated by inhomogeneities of the material structure, and the propagation of electrical discharges is
supported by the particular configuration of the insulation.
Interfaces and surface discharge configurations can occur between all types of insulating materials. The
most important combinations are solid-solid (e.g. a rubber stress cone of the cable termination on an
XLPE cable), solid-liquid (e.g. pressboard barriers in transformer oil), and solid-gaseous (e.g. the outer
surface of a bushing or a cable termination). Because of the low dielectric strength, the low permittivity
and the low conductivity of gases, special problems arise at interfaces between gases and solid or liquid
insulating materials, respectively. They are further intensified if pollution and water are present. These
interfaces are often referred to as surfaces. It belongs to the most important and frequently most
difficult tasks of high-voltage engineering either to avoid surface discharge arrangements (i.e. such with
significant tangential electrical stress) at all or at least to relieve them by adequate field control
measures. The following methods are available for this purpose:
 geometric field control by the geometric design of electrode contours;

7
TB 794 - Field grading in electrical insulation systems

 capacitive field control by conductive layers with capacitively determined voltage distribution (at
alternating or transient voltages);
 refractive field control by high-permittivity insulating materials (at alternating and surge voltage);
 resistive field control by semiconductive coatings with resistively determined voltage distribution;
 nonlinear resistive field control by materials that relieve themselves in areas of high electric field
stress by an increasing conductivity.
A need for electric field control is also given when potential and field distributions along insulator
surfaces are affected by stray capacitances to ground or adjacent live structures (at alternating and
transient voltage) or resistive surface currents (at direct voltage), e.g. for long insulators, surge arresters
or HVDC bushings.
Although it would have been most likely beneficial for some end users of field grading materials to base
their requirements on standards, no attempt was undertaken so far to standardize neither field grading
materials nor their components. A reason may be that the general requirements are too diversified to
establish a single standard being useful to a reasonable number of users.
In this Technical Brochure the basics of electric field control, the different materials used, the numerous
methods of characterization, simulation and verification approaches and typical applications are
presented and explained. It shall thus help improving the understanding of field grading systems and
recognizing the relevance of their individual components, and by this, give technical guidance to
designers as well as to users of field grading applications as long as standards do not exist. It may also
serve as an input to future standardization.

8
TB 794 - Field grading in electrical insulation systems

2. Basics of electric field grading


2.1 Capacitive and resistive electric field distributions
2.1.1 Basics
Free charge carriers within a defined space and exposed to an electric field will move along the electric
field lines within this space. The direction of the movement depends on the polarity of the charge and
the orientation of the electric field lines. The resulting current density ⃗⃗𝐽c of the moving charge is defined
by the electric field 𝐸⃗ and the electric conductivity 𝜎 of the material:
⃗⃗𝐽c = 𝜎𝐸⃗
Equation 2.1

This type of electric current is basically relevant to both, AC and DC, and therefore also to mixed electric
fields. However, for AC fields another current, the displacement current, is usually dominant.
Maxwell introduced the displacement current when he derived the electromagnetic wave equations. The
total current density 𝐽 then is:
𝐽 = ⃗⃗𝐽c + ⃗⃗𝐽d
Equation 2.2

where ⃗⃗𝐽c represents the conduction current by moving charges, and ⃗⃗⃗
𝐽d represents the displacement
current caused by the time-dependent electric field.
The displacement current density ⃗⃗⃗
𝐽d is defined by the permittivity 𝜀 of the material and the rate of
∂𝐸⃗
change of the electric field :
∂𝑡

𝜕𝐸⃗ ⃗
𝜕𝐷
⃗⃗𝐽d = 𝜀 =
𝜕𝑡 𝜕𝑡
Equation 2.3

provided that the permittivity 𝜀 is time-invariant. For rather high frequencies permittivity will decrease
(see e.g. Figure 3.1 and Figure 3.2).
Electric conductivity 𝜎 and permittivity 𝜀 both describe material properties related to the response of
materials to applied electric fields. On the basis of the formula for conduction current density ⃗⃗𝐽c the
permittivity 𝜀 is sometimes also called the “dielectric conductivity”, which underlines the similarity
between conduction and displacement current.
Traditionally, permittivity 𝜀 of a material is specified as a multiple of the permittivity of the vacuum ε0 .
The factor is called the relative permittivity 𝜀r .
So the total current density 𝐽 becomes:
𝜕𝐸⃗
𝐽 = ⃗⃗𝐽c + ⃗⃗𝐽d = 𝜎𝐸⃗ + 𝜀r 𝜀0
𝜕𝑡
Equation 2.4

In general the total current field should be considered when AC electric fields are applied to insulating
materials. However, the displacement current ⃗⃗⃗ 𝐽d usually dominates this case, and therefore the
conduction current ⃗⃗𝐽c is often neglected.
For pure DC electric fields only the conduction current ⃗⃗𝐽c needs to be considered as far as steady state
conditions are concerned. However, transitions between different steady state conditions, like polarity
reversal etc., may take significant time and should, therefore, be carefully considered.
In a homogeneous isotropic material of well defined linear electric conductivity 𝜎 and permittivity
𝜀𝑟 respectively, both current fields, ⃗⃗⃗
𝐽d and ⃗⃗𝐽c are oriented in the same way. Further, in case of a purely
sinusoidal time-depedency of the electric field 𝐸⃗ with an angular frequency of 𝜔 and no DC bias the
time-dependent electric field may be expressed as a complex quantity

9
TB 794 - Field grading in electrical insulation systems

𝐸⃗ = 𝐸⃗ ∙ ej(𝜔𝑡+𝜑)
Equation 2.5
∂𝐸⃗
with 𝐸⃗ representing the amplitude and direction of the sinusoidal field. The rate of change is:
∂𝑡

𝜕𝐸⃗
= 𝑗𝜔𝐸⃗ ∙ ej(𝜔𝑡+𝜑) = 𝑗𝜔𝐸⃗
𝜕𝑡
Equation 2.6

The total current 𝐽 can then be expressed in complex quantities, that is:
𝐽 = ⃗⃗𝐽c + ⃗⃗𝐽d = 𝜎𝐸⃗ + 𝑗𝜔𝜀𝐸⃗ = 𝐸⃗ (𝜎 + 𝑗𝜔𝜀)
Equation 2.7

Both current amplitudes, 𝐽c and 𝐽d , are proportional to the amplitude 𝐸 of the electric field under these
conditions. It has to be kept in mind, however, that the displacement current ⃗⃗⃗ 𝐽d is not in phase with the
𝜋
conduction current ⃗⃗𝐽c and the electric field 𝐸⃗ but phase-shifted by a time angle of . The amplitude 𝐽 of
2
the total currrent density 𝐽, therefore, calculates to

𝐽 = √𝐽c2 + 𝐽d2 = 𝐸 √𝜎 2 + 𝜔 2 𝜀 2

Equation 2.8

The phase angle between the total current 𝐽 and the displacement current ⃗⃗𝐽d is called loss angle 𝛿. It is
typically used to characterize the ratio between power losses P and reactive power Q at a specified and
constant frequency.
𝑃
tan δ =
𝑄
Equation 2.9

tan 𝛿 is also called “dielectric dissipation factor” or “dielectric power factor”.


Apart from the losses caused by the conduction current, dielectrics also show frequency dependent
polarization losses. The total dielectric dissipation factor tan 𝛿 is then represented by the sum of both
contributing components, tan 𝛿c due to conduction, and tan 𝛿pol due to polarization:
tan 𝛿 = tan 𝛿c + tan 𝛿pol
Equation 2.10

The dissipation factor due to conduction can be expressed as


𝜎
tan 𝛿c =
𝜔𝜀0 𝜀r
Equation 2.11

The polarization losses can be characterized by introducing the complex permittivity 𝜀r∗ :
𝜀r∗ = 𝜀r′ + 𝑗𝜀r′′
Equation 2.12

where 𝜀r′ represents the earlier introduced relative permittivity 𝜀r . With this definition the dielectric
dissipation factor tan 𝛿pol of the polarization becomes:
𝜀r′′
tan 𝛿pol =
𝜀r′
Equation 2.13

2.1.2 Fields at interfaces of layered materials


As long as electric interfaces between layered materials are stressed in normal direction (perpendicular
to the interface between them), the electric field can be described by comparatively simple equations.

10
TB 794 - Field grading in electrical insulation systems

As soon, however, as tangential components of the field exist, the typically lower dielectric strength of
the interfaces and the refraction of field lines have to be considered.
Electric fields at interfaces of dielectrics of different permittivities 𝜀𝑟 and/or conductivities 𝜎 will be
refracted depending on the ratios of the mentioned parameters. As long as either permittivity or
conductivity are the dominating parameters in both dielectrics the refraction can be easily approximated.
The refraction index is defined as:
tan 𝛼1 𝜀𝑟 1
=
tan 𝛼2 𝜀𝑟2
Equation 2.14

or
tan 𝛼1 𝜎1
=
tan 𝛼2 𝜎2
Equation 2.15

depending on the dominating parameter 𝜀𝑟 or 𝜎 of the applied stress (DC or AC). Note that in these
equations 𝛼1 and 𝛼2 are the angles between the electric field vectors and the normal vectors of the
boundary surface, see Figure 2.1.
In case of dielectrics with mixed and different dominating properties the refraction becomes more
difficult to describe, since accumulating charges trapped at the interface need to be considered as well.

Figure 2.1: Tangential and normal electric field components at the interface between two dielectrics [Küchler 2017]

2.2 Electric field grading concepts


2.2.1 General
Electric field grading aims at reaching a more or less homogeneous field distribution in the bulk or at
the interfaces of an electric insulation. By the application of field grading measures it is possible to
increase the operating voltage or to reduce the volume of the insulation as the insulating material can
be used in a more efficient way. Basically, there are three different field grading concepts, which can
be further sub-divided depending on the type of electrical stress (DC, AC, or transient) and the main
parameters influencing the field grading properties.
Table 2.1 provides an overview of these concepts. It is extended by two more concepts, based on the
basic three ones. For the extended concepts material properties are no longer assumed to be linear nor
isotropic nor homogeneous.
It can further be distinguished between two basic modes of operation: Surface field grading concepts
aim at defining a smoothly changing potential distribution along a surface, whereas bulk field grading
concepts involve the total volume of the relevant insulating materials. Although functionally graded
materials and nonlinear materials applied in the bulk volume will affect the electric field at the interfaces,
this approach is considered as bulk field grading concept in this context.

11
TB 794 - Field grading in electrical insulation systems

Table 2.1: Overview of electric field grading concepts (red dashed lines = field lines, blue dashed lines = equipotential
lines)

Significant parameter
for

Concept DC AC

Geometry (bulk field grading)

shape of shape of
electrode and electrode and
total current total current
density along density along
electrodes electrodes

Potential grading (capacitive/resistive field grading, bulk field grading)

conductivity of permittivity of
layer between layer between
adjacent adjacent
grading foils grading foils

Potential grading (surface field grading)

conductivity
conductivity &
permittivity

Nonlinear materials with field dependent conductivity/permittivity permittivity =


conductivity
(in combination with any of the other concepts) f(E)
=
conductivity =
f(E)
f(E)

Continuous or stepwise discrete functionally graded materials permittivity =


(in combination with any of the other concepts) conductivity f(x, y, z, E, ϑ,
= …)
f(x ,y, z, E, ϑ, conductivity =
…) f(x, y, z, E, ϑ,
…)

2.2.2 Bulk field grading


These concepts aim at defining equipotential surfaces in a way that the field stress within the dielectric
located between those surfaces remains below critical levels. The simplest kind of this concept consists
of two, either metallic or metallized surfaces, where both are usually connected to defined levels, i.e.

12
TB 794 - Field grading in electrical insulation systems

high voltage and ground, respectively. The (linear) electrical properties of the dielectric are not relevant
to the field distribution between these equipotential surfaces.
In addition to these two equipotential surfaces with fixed potential, further, floating equipotential
surfaces can be introduced. Their potential is defined by the capacitance/resistance between the
adjacent equipotential surfaces and the total capacitance/resistance between the grounded and
energized surfaces (principle of voltage divider). The (linear) electrical properties of the specific
dielectric are again not relevant to the field distribution within the region between the adjacent
equipotential surfaces. However, the properties influence the level of floating potential of the additional
surfaces, and by selecting a dielectric with higher permittivity/conductivity for a highly stressed region,
the maximum field stress in that specific region can be reduced.
The principle of layering dielectrics with different properties can also be applied without the introduction
of additional metallic or metallized equipotential surfaces. However, the shape of the individual layers
needs to be well defined since the refraction of field lines at the interface between the layers may have
an additional impact on the field distribution where two or more interfaces intersect or converge (triple
point effect).

2.2.2.1 Geometric field grading


Provided that there is enough space available within an electric apparatus it is possible to influence the
electric field by introducing additional conductive surfaces, which are connected to well defined
potentials (typically ground or high voltage). These electrodes guide the equipotential surfaces such
that the corresponding dielectric stress remains within a tolerable limit. This way of field grading is
called geometric, since suitable geometries are used to influence the electric field. An example for
geometric field grading is shown in
Figure 2.2. In this case of a gas-filled bushing field grading must be achieved not only inside the bushing,
where an insulation gas of high electric strength is used, but also at the outer insulator surface.
Particularly position and shape of electrode (6) must, therefore, be carefully chosen.

Figure 2.2: Gas-filled bushing as an example for geometric field grading [HSP]

13
TB 794 - Field grading in electrical insulation systems

Another example of geometric field grading is the stress control1 element typically used in terminations
for solid dielectric cables, provided in Figure 2.3.

Figure 2.3: Typical field control element (“stress cone”) for solid dielectric cables (left) and distribution of
equipotential lines in the cable termination (right) [G&W Electric Company]

The ground electrode (“deflector”) provides extension of the power cable insulation screen and is
geometrically shaped to distribute the electrical stress as evenly as possible from the edge of cable
screen to the point where the deflector ends. The regions of highest electrical stress are embedded in
insulating material with high electric strength, such as silicone rubber or EPDM in case of solid dieletric
cables, or oil-impregnated paper in case of laminar insulated cables.
An example of geometric field control for laminar dielectric systems is provided in Figure 2.4. Here the
same principle is used as in the case of solid dielectric cables, but the “ideal” geometry of the electrode
is determined by calculation of the longitudinal component of the electrical stress that is critical for
laminar (layered) dielectric systems such as oil-impregnated paper insulation.

Figure 2.4: Example of geometric field grading in laminar dielectric systems, in this case oil-impregnated power cable
[G&W Electric Company]

Geometric field grading is in principle applicable to AC and DC electric fields. However, it needs to be
considered that in case of layered dielectrics between the electrodes any refraction of field lines at the
interfaces depends on the type of field and may change the relevant electric field significantly. At AC
and transient stress the refraction is controlled mainly but not exclusively by the ratios of the
permittivities whereas at DC the ratios of the electric conductivities become the only relevant parameter
[Hinrichsen 2011]. Hence in the DC case it has to be taken into account that conductivity of a material
is strongly dependent on temperature and also field strength, which makes field control rather
challenging for high DC voltages, refer e.g. to [Wirth 2015].

1
As mentioned earlier, the term “stress grading” is often used synonymously for “field grading”. This is especially the case for
cable accessories.

14
TB 794 - Field grading in electrical insulation systems

2.2.2.2 Capacitive field grading (Condenser field grading)


Capacitive (condenser) field grading can be applied if available space is limited or if size and hence
weight shall be kept low. This type of field grading is not limited to AC or impulse electric fields, where
the permittivity of the material defines the capacitances between the grading elements. It can be applied
to DC as well, were the conductivity of the material defines resistive grading elements (see Figure 2.5).
Further, the condenser type field grading is usually limited to rotational-symmetric apparatus for
manufacturing related reasons.
As in the geometric field grading concept additional equipotential surfaces (electrodes) are introduced
to the dielectric. In contrast to the geometric field grading, the electric potential of these additional
electrodes is not fixed but floating and will take a level, which ensures that the electric field in all layers
between adjacent electrodes (and in their proximity) remains below critical limits. The potential
difference between adjacent grading electrodes is given by the voltage divider ratios of the impedances
between the layers (see Figure 2.5). E.g., by changing the lengths of the cylindrical electrodes of a
condenser type bushing (see Figure 2.6) the individual capacitances can be adjusted as needed, but it
has to be kept in mind that the change of a single electrode also impacts the potential difference of the
other layers, which typically requires an iterative approach to find the optimal design. An example for
the radial field distribution in an optimized condenser bushing is shown in Figure 2.7. Further information
is given in 5.4.
In order to achieve identical capacitances between the individual electrodes, the lengths of the
electrodes usually change from inside to outside of the cylindrical structure with the longer one being
located closer to the center of the cylindrical structure. This means that a portion of the electrode is
exposed to the surrounding electric field and can, therefore, also be used as a geometric field grading
electrode. This feature is actually used with air-to-oil and air-to-SF6 condenser bushings, where the
exposed length is larger at the air side of the bushing to provide a larger flashover distance in air
compared to the other, electrically stronger side (Figure 2.6). Hence the capacitive field grading does
provide both, a radial (Figure 2.7) and an axial field grading on the surface of the bushing envelopes,
which can be seen at the equipotential lines shown on the right side of the bushing in Figure 2.6.

Figure 2.5: Equivalent circuit of a condenser type field grading arrangement

Figure 2.6: Condenser type bushing with asymmetric arrangement of electrodes for additional boundary field grading;
acc. to [Küchler 2017]

15
TB 794 - Field grading in electrical insulation systems

Figure 2.7: Typical radial field stress within a cylindrical condenser type bushing with homogeneous dielectric
properties

Since the permittivity of the material used in condenser bushings for AC application is dominating over
its resistivity, the voltage drop across an individual layer between two adjacent electrodes is bascially
not influenced by the frequency of the applied electric field. Therefore, this type of field grading is also
suitable for transient electric fields (impulse voltages). Special care should be taken for very fast
transients because they can be split according to the different surge impedances inside and outside the
dielectric. This is important especially for bushings in gas insulated switchgear (GIS), because the rate
of voltage change is exceptionally high in case of switching, and the dielectric can be overstressed.
Theoretically, there is another way of changing the capacitance of individual layers between the
electrodes (see Figure 2.14): Materials of different dielectric properties could be applied, which would
represent a stepwise functional field grading. However, for manufacturing-related reasons this may be
done only in very special cases. Even nonlinear field grading material could be used as well, but no
related product is known so far.

2.2.3 Surface field grading


This concept aims at defining a suitable potential distribution along a surface or interface between
different materials to prevent surface discharges. Theoretically, the thickness of the layer controlling
the potential distribution could be infinitesimal. In practice layer thickness varies, which may affect the
efficiency of the layer at the interface. However, in most cases the field distribution can be approximated
with sufficient accuracy by assuming surface properties only.
2.2.3.1 Creeping discharge configurations
Surface discharges develop along the boundary between two insulating materials with different states
of matter e.g. along the surface of a solid insulator kept in air. The main conditions responsible for this
phenomenon are electric field lines mainly in parallel to the surface and strong capacitive coupling of
conductive surface layers with the counter electrode. Typical examples of such configurations are
bushings, cable ends, or the winding terminations of electrical machines. A classical model used for the
investigation of surface discharges is the Toepler's configuration [Toepler 1921], which helps
understanding the phenomena. As shown in Figure 2.8, it consists of a simple glass plate connected to
a rod electrode at HV potential. The ground electrode is formed by a metal surface on the back side of
the glass plate.

Figure 2.8: Surface discharge configuration according to Toepler: experimental setup (left) and cross section
depicting surface discharges and surface capacitances (right)

Characteristical for such a configuration is that no direct puncture (breakdown) between the electrodes
will occur due to the high electric strength of the insulator between the electrodes. Nevertheless, such
kind of discharges are quite a strong load for insulating materials, as they can erode the material and
lead to a breakdown after a certain time. These discharges will occur mainly along the insulator surface,

16
TB 794 - Field grading in electrical insulation systems

fed by capacitive displacement currents through the insulator. The development of breakdown decisively
depends upon the specific surface capacitance of the dielectric. Surface discharges can form only when
this capacitance has a certain minimum value. Another precondition is that pre-discharges develop at
one of the electrodes. Prevention of these pre-discharges is the most effective measure to prevent
surface discharges.

2.2.3.2 Development of surface discharges


The inception voltage U i especially of the above stated Toepler's configuration depends on the thickness
s of the electrode as:

𝑈i = 𝐾 ∙ √𝑠⁄𝜀r

Equation 2.16

with U i in kV and s in cm.


The factor K can be calculated theoretically. But many additional factors (such as the shape of the
electrode or the surface insulation resistance) have an influence. Thus it is appropriate to use the
empirical values determined for K, listed in Table 2.2:

Table 2.2: K values for different configurations; acc. to [Böning 1953]

Configuration K
in air 8
Metal edge
in SF6 21
Metal or graphite edge in oil 30
Graphite edge in air 12

If the voltage is increased above the inception voltage, streamer discharges occur. In the Toepler's
configuration, these streamers spread out radially around the center of the electrode, whereas in a
coaxial cable arrangement (e.g. at cable ends) these streamers spread out in axial direction along the
dielectric. This is shown in Figure 2.9. In principle, this configuration is also found e.g. on the winding
rods of rotating electrical machines, where they leave the stator.

Figure 2.9: Appearance of surface discharges in Toepler’s configuration (left) and at a


cable end (right; acc. to [Küchler 2017])

The surface discharges on the insulator surface form a capacitance with the counter electrode. The
value of the specific surface capacitance csurface of the covered surface Asurface is given by:

17
TB 794 - Field grading in electrical insulation systems

Csurface  0   r
csurface  
Asurface s
Equation 2.17

where Csurface represents the surface capacitance and s the thickness of the dielectric.
The current in the discharge channel can be expressed as follows:
dQsurface d(usurface  Csurface )
ichannel  
dt dt
Equation 2.18

The voltage usurface is thus the voltage between the considered location of the insulator's surface and the
counter electrode. In a first approximation, it is same as the voltage u applied across the electrodes, so
that finally, using the product rule, for the applied current in the discharge channel the following holds:
du dCsurface
ichannel  Csurface  u
dt dt
Equation 2.19

For alternating voltages, it can be assumed that the voltage in the interesting time intervals ( 10 ns)
is constant (i.e. du/dt = 0). This leads to:
dCsurface dA
ichannel  u   u  csurface  surface
dt dt
Equation 2.20

Thus, besides voltage, the discharge current is dependent upon the specific surface capacitance and
the rate of change of the surface area covered by surface charges.
However, for impulse voltages, the term du/dt also contributes significantly and cannot be ignored.
Hence the surface discharge is more intense with impulse voltages than with alternating voltages.
Similarly, for direct voltages, both the derivative terms result in zero values. Theoretically there should
thus not be any problem of surface discharges in case of DC configurations. However, in reality, there
exists a problem of discharges (e.g. bushings or terminations of DC cables) because of non-zero du/dt
high frequency components superimposed to the DC voltage as soon as pre-discharges develop. Such
pre-discharges are formed due to the low electric strength of the ambient media (e.g. air) covering the
surface of the solid insulation.
Already streamer lengths of a few centimetres will result in thermal ionization because of the high
displacement currents due to the high specific surface capacitance. As a result, the initial streamer
discharges slowly transit into leader discharges even at such small distances of around few centimetres.
(It has to be noted that, in a conventional rod-plane arrangement in air, at least one to two meters of
discharge length is required for leader discharges.)
Leader discharges have low voltage demands because of their negative differential resistance
characteristics (decreasing voltage with increasing current). As a comparison:
 specific flashover voltage of a post insulator in air (peak value):  5 kV/cm
 inception voltage of a streamer surface discharge (peak value):  6 kV/cm
 field intensity for surface leader propagation (peak value): = a few 100 V/cm
Once the streamer surface discharge transforms into a leader discharge, it requires only a slight further
increase in voltage to initiate a flashover. Hence, in the very first place, voltages equal to or above the
surface discharge inception voltage Ui must be avoided in practical configurations. An empirical formula
for Ui in air is as below:
13.5 ∙ 10−5
𝑈𝑖 = 0.44
𝑐surface
Equation 2.21

with U i given as the r.m.s. value of voltage in kV and csurface in F/cm².

18
TB 794 - Field grading in electrical insulation systems

In [Werdelmann 2009] an inception field strength of 0.64 kV/mm is determined for a clean surface in
air. This reference also provides measurement values for different kinds of surface pollution. In [Marusic
1994] [Wheeler 2005] an inception field strength of 0.6 kV/mm is stated, and [Baumann 2011] discusses
values between 0.4 – 0.5 kV/mm. In a theoretical approach [Thienpont 1964] derives an equation for
inception voltage of glow and brush discharges at the slot exits of generator stator windings and
determines a field strength of 0.39 – 0.41 kV/mm.

2.2.3.3 Measures to avoid surface discharges


To understand, how surface discharges can be avoided, it is necessary to recall that surface discharges
are discharges in the gaseous phase. These discharges creep along the surface in the described manner,
and just increasing the length of the surface will not be conducive to avoid or stop the discharges. So,
as mentioned already, prevention of pre-discharges is most important. This can be achieved using
advanced measures of field control (like using embedded electrodes in the insulating material,
metallization of the insulator at the interfaces to the flanges of bushings or capacitive grading like in
bushings). Another frequently used counter-measure is resistive field control by using a weakly
conductive layer on the insulating surface to limit the power available for the development of a leader
discharge, which is exemplarily and principally shown in Figure 2.10 and is explained in detail in 2.2.3.5.
In many cases, such as at the winding terminations of electrical machines, not many other measures
do exist for suppressing surface discharges.

Figure 2.10: Example of a counter measure against surface discharges: resistive field grading
[Hinrichsen 2011]

2.2.3.4 Refractive field grading


The least space is typically required for refractive field grading. This type of field grading is mainly used
to avoid surface discharges along tangentially stressed interfaces (see Figure 2.11), like the cable
insulation at the end of the semicon shield within cable terminations or the end corona protection of
stator windings. The name “refractive” reflects that the field lines at the interface between two different
materials having different electric properties (permittivity and conductivity) are refracted (see also
2.1.2). Although the refraction of the electric field is not limited to interfaces between two mainly
capacitive materials, the name refractive field grading is commonly used for this combination of
materials only. The field grading for capacitive-to-resistive or resistive-to-resistive interfaces is usually
called “resistive”, but it should be kept in mind that the basic concept is the same.
Due to naming conventions the term “refractive field grading” is used for AC electric field grading only,
though the physical effect is also present in the DC case.
Concept of this type of field grading is to provide a spatially distributed voltage divider (lattice network)
along the electrically highly stressed interface. A simplified version of the divider is shown in Figure
2.11. This model can be applied provided the thickness of the field grading layer is small compared to
the thickness of the remaining insulation layer. For thick field grading layers permittivity and conductivity
perpendicular to the interface have to be taken into account additionally. The equivalent lattice network
may be extended in this case by another lattice network, representing the bulk of field grading material.
Efficiency of the field grading layer depends on the characteristic admittance (permittivity and
conductitivty) per unit length of the field grading layer versus the insulation layer. The higher the
difference the lower will be the electric stress along the interface. In order to keep the field grading
layer thin, high permittivity materials are preferred.

19
TB 794 - Field grading in electrical insulation systems

Figure 2.11: Thin mixed “refractive” & resistive (semiconducting) field grading and equivalent electric circuit (spatially
distributed mixed dividers); acc. to [Hinrichsen 2011] and [Küchler 2017]

The equivalent electric circuit suggests that in case of a mainly capacitive field grading material the
grading will not depend on frequency and therefore also applies to transient fields. However, typically
high permittivity materials are significantly frequency dependent. The effectiveness of a refractive field
grading system exposed to transient fields should therefore always be checked carefully against the
frequency dependance of the electric properties. For that reason it may be required to use larger
thicknesses of lower permittivity materials or to use a combination of geometrical and resistive field
grading as implemented in some cable joints, refer to 5.1.3.

2.2.3.5 Linear and nonlinear resistive field grading systems


As mentioned in the previous chapter, a resistive field grading is based on the same concept as the
refractive field grading. However, the conductivity of the field grading layer becomes the dominant
property, and therefore the thermal balance of the field grading systems needs to be considered
carefully.
As shown earlier, the electric field lines are refracted also in this case. The difference is that additionally
space charge accumulation at the interface has to be considered. Depending on the subjacent insulation
material the resistive field grading can be applied to AC or DC electric fields. For DC field grading the
subjacent insulation must exhibit a controlled conductivity in order to achieve a well-defined field
grading. This may represent a significant technical challenge since the conductivity is typically highly
influenced by temperature.
For certain service conditions like polarity reversal and transient electric stresses, also for DC applications
the permittivities of the involved materials may need to be coordinated. Otherwise, the field grading
system may fail under these conditions. In practice resistive field grading is, therefore, often
accompanied by refractive/capacitive or geometric field grading.
In order to improve the thermal balance or the response to transient electric fields, the refractive field
grading can be extended by using nonlinear field grading material (see Figure 2.12), which changes its
properties to become dominantly conductive when e.g. a certain field strength is exceeded. The losses
generated at low or continuous stresses can thus be reduced without affecting the field grading
performance at higher overall stresses including transients.
An example for such a field grading system are microvaristors, mixed into either elastomers, hot melt
compounds or thermosetting resins to allow for easy application. In addition to the nonlinear resistive
field grading, microvaristors also provide a suitable refractive field grading due to the high permittivity
of the varistor grains.

Figure 2.12: Thin nonlinear field grading and equivalent electric circuit (spatially distributed nonlinear dividers); acc. to
[Hinrichsen 2011]

In principle nonlinear materials are not limited to resistive field grading. A dielectric with field dependent,
increasing permittivity could also be utilized to improve a linear refractive field grading without the
negative impact of higher losses, which is typical for nonlinear resistive field grading. However, such

20
TB 794 - Field grading in electrical insulation systems

kind of material with sufficiently high nonlinear permittivity is currently not available, but anti-
ferroelectrica are possible candidates and under investigation, e.g. [FLAME 2019].
Current technology is based on materials with electric field-dependent conductivity. Their response to
a field stress exceeding a certain threshold is practically instantaneous. However, other indirect
dependencies like increased local temperature due to electric stress could theoretically be exploited as
well. They may, however, react slower to high stresses than the currently used materials.
Typically, a high level of nonlinearity is beneficial, since power losses will be low with currently used
materials, as long as the threshold is not exceeded. But when exceeded, the impact on field grading
performance is strong. However, a high level of nonlinearity also affects adjacent areas in several ways.
Firstly, the adjacent areas will experience a higher stress once the threshold is exceeded somewhere.
As long as the initial field distribution is distinctly inhomogeneous the stress will be gradually displaced
to adjacent areas. In case of a rather homogeneous field, where all regions are likely to simulaneously
exeed the threshold, this concept may not provide an improvement. Secondly, the electric field stress
in the adjacent areas will become non-sinusoidal, which may have a negative influence on the field
grading properties of the material. The higher the nonlinearity, the larger will be the distortion of the
electric stress in the adjacent areas.
Therefore, the design of a nonlinear field grading is much more difficult than it is for linear systems.
Even if the applied voltage is sinusoidal, the overall displacement and conduction currents and as a
result also the electric field may deviate significantly from a sinusoidal shape. The harmonic content of
the electric field should be taken into account within the field grading material itself but also in the
adjacent insulation since it potentially may cause higher power losses. In 3.4.5 an example of a design
procedure for nonlinear field grading material is given for the case of a DC application.

2.2.4 Field grading in coaxial cylinder configurations by layered dielectrics


2.2.4.1 General principle
Basically the electric field distribution can be optimized by chosing permittivity and geometry such that
maximum field strength values are kept below the specified ones. Especially in coaxial cylinder
configurations use can also be made of coaxially layered dielectrics. In each individual layer the electric
field decreases inversely proportional to the radius and makes a step upwards at the next layer if the
dielectric constant makes a step downwards, and by varying thickness and the dielectric constant of
each layer, taylored field control is possible. A basic configuration consiting of two different layers is
shown in Figure 2.13.

Figure 2.13: Coaxially layered dielectrics; r0…radius of inner conductor, r1….outer radius of layer 1, r2….outer
radius of layer 2, εr1 and εr2….dielectric constants of layers 1 and 2, respectively

The maximum field strength, En, max, in each layer n is found at its inner radius and can be expressed
as:
𝑈
𝐸1𝑚𝑎𝑥 = ∙𝐾
𝑟0 ∙ 𝜀𝑟1 𝑐
Equation 2.22
𝑈
𝐸2𝑚𝑎𝑥 = ∙𝐾
𝑟0 ∙ 𝜀𝑟2 𝑐

21
TB 794 - Field grading in electrical insulation systems

Equation 2.23

where Kc is a constant, determined by the radii and dielectric constants of the overall configuration of
N layers:
1
𝐾𝑐 = 1 𝑟𝑛
∑𝑁
𝑛=1 ( ∙ ln 𝑟 )
r𝑛 𝑛−1
Equation 2.24

The maximum field strengths, En,max, in each layer (in this case only two) are identical if r 0·εr1 = r1·εr2,
or in general if rn-1·εrn = const.
Following this principle, an ideally constant electric field can be achieved if a material is used that has
a dielectric “constant”, which continuously decreases proportional to 1/r (see Figure 2.14). This
approach has become known as “functionally graded materials” and is explained in detail in 2.2.4.2 and
5.6.3.2. In some practical applications, like bushings, the ratio between the absolute radius of a grading
foil and the distance to its neighboured one is so high, that the field in between can be regarded as
nearly constant.

Figure 2.14: Magnitude of electric field strength in a coaxial cylinder configuration; black: not layered, only one
homogeneous dielectric; red: four layers of dielectrics optimized such that all are exposed to the same maximum field
stress (by decreasing εr in four steps from the inner to the outer conductor); blue: functionally graded insulation
system with εr  1/r, which corresponds to an infinite number of grading layers; note: areas below the three curves are
identical

The same grading principle can also be applied to DC insulators in the emerging DC GIS and GIL (gas
insulated switchgear, gas insulated lines). For insulation layers of different electrical conductivities, the
same rules apply as for layers of different permittivities in the AC case. In the equations given above
just the dielectric constants have to be replaced by electric conductivities. Well controlled conductivities
in DC systems are much more important than well controlled permittivities in the AC case. This is due
to the fact that conductivity has a very strong dependance of temperature. Close to the inner conductor,
which may take temperatures up to 90 °C, conductivity of the insulators may be much more than one
order of magnitude higher than at the outer conductor, where the temperature can be significantly
lower. Consequently, the electric field strength is reduced close to the inner conductor and increased at
the outer conductor. In extreme cases the field stress at the outer conductor becomes higher than that
at the inner conductor (“field inversion”). This phenomenon is also well known for DC cables and their
accessories, or for DC bushings [Wirth 2015].

2.2.4.2 Application of functionally graded materials


Functionally graded materials are characterized by the spatial distribution of dielectric permittivity (r),
denoted here as “ε-field grading”, and/or electrical conductivity (), denoted here as “-field grading”,
of solid insulators in order to change these electrical properties e.g. inversely proportional to the distance
to the inner conductor [Kurimoto 2002] [Kurimoto 2010] [Ju 2009]. The concept of ε-field grading with
a graded permittivity distribution is shown in Figure 2.15. Compared with the uniform material in Figure
2.15 (a), the ε-field grading in Figure 2.15 (b) has a spatial grading of filler content from material-A to
material-B in a solid insulator. In gas/solid insulation systems, the electric field stress under AC or

22
TB 794 - Field grading in electrical insulation systems

impulse voltage application is generally intensified in the gas region, because the permittivity of gas is
lower than that of solid material. In order to relax the stress intensification, application of ε-field grading
is expected to be effective by giving a suitable permittivity distribution inside the solid insulator. Thus,
the intensified electric field stress in the limited region can be relaxed and homogenized in the wide
region, i.e. ε-field grading can drastically raise the electric field utilization factor in the insulating
materials.
Electric field stress
Material-A Material-B Low High

 HV electrode

Gas

GND electrode
(a) Uniform material
Electric field stress
Low High
 HV electrode

Gas

GND electrode
(b) -FGM

Figure 2.15: Concept of ε-field grading [Hayakawa 2014]

Figure 2.16 shows the fabrication concept of ε-field grading with grading to lower permittivity (GLP-
FGM) by using the centrifugal force. Two kinds of fillers, such as Al2O3, SiO2, TiO2, are uniformly mixed
into the uncured epoxy resin. One filler has a small particle diameter and high permittivity, and the
other one has a large diameter and low permittivity. By application of a centrifugal force, the large
diameter fillers (with low permittivity) move towards the direction of centrifugal force during the curing
process, whereas the small diameter fillers (with high permittivity) do not move under the centrifugal
force. But they can be pushed away by the large diameter fillers and, therefore, relatively move towards
the direction opposed to the centrifugal force. Once the epoxy resin is cured this distribution of filler
particles is “frozen”. As the result, a GLP-FGM with grading to lower permittivity along the centrifugal
direction can be obtained. Following the same principle, other kinds of ε-field grading with grading to
higher permittivity (GHP-FGM) or U-shape permittivity distribution (U-FGM) can be fabricated by
centrifugation [Hayakawa 2014].

23
TB 794 - Field grading in electrical insulation systems

Figure 2.16: Fabrication concept of GLP-FGM with grading to lower permittivity along the centrifugal direction
[Hayakawa 2014]

[Hayakawa 2014] shows a calculation model to simulate the filler particle movement in viscous fluid
(epoxy resin mixed with hardener) under the impact of a centrifugal force.

Figure 2.17: Calculation model of filler particle movement under centrifuging force [Hayakawa 2014]

Using a one dimensional model for the post-type ε-functionally graded materials spacer (Figure 2.17),
the flow of particle density due to the centrifugal force is calculated in each region. The balance of three
forces working on the filler particles, i.e. centrifugal force, buoyancy and drag force, is given by the
following equation:
d𝑣r 𝜌f
𝑀 = 𝑀𝑟𝜔2 − 𝑀𝑟𝜔2 − 𝐹D
d𝑡 𝜌p
Equation 2.25

where M is the mass of a filler particle, vr is the relative velocity of a filler particle, r is the radius of
rotation,  is the angular speed,  f and  p are the mass densities of fluid and particle, and FD is the
viscous force. Given dvr/dt = 0, the terminal velocity vt of filler particles is expressed as follows:
𝐺𝐷p2 (𝜌p − 𝜌f )
𝑣t =
18
Equation 2.26

where G is the acceleration by the centrifugal force, Dp is the diameter of a filler particle,  is the viscosity
of the epoxy resin mixed with filler particles.  is estimated and experimentally verified, considering
material, temperature, loading with filler, etc. The permittivity distribution of the ε-functionally graded
material can be calculated by the density distribution of the filler particles in the epoxy resin [Hayakawa
2016].

24
TB 794 - Field grading in electrical insulation systems

Using TiO2 and SiO2 with the specifications shown in Table 2.3, a sample of a GLP-FGM post-type
insulator was fabricated. The permittivity distribution was calculated and verified on the sample, with the
results shown in Figure 2.18 [Hayakawa 2016]. Permittivity distributions of GHP-FGM and U-FGM
samples were also verified by calculation and measurement. An arbitrary permittivity distribution of an
ε-functionally graded material is expected to be obtained by optimizing the specifications of fillers
(material, size, specific gravity, relative permittivity, loading, etc.) and the manufacturing conditions
(centrifugal acceleration, time, temperature, their combination pattern with time, etc.).
Table 2.3: Specifications of filler particles for ε-functionally graded material

Figure 2.18 : Permittivity distribution of post-type GLP-FGM (Filler loading: 10 vol% TiO2 and 40 vol% SiO2, centrifugal
conditon: 4 000 G and 60 minutes) [Hayakawa 2016]

An alternative approach to achieve “functional grading” especially in DC systems is to make use of the
electric field dependance of functional fillers that provide field dependent conductivity and only
moderate and well controlled temperature dependance. Typically, conductivity increases with the
applied electric field as well as with temperature, which is exactly the desired behavior (this would not
work with permittivity as it typically decreases with increasing electric field). The comparatively high
conductivity close to the inner conductor is achieved both by the electric field dependance and the
temperature dependance. The fillers may be based on ZnO microvaristors or on antimony doped tin
oxide (ATO) and possibly other materials in the future. They are filled into the polymer matrix at a
densitiy above the percolation threshold. Such systems, which are still under development, can be used
to optimize DC GIS/GIL insulation systems.
One practical approach of ATO fillers are MFF (“MFF” = Minatec® functional fillers; since 2015 named
lriotec 7000®) [Rüger 2012] [Greb 2015]. The electric bulk conductivity of an MFF filled epoxy is in the
range (10-13… 10-11 S/m) (Figure 2.20), which is an appropriate range for DC compact GIS insulators in
order to achieve the desired field distribution and low losses at the same time. An example of achievable
field distributions in a DC GIS insulator is shown in Figure 2.19.

25
TB 794 - Field grading in electrical insulation systems

A further benefit of the controlled conductivity is a faster decay of surface and volume charges, which
typically develop in DC insulation systems [Tenzer 2011] [Tenzer 2013c] [Tenzer 2015] [Winter 2011a]
[Secklehner 2013] [Winter 2014] [Winter 2015] [Secklehner 2015]. It is still an open question and under
investigation if such sophisticated approaches are necessary or if “conventional” insulating materials of
comparatively high intrinsic conductivity may be sufficient for HVDC applications.

Figure 2.20: Measured bulk conductivity of MFF filled epoxy [Winter 2014]

Figure 2.19: Normalized tangential electric field distribution at a disc-shaped DC GIS insulator surface for the two
cases: “hot” with a temperature gradient
(ϑconductor = 80 °C/ϑenclosure = 50 °C) and “cold” with constant low temperature
(ϑconductor = ϑenclosure = 30 °C); MFF = “Minatec® functional fillers” [Winter 2014]

26
TB 794 - Field grading in electrical insulation systems

3. Material systems for field grading systems


3.1 General
Most field grading materials used today as bulk material or coatings consist of a blend of different
unitary materials since the properties of single unitary materials usually do not meet all the requirements
set for modern field grading systems. The typical base material – the host matrix – is a polymer (resin,
elastomer, hot melt compound, etc.) into which particulate fillers like carbon black are homogeneously
dispersed. Inorganic, ceramic type fillers may also be used either together with carbon black or
independently. The host matrix is typically selected with the focus on the required mechanical properties
for easy processing and acts as a carrier and fixation for the filler. It usually encloses the particles almost
completely. The filler is typically selected to provide or improve the electrical properties of the host
matrix.
The effectiveness of a field grading material in general depends on its permittivity or conductivity,
respectively. Usually one of these properties is dominating with respect to field grading, although
sometimes a well balanced combined system has to be chosen or can provide some additional benefits.
In the following these two parameters are considered, and unitary materials are addressed first.

3.1.1 Conductivity
Depending on their level of conductivity, materials are typically classified into groups:
- Insulators
- Semiconductors
- Conductors
For solid unitary conductors and semiconductors the conduction mechanism is often determined by the
electronic band structure. Charge carriers are mainly electrons and electron-holes.
The conduction mechanism in polymers, which are commonly used as host matrix for field grading
systems, is rather complex and still under investigation. At low electric fields ionic conduction seems to
dominate. Only at high fields close to the breakdown strength electronic conduction takes over and
becomes dominant [Bärsch 2008].
Polymers are usually good insulators due to the absence of free and mobile electrons. Depending on
the purity of the polymer, free ions may, however, be available, increasing the conductivity. Other
reported sources of ions are ionic dissociation by thermal agitation. The number of ions from this source
increases exponentially with temperature as shown in the following Arrhenius equation [Warfield 1961]:
𝑊𝑎
𝑓
(− )
k∙𝑇
𝑁 = C1 𝑒
Equation 3.1

where
𝑁 is the number of ions formed
C1 is a constant
𝑊𝑎𝑓 is the activation energy for the formation of an ion
𝑘 is the Boltzmann constant
𝑇 is the absolute temperature.
Furthermore, some ions are formed by background radiation.
Conduction in polymers depends not only on the number of ions present but also upon the mobility of
the ions. This mobility, which is proportional to the reciprocal of the viscosity of the system, increases
exponentially with temperature following also an Arrhenius equation [Warfield 1961]:
𝑊𝑎𝑚
(− )
𝜇 = C2 𝑒 𝑘∙𝑇

Equation 3.2

where

27
TB 794 - Field grading in electrical insulation systems

𝜇 is the mobility
C2 is a constant
𝑊𝑎𝑚 is the activation energy for the mobility of an ion
𝑘 is the Boltzmann constant
𝑇 is the absolute temperature.
Hence electrical conductivity of a polymer is the product of the number and the mobility of the ions
present [Warfield 1961]:
𝐸1 +𝐸2
𝜎 = C1 C2 𝑒 (− 𝑘∙𝑇
)

Equation 3.3

where
𝜎 is the conductivity at absolute temperature 𝑇.
Conductivity is also influenced by the chain-length of the polymer, the degree of cross-linking and its
crystallinity. All of them reduce the mobility of ions, which leads to lower conductivity. This is the reason
why XLPE cable insulation has a rather low conductivity, and hence the problem of volume charges
trapped inside the dielectric has to be considered for a proper design.
Special polymers with alternating single and double bonds along the molecular chain, so-called
conjugated polymers like e.g. polyacetylene, show a small intrinsic electric conductivity already at room
temperature. It is based on the electrons being able to move along the polymer backbone [Dai 2004].

3.1.2 Permittivity
With respect to permittvity, materials are usually not classified into specific groups like conductive
materials. However, the polarization modes contributing to the overall permittivity lead themselves to a
classification in cases where a specific mode of polarization is dominant.
Polarization describes the ability of a dielectric material to generate and/or modify electric dipoles under
the influence of an electric field. It results from the separation and alignment of electric charges brought
about by that field. The larger the dipole moment arm (the separation of charges in the direction of the
field) and the larger the number of these dipoles, the higher is the material’s permittivity [Ulrich 2000].
The following modes of polarization can be distinguished [Bärsch 2008] [Ulrich 2000]:
- Orientational polarization – existing polar molecules (dipoles) aligned by the electric field
- Displacement or deformation polarization
o Electronic polarization – slight displacement of electrons with respect to the nucleus
o Ionic and atomic polarization – slight displacement or orientation of ions and nuclei in a
molecule or lattice
 e.g. ferroelectric materials – with spontaneous ionic polarization
- Interface and space charge polarization

All of these modes contribute to the material’s permittivity, depending on the mechanisms relevant to
the given dielectric. They do not only determine the value of the relative permittivity, εr, but also how
it varies with frequency, temperature, bias, impurity concentration, and crystal structure [Ulrich 2000].
The four general mechanisms important to candidate materials for field grading are electronic, atomic,
ionic and last but not least interface charge polarization. Since the orientational polarization requires
mobile polar molecules it is less important to solid dielectrics.
Electronic polarization involves charge symmetry distortion of an atom or molecule. Under the influence
of an applied electric field the nucleus and the negative charge center of the electron cloud shift in
opposite directions, creating a small dipole. This induced dipole effect occurs in all materials, but is
usually relatively small compared to other polarization mechanisms since the moment arms of these
dipoles are very short, usually a fraction of the size of an atom [Ulrich 2000]. The described mechanism
is very fast, and, therefore, the polarization is almost independent of the frequency of the applied field.
Atomic polarization occurs in substances made up of more than one type of non-ionic atoms, which is
why different elements will normally not share the electron cloud equally. The negative charge center
will be shifted towards the more electronegative atoms resulting in a permanent dipole. An externally

28
TB 794 - Field grading in electrical insulation systems

applied electric field generates opposite forces on various parts of the molecules causing electronic
polarization but also the nuclei to move relative to each other and to align with the field. It is this relative
movement of the nuclei within the molecule which is attributed with the atomic polarization. The atomic
polarization is, with (1…15) % compared to the electronic polarization, usually not very strong, but can
reach up to 30 % under special circumstances [Ulrich 2000] [Gor 2003].
Ionic polarization is similar to atomic polarization but involves the movement of ionic species under the
influence of the electric field. It can result in very high relative permittivity, up to several thousands,
but is, compared to electronic and atomic polarization, much slower.
Orientational polarization requires permanent dipoles to be present, which align themselves with the
orientation of the applied electric field. This rotation occurs on a timescale that depends on the torque
and surrounding local viscosity of the molecules. This type of polarization is, therefore, mostly effective
with liquids and gases. The most prominent example is the water molecule H2O, which has a high dipole
moment. This is used e.g. in microwave ovens for heating due to the dielectric losses.
Each mode of polarization has got a typical dispersion area limited by the relaxation frequency, up to
which it is effective (see Figure 3.1). At frequencies above the dispersion area the corresponding mode
of polarization is unable to contribute any more to the overall permittivity of the material. Reason is that
the electric field is not applied for a time duration long enough to allow an earlier polarization of opposite
polarity to relax. As a consequence, relative permittivity 𝜀𝑟 as well as the dissipation factor tan 𝛿 are
frequency dependent [Bärsch 2008].

Figure 3.1: A dielectric permittivity spectrum for unitary materials over a wide range of frequencies. Real and imaginary
parts of permittivity  are shown, and various processes are depicted; acc. to [HP 1992]

Frequency dependance of permittivity may impact the efficiency of a field grading material in the time
domain significantly, depending on the intended application. In case the field grading shall be effective
at transient stresses (i.e. lightning impulse voltage, commutation impulses, etc.) as well as at power
frequency (50/60 Hz), a material of lower but balanced overall permittivity may be better suited than
one of very high permittivity at low frequencies only.
The relaxation frequencies for electronic as well as atomic polarization are typically above the frequency
spectrum relevant to HV apparatus. Often they only matter to optical physics applications. For materials
with significant ionic and orientational polarization mechanisms, however, relaxation frequencies may
have to be considered.
Within the group of dielectrics with dominating ionic polarization two different kind of behaviours can
be distinguished. So-called ferroelectric materials will not lose their polarization even when the electric
field is removed (often described as spontaneous electric polarization), whereas paraelectric materials
will. In the lattice of a ferroelectric material ions may take several different states. An electric field can
pull some ions into configurations that do not relax back to the previous state once the field is removed.
As a result, ferroelectric, analogous to ferromagnetic materials, can have a residual polarization after
the field has been removed. These materials usually provide the highest permittivities and are,
therefore, in the focus of scientific research in the field of power energy storage devices.

29
TB 794 - Field grading in electrical insulation systems

Barium titanate is the most popular member of ferroelectric ceramics, although there are others
available, which exceed barium titanate’s permittivity by a factor of two (e.g. lead magnesium niobium-
lead titanate: PMN-PT; however, lead-free materials need to be developed).
Ferrorelectric behaviour can also be observed with polymers where atomic polarization is dominating
and where the dipoles can be aligned by the applied electric field. Polyvinylidene fluoride and
polytrifluoroethylene are two representatives of this subgroup. Since atomic polarization is weaker
compared to the stronger ionic one permittivities of these polymers are somewhat lower.
A commonly named representative for orientational polarization is water. Due to the strong permanent
dipole of its molecule the relative permittivity can reach values of more than 80.
Unitary high permittivity materials and materials with nonlinear electric properties are typically of
ceramic type and, therefore, difficult to apply directly in field grading systems with respect to
processability, durability, flexibility and stability. Conducting and semiconducting materials are likewise
not directly applicable due to either their excessive conductivity or their brittle structure. Only after
being dispersed in a suitable host matrix, which is selected mainly based on mechanical requirements,
an easy-to-apply composite is formed. Typical host matrices are varnishes (resins), EPDM and VMQ
(elastomers) as well as thermoplastics and thermosets.
For these composites interface charge polarization is typically the most dominant polarization
mechanism. It is present when charge carriers can migrate an appreciable distance through the
composite but then get trapped e.g. at interface boundaries before they can reach an electrode. This is
the case when materials with different dissipation factors tan 𝛿 are combined or micro- and macroscopic
interfaces like cellulose fibres do exist. Therefore, this mode of polarization applies to almost every
material mix like suspensions or colloids, biological materials, phase separated polymers, blends and
crystalline or liquid crystalline polymers and the commonly used oil-paper insulation system.
The relaxation frequency for interface charge polarization is quite low and is determined basically by
the conductivity and permittivity of the involved dielectrics (see Figure 3.2). In simplified terms the
relevant process can be compared to a low-pass RC filter where the conductivity of the material with
the higher dissipation factor and the permittivity of the other material determine the limiting frequency.

Figure 3.2: A dielectric permittivity spectrum for composite materials over a wide range of frequencies. Relative
permittivity εr and dissipation factor tan δ are shown. Compared to unitary materials interface charge polarization
contributes at low frequencies; acc. to [Beyer 1986]

The electrical properties of a composite are in general difficult to predict due to interaction effects
between the filler particles’ surface and the host matrix. Also statistical effects have a significant impact.
A number of “effective medium approximation” models have been published, from which only the
Maxwell-Garnett approximation, the Bruggeman’s model and the Maxwell-Wagner-Sillars model shall be
mentioned in this context.
Effective medium approximations are based on the assumption that each filler particle is, in average,
surrounded by the host matrix that has the assumed homogeneous medium property. With increasing
filler level, however, percolation of the filler particles has to be taken into consideration.

30
TB 794 - Field grading in electrical insulation systems

The Maxwell-Garnett approximation usually provides valid results at low volume fractions of fillers with
dilute conductivity (below the percolation threshold) of up to 10 % or 20 %, respectively [Steeman 1992]
[Beek 1967]. It is based on a mean field approximation covering spherical inclusions according to the
Maxwell’s relation embedded in a continuous matrix of polymer [Barber 2009]. The formula can be
extended to complex permittivities and then also covers dielectric losses of the composite.
Bruggeman’s model assumes a binary mixture, composed of repeated units, which in turn are composed
of the host matrix and a spherical or ellipsiod inclusion in the center [Barber 2009]. This mixing model
yields better results even though the dispersion is not strongly diluted.
The Maxwell-Wagner-Sillars model is finally able to also take into account the effects of interface charge
polarization, which may occur at inner dielectric boundary layers between filler particles and the matrix
[Zulkifli 2012] having different permittivities and/or conductivites [Romasanta 2011].
It was found that with a larger surface area at the interface between filler particles and host matrix the
interfacial polarization increases [Salaeh 2012]. Investigations of functionalized graphene sheets with
high aspect ratios [Romasanta 2011] and high aspect ratio fillers like barium titanate fibres and
graphene platelets [Wang 2012] confirm that a large interface surface area is able to provide higher
permittivity levels. It is also reported that by chemically modifying the filler’s surface the typical high
power loss related with the Maxwell-Wagner-Sillars process can be significantly reduced [Romasanta
2011]. Such an additional process step can, however, significantly raise the production costs, and
thermally expanded graphene sheets are presented as an alternative. The thermal reduction of graphite
oxide is reported to provide a chemically modified graphene sheet or so-called functionalized graphene
sheet without the need for further modification steps [Romasanta 2011].
Carbon black is the most commonly used filler to increase conductivity as well as permittivity of
polymers. At low filling levels and well dispersed particles conductivity usually does not change much
since the average distance between the conductive particles is too large to form a continuous conductive
path. Maxwell-Garnet approximation may be applied to calculate permittivity and conductivity of such a
composite. At increasing filling levels, however, conductivity starts to increase rapidly as the distance
between the conductive particels becomes less than 10 nm [Bärsch 2010]. Further details about carbon
black, its application as filler and percolation theory are given in 3.3.2.
The effective medium approximations and the percolation theory can be combined into a “general
effective media (GEM) equation”, which covers most aspects of both theories. Adapated equations have
been proven to provide good results with respect to AC and DC conductivity as well as complex
permittivity of composites [Wu 2003].
So far it has been assumed that the particulate fillers are homogeneously dispersed in the composite.
Recent development investigates possibilities to disperse fillers inhomogeneously but in a controlled
way within an insulation or field grading system. This is explained in greater detail in 2.2.4.2.

3.2 Matrix materials


3.2.1 Glass (resistive glazing)
Theoretically, the overall axial potential distribution of an insulator can be improved by resistive grading
in form of a conductive surface layer. Since the 1970s ceramic insulators with weakly conductive glazing
have been in use. The first patent was granted in 1940 [Forrest 1940] and the application published
shortly afterwards [Forrest 1942]. Semi-conducting glaze properties are obtained by a conducting phase
in the glass matrix, usually metal oxide crystallites. In the beginning metal oxides, especially iron oxide,
were applied. Partially reduced titanium oxide improved thermal properties but turned out to be instable
in the long term [Gubanski 2005]. Finally, an insulator manufacturer developed a glazing system that
contains antimony-doped tin oxide, resulting in a constant leakage current flow of (0.5…1) mA across
the insulator surface [Fukui 1978] [Naito 1997] [Mizuno 1999]. Further information is given in 5.6.1

31
TB 794 - Field grading in electrical insulation systems

3.2.2 Thermoplastic polymers


High molecular polymers consist of one or several different monomers bonded together by
polymerization, polyaddition or polycondensation. For thermoplastic polymers the resulting long chain
molecules mostly have a linear structure (examples in Figure 3.3), interconnected by van der Waals
bonds. When heating the material the van der Waals bonds split up, and the polymer melts. When
cooling the substance sets again. This reversible behavior allows shaping by thermal treatment, e.g.
extrusion processes. Among others, polyethylene (PE), polyvinyl chloride (PVC) and polypropylene (PP)
are widely used thermoplastics in high voltage engineering, e.g. as insulators or spacers.

Figure 3.3: Molecular structure of polyethylene, polypropylene and polyvinyl chloride; acc. to [Küchler 2017]

The copolymer of ethylene and vinyl acetate is called ethylene-vinyl acetate (EVA), or more precisely
poly(ethylene-vinyl acetate) (PEVA). Figure 3.4 shows the molecule structure. The content of vinyl
acetate determines the chemical and physical properties and thus, the application of EVA. The technical
properties of the polymer with about 20 wt% 2 vinyl acetate are very similar to low density PE.
Increasing the fraction to (40…50) wt% of vinyl acetate leads to rubber-like materials. EVA is used for
cable sheathing and for sheds of various types of insulators as well as for heat shrinking cable
accessories [Brinkmann 1975].

Figure 3.4: Molecule structure of EVA consisting of ethylene (right) and vinyl acetate (left) monomers; acc. to
[Brinkmann 1975]

3.2.3 Thermoset polymers


As for the thermoplastics the thermosets originate from polymerization, polyaddition or
polycondensation reactions. But the resulting long chain molecules are chemically cross-linked to each
other, forming a network structure. Thus, those materials do not melt with increasing temperature but
soften and show an elastomeric behavior when reaching the glass transition temperature. After curing
or crosslinking, respectively, thermosets can only be shaped by mechanical processes. Thermosets,
typically used as matrix materials, are epoxy resin, alkyd resin and polyurethane.
Epoxy resins (EP) are widely used for electrical engineering purposes mostly as cast resins or as
adhesives. They can also be used as basis for varnishes, coatings and tapes. The resin contains more
than one epoxide group per molecule. The epoxide group consists of an oxygen atom bonded to two
carbon atoms, which are further connected, resulting in a reactive ring structure (Figure 3.5 left). For
applications the epoxy resin is mixed with a hardener.

2
wt% = percentage by weight

32
TB 794 - Field grading in electrical insulation systems

Figure 3.5: Structure of epoxy group and urethane group; acc. to [Küchler 2017]

Alkyd resins form the basis of varnishes, coatings and tapes. The term “alkyd” reflects the combination
of “alcohol” and organic “acid”, the main components of the material. Mostly, fatty acids are used. The
number of double bonds of the unsaturated fatty acids determines if the resin is air drying or needs
baking.
Polyurethane (PUR) is synthesized by polyaddition from polyisocyanate with polyhydric alcohols.
Polyurethanes are characterized by the urethane group (Figure 3.5 right). If the linear molecules of
thermoplastic polyurethanes are cross-linked, this results in a thermoset polymer. Polyurethane is used
as basis for varnishes and tapes.
As described above, polyethylene (PE) is a thermoplastic polymer, thus showing low long term
temperature stability and creeping of the material. Crosslinking the extruded polyethylene in a
subsequent process step by e.g. peroxide crosslinking or electron beam forms the thermoset XLPE.
XLPE is commonly used for cable insulations and can be used for warm shrink applications.

3.2.4 Elastomers
Elastomers are wide-meshed thermoset polymers, thus being elastic over a large temperature range.
The shaping process is identical to thermosets. Due to their shape memory elastomers can be
permanently expanded or compressed without losing their resilience. This behavior is especially useful
for cable accessories where a permanent contact pressure between the body of the accessory and the
cable insulation is necessary to maintain the dielectric strength.
Silicone rubbers (VMQ) are elastomers widely used in high voltage applications. The silicone
macromolecules consist of an inorganic backbone of silicon and oxygen atoms surrounded by organic
side groups (Figure 3.6). Depending on the vulcanization process the relevant silicone rubbers are
divided into three groups: high temperature vulcanizing silicone rubbers (HTV silicone), room
temperature vulcanizing silicone rubbers (RTV silicone), and liquid silicone rubbers (LSR). The wide-
meshed structure of silicone rubbers allows a comparatively high diffusion of gases, water vapor, and
oil molecules. This is an advantage for the use of VMQs in cable accessories, as voids in the interface
of cable insulation and insulation body disappear rapidly after installation. Suspension insulators, post
insulators and hollow core insulators benefit from the tracking resistance and the excellent hydrophobic
surface properties of silicone rubbers. Additionally, silicone rubbers can be used for tapes, coatings and
varnishes.

Figure 3.6: Chemical structure of silicone with organic side groups

Beside XLPE and silicone rubber, ethylene propylene rubber (EPM, often incorrectly denoted as EPR) is
a standard elastomer used for extruded cable insulation and cable accessories. Monomers of ethylene
and propylene are randomly combined to synthesize EPM, resulting in a rubber resistant to heat,
oxidation, ozone, and weather impact. Due to the comparatively high dielectric losses the use of EPM
in AC systems is typically limited to the distribution and sub-transmission voltage levels.
Closely related to EPM, the ethylene propylene diene rubber (EPDM rubber) takes diene monomer as
third partner in the polymerization process. The resulting elastomer has similar properties as EPM.
Widely used as sealing material (e.g. O-rings), it is also taken as insulating material. Further, beside
silicone rubber, EPDM is commonly used as cold shrink material.

33
TB 794 - Field grading in electrical insulation systems

3.3 Fillers
3.3.1 Goal
Fillers are particles added to a matrix to modify its physical properties (color, strength, permittivity,
resistivity, etc.). In the context of field control, fillers are used to tune either the permittivity or the
conductivity of a material. Fillers can be added homogeneously throughout the material or with a spatial
gradient of the concentration (Figure 3.7). In some cases multiple fillers are used; either to obtain
specific properties, for example gradient materials (see 2.2.4.2) or to improve the manufacturability,
reduce the cost, etc.

Figure 3.7: Schematic representation of different filler types: a) single filler homogeneously dispersed, b) single filler
homogeneously dispersed above percolation limit (the arrow illustrates one possible continuous path from one
particle to the next), c) single filler with spatial gradient, d) system of two fillers with spatial gradients

To adjust the electrical conductivity, it is necessary that the filler concentration is above the percolation
limit. If the filler is conducting (for example carbon black) the level of conductivity is controlled by the
concentration (see 3.3.2). For semiconducting fillers such as SiC, ZnO microvaristors, or antimony-tin
oxide (ATO – SnO2/Sb2O3) the conductivity level is determined by the properties of the filler (see 3.3.3).
With their help it is possible to provide a kind of self-adaptive field grading systems [Brandes 2008]
[Donzel 2011] [Seen 2012] [Rüger 2012] [Greb 2015] [Tenzer 2011] [Tenzer 2013c] [Tenzer 2015]
[Winter 2011] [Winter 2014] [Winter 2015] [Secklehner 2013] [Secklehner 2015] [Secklehner 2017a]
[Secklehner 2017b]. The commonly applied concept is that the filler particles become more conductive
after a certain threshold level of the electric field is exceeded. Further new approaches are addressed
in 3.3.4.
The permittivity of a composite is tuned by modifying the concentration of high permittivity particles
(for example SrTiO3,TiO2, etc.). It is not necessary to be above the percolation limit as the effect relies
on polarization.

3.3.2 Carbon black and carbon nanotubes


For many field grading applications materials with a comparatively low specific volume resistivity of less
than 100 cm is required. These materials are usually compounds of a matrix polymer and an
electrically conductive filler. The most widely used fillers are carbon blacks, which, with a carbon content
of more than 96 %, in principle demonstrate a variety of carbon (C). Carbon blacks with standard
conductivity are produced in the so-called acetylene-black-process, a controlled incomplete combustion
of acetylene in a special reactor. Other gases and specific technologies are used to produce carbon
blacks with a very high electrical conductivity.
The size of primary particles of conductive carbon blacks is typically in the range of (1…100) nm.
Primary particles in the carbon black do not exist in an isolated form, but rather aggregate in a beads-
like manner into aggregates of different sizes. Smaller particles have larger contact surfaces, which lead
to a self-aggregation behavior. The individual primary particles are coupled by van der Waals forces and
can only be separated by applying very strong mechanical forces. Thus, an aggregate can be defined
as a separate, solid, colloidal particle, which is the smallest dispersible unit in the carbon black and,
therefore, must be considered as the actual primary particle in a carbon black as it is delivered.

34
TB 794 - Field grading in electrical insulation systems

Aggregates tend to form agglomerates. The agglomerates of conductive carbon blacks typically show a
less spherical appearance then blacks used for pigments. They are rather branched agglomerates (Table
3.1) with a high specific surface of (80…1200) m2g-1. Carbon blacks consisting of branched aggregates
are easier to incorporate into the bulk material and show a better electrical conductivity in comparison
to other aggregate types.
Table 3.1: Types of carbon aggregates

Spherical aggregate Elliptical aggregate

Linear aggregate Branched aggregate

Conductivity of carbon black filled polymer composites changes rapidly when the concentration of the
homogeneously dispersed carbon black reaches the so-called percolation threshold. Well below and well
above the percolation threshold, change of conductivity is negligible. The percolation threshold level in
turn is depending on the conductivity of the carbon black, the matrix polymer and the mixing process
parameters, but more importantly on the carbon black’s micro structure. Therefore, it is typically
required to evaluate the parameters experimentally and then adapt a suitable mixing theory for
predicting the electric properties.
The objective of percolation theory is to characterize the connectivity properties in complex and
randomly mostly disordered structures that are composed of simple components. It provides a natural
frame for the theoretical description of random composites. The percolation theory gives a
phenomenological power dependance of the effective property of a diphase composite in the volume
fraction range, where one phase has just formed or is about to form a continuous percolation network
or infinite cluster. However, the main problem with using the percolation theory in practical cases is
that it is only strictly valid when the ratio of the properties of the two phases is infinite or zero [Wu
2003].
The resulting typical behavior of the specific resistivity of a carbon black filled compound is shown in
Figure 3.8. The figure shows the resistivity of three polypropylene-carbon black-compounds with
different specific surfaces in dependance on the percentage of the added carbon black.

35
TB 794 - Field grading in electrical insulation systems

Figure 3.8: Influence of the specific surface of carbon blacks on the electrical resistivity of a carbon black-PP-
compound (sketch after R. Gilg, DEGUSSA)

Taking into consideration that the electrical conductivity of these compounds is provided by small
contact areas between the aggregates, it is easy to understand that the ability of carbon-black-filled
materials to carry an electrical current is very limited.
Even if the earlier mentioned preconditions for the application of the percolation theory are not perfectly
met for carbon black and the typical polymers, it still can provide some useful insight into other effects
relevant to e.g. heat- and cold-shrink field grading tubes filled with carbon black particles. When these
are mechanically expanded the probability to form a continuous conduction path increases, since
expansion in one direction is accompanied with reduction of thickness in other directions. Existing but
separated small clusters of carbon black may get connected to each other forming larger conductive
clusters. Both, the probability to form a continuous conduction path and conductivity, will increase. This
effect is reversible, which means conductivity decreases again when the material relaxes. The effect
may also cause anisotropic changes of the electric parameters. It should also be taken into consideration
that the compounding of carbon black and polymers can lead to a strong increase in the viscosity of the
mixture, which may limit the usability.
A promising alternative for future developments with good electrical conductivity are carbon nanotubes
(CNT). CNT are microscopic tubular structures (molecular nanotubes) consisting of carbon atoms that
are commercially available but have not reached a broad range of application yet. Their walls consist of
carbon only, where the carbon atoms form a honeycomb-like structure with hexagons and three binding
partners each. The diameter of the tubes is usually in the range of (1….50) nm. In comparison, the
longitudinal dimension of the tubes is much higher. Thus, they are more a fiber-like material in
comparison to carbon black, but in a nano-scale. A distinction is made between single and multi-walled
CNT. Because of the limited data available, the following paragraph should be understood as guiding
information.
The high electrical conductivity of CNT allows to use rather low amounts of less than 5 % of CNT to
achieve about the same electrical resistivity as by conventional filling of a compound. As known from
carbon blacks, there is a strong dependance of the resistivity on the filling degree (Figure 3.9). Filling a
low viscosity bulk material, e.g. silicone polymer, with CNT leads to a very strong increase of the viscosity
of the mixture. This may become a limiting factor for the application, so that the very low resistivity
values of compounds reported in the literature are not possible to reach with standard mixing
technologies, such as stirring or role-milling. Another measurable effect is the increase of resistivity of
a compound with increasing number of mixing cycles (Figure 3.10). It can be assumed that the mixing
process leads to changes of the orientation of the CNT and the number of contact points to each other,
so that the resulting specific volume resistivity increases.

36
TB 794 - Field grading in electrical insulation systems

Figure 3.9: Specific volume resistivity of a VMQ-CNT-compound in dependance on the amount of CNT [Wacker Chemie
AG, Munich]

Figure 3.10: Specific volume resistivity of a VMQ-CNT-compound in dependance of the number of mixing cycles of the
compound on a triple-roll-mill [Wacker Chemie AG, Munich]

3.3.3 Nonlinear materials


As fillers for nonlinear field grading materials either silicon carbide particles or zinc oxide microvaristors
are used in commercially available products. In the following both systems are described. The use of
ATO is still under investigation.
3.3.3.1 Silicon carbide (SiC)
SiC is a semiconductor that can take many different crystallographic structures. Depending on the
doping (on purpose or due to impurities) SiC is either an n- or a p-semiconductor. Green SiC (doped
with nitrogen or phosphor) is an n-semiconductor, while black SiC (doped with Al, B, Ga) is a p-
semiconductor [Senn 2012]. The SiC powder used for field control is of the same quality that is produced
for grinding application. The SiC grains are typically sharp edged bulky particles and are classified
according to their particles size, see Figure 3.11 [Donzel 2011].

37
TB 794 - Field grading in electrical insulation systems

Figure 3.11: Scanning electron microscopy picture of SiC particles [Donzel 2011]

The macroscopic electrical properties of the SiC based compound depend on the microscopic properties
of the contact zones between SiC grains, see Figure 3.12 [Donzel 2011]. It was shown that these
contacts can be modeled by Schottky-like barriers. Tunneling by field emission is the dominant
conduction mechanism [Mårtensson 2003]. The size of the SiC particles has direct influence on the
electrical properties through the number of contacts of the current path [Onneby 2001], see Figure
3.13. Usually SiC powder with a mean particle size of a few micrometers to few tens of micrometers are
used. The percolation limit for SiC based composite lies at about 32 vol% 3 [Seen 2012].

Figure 3.12: Schematic illustration of an SiC based composite; interparticle contacts (responsible for the macroscopic
electrical properties of the composite) highlighted in red [Donzel 2011]

Figure 3.13: Dependance of the electrical properties (here: resistivity) of SiC based composites on the particle size
[Seen 2012]

The doping of the SiC particles changes their intrinsic conductivity and, therefore, also influences the
electrical properties of powder bed4 or composites [Seen 2012] [Mårtensson 2003]. The large range of
U-I-characteristics measured on SiC of different doping and grain size is illustrated in [Brandes 2008],

3
vol% = percent by volume
4
powder bed: a layer that embeds any solid substance reduced to a state of fine, loose particles by crushing, grinding,
disintegration, etc.

38
TB 794 - Field grading in electrical insulation systems

see Figure 3.14. The U-I-characteristic of SiC powder is slightly nonlinear with exponents of nonlinearity,
α, between 2 and 4 (I = kU α).

Figure 3.14: Current vs. voltage characteristics of six different grades of SiC particles, illustrating the large spread of
electrical properties; measurements made on powder beds [Brandes 2008]

The strong effect of impurities on the electrical properties of SiC is an issue for engineering application:
“nominally identical” SiC can have very different properties, see Figure 3.15 (acc. to [Seen 2012]). This
is one reason for the large tolerance in the specification of field control elements based on SiC, for
example end corona protection tapes, see Figure 3.16 [Brandes 2008].

Figure 3.15: Resistivity vs. electric field measured on two black-SiC powders of identical grain size but delivered by
two different suppliers; acc. to [Seen 2012]

39
TB 794 - Field grading in electrical insulation systems

Figure 3.16: Boundaries for the current-voltage characteristics of two commercial SiC tapes, named Type “A” and Type
“B” here [Brandes 2008]

SiC is a hard material (9 on the Mohs scale; diamond hardness is 10). This has a practical impact on
the production of field control compounds and tapes based on SiC: the equipment is subject to severe
abrasion.

3.3.3.2 Metal oxides (MO)


Diverse nonlinear ceramics such as for example ZnO, SnO 2, or SrTiO3 are known, however doped ZnO
is the material of choice for MO surge arresters and state-of-the-art field control industrial applications
due to its unmatched properties. The nonlinearity of ZnO originates in the microstructure of the ceramic
[Greuter 1995] [Donzel 2011], see Figure 3.17: conductive grains (<1 Ωcm) separated by electroactive
grain boundaries. At the grain boundaries electrostatic potential barriers (double Schottky barriers) are
built up, which control the current flow through the material. The switching voltage of one boundary is
about (3.2…3.4) V. As long as the potential difference across the boundary is below this value the
boundary is insulating, whereas above this value the boundary “breaks down”, and the current flow
increases exponentially up to a value determined by the conductivity of the ZnO grains.

Figure 3.17: Schematic illustration of a ZnO microvaristor based composite; grain boundaries (responsible for the
macroscopic electrical properties of the composite) highlighted in red [Donzel 2011]

The electrical properties of ZnO thus mostly depend on the number of boundaries, i.e. the size of the
grains. ZnO ceramics are manufactured by sintering, and the grain size can be tuned by varying
processing parameters such as doping elements, sintering temperature and time (Figure 3.18). ZnO can
be manufactured as varistor blocks (for MO surge arresters) or as microvaristor powder (as filler for
field control composite, see Figure 3.19).

40
TB 794 - Field grading in electrical insulation systems

Figure 3.18: Effect of sintering temperature (top) and time (bottom) on the electrical properties of ZnO varistors; acc. to
[Gupta 1990]

Figure 3.19: Scanning electron microscopy picture of a ZnO microvaristor [Matsusaki 2012]

The large range of electrical properties that can be achieved with ZnO microvaristor based field control
materials is illustrated in Figure 3.20 [Donzel 2011]. ZnO microvaristors feature very high exponents of
nonlinearity (α-values between 20 and 50).

Figure 3.20: Resistivity vs. electric field for various ZnO-microvaristor based composites, illustrating the wide range of
electrical properties achievable with ZnO microvaristors [Donzel 2011])

ZnO microvaristors are manufactured specifically for electrical applications, therefore the electrical
properties of ZnO particles are better controlled than those of SiC, which is primarily produced for
grinding applications. The hardness of ZnO is 5 on the Mohs scale.

41
TB 794 - Field grading in electrical insulation systems

3.3.4 Further approaches under development


3.3.4.1 Intrinsic conductive polymers: Polyaniline
Polyaniline is an intrinsic conducting polymer (ICP) that was first made in 1834 [Feast 1996] but has
been studied extensively over the past fifty years. It is polymerized from the aniline monomer, shown
in Figure 3.21 (a), and can be found in three different oxidation states. Leucoemeraldine which give rise
to a white/clear compound, emeraldine which will be green for the salt and blue for the base and last
(per)nigraniline which will have a blue/violet color [Feast 1996]. Polyaniline has a special structure
containing an alternating arrangement of benzene rings and nitrogen atoms. The nitrogen atoms can
exist either as an imine (in an sp 2 hybridized state) or an amine (sp 3 hybridized). Depending on the
relative composition of these two states of nitrogen, and also depending on if they are in a quaternized
state or not, different forms of polyaniline can result as shown in Figure 3.21 (b). The only form of the
four polyaniline compositions in Figure 3.21 (b) that is conducting is the green protonated emeraldine
form [Ramakrishnan 1997], and the conductivity can vary from 10-10 to 102 Scm-1 [Libert 1997].

Figure 3.21: (a) Aniline monomer and (b) the three forms of polyaniline and doped PANI-EB; acc. to [Ramakrishnan
1997]

There are several kinds of ICP, i.e. polypyrolle, polythiophene and polyaniline. Among the ICPs,
polyaniline is considered to be one of the most environmentally stable conductive polymers, which
makes it a promising candidate for the demands of high voltage insulation systems. Applications, in
which polyaniline have been mentioned, are semiconducting layers in cables as well as fillers in
polypropylene films used in power capacitors [Cottevielle 1999a] [Cottevielle1999b] [Kieffel 2000].
The emeraldine base form (PANI-EB) is an intrinsic semiconductor with a low field DC conductivity of
(10-9….10-8) S/m (Figure 3.22) and an optical bandgap energy of Eg = 3.8 eV [Aladenize 1995]
[Pelto 2004].

42
TB 794 - Field grading in electrical insulation systems

Figure 3.22: Conductivity for oxidation ratio vs. polyaniline [Aladenize 1995]

The nonlinear behavior was first reported in [Aladenize 1995] with a power law exponent of 2.8.
Nonlinear behavior mainly has its use in field grading applications, something which is used in, for
instance, cable applications and electrical machines. Other uses could also be foreseen, but these are
the major areas of application at the moment [Sonerud 2013].
In thin films of PANI-EB a set of conduction mechanisms are reported to be involved, which are both
electric field and temperature dependent. Both the Arrhenius law and Mott’s variable range hopping
model has been used to describe the temperature dependent conductivity of polyaniline [Mzenda 2002].
It has been reported that Mott’s 1D-variable range hopping (1D-VRH) is the dominant conduction
principle at fields < 100 V/mm and dominated by space charge limited current (SCLC) behavior at fields
higher than 1 kV/mm [Nazeer 2004]. It has also been reported that Schottky type conduction (metal-
insulation-metal) is dominant [Mathai 2003]. Even though many conduction mechanisms are shown for
films in different regions, the conductivity in bulk PANI-EB cannot be fully explained by the models for
thin films [Pelto 2004].
A PANI-EB based compound with nonlinear J-E-characteristics and high processability was developed
by Pelto et al. [Pelto 2004] for field grading purposes. The onset of the strong nonlinearity appears in
the range of 10 kV/mm for pure PANI-EB as well as for the silicone rubber composite also including
graphite. Figure 3.23 (b) shows the J-E-characteristics obtained for PANI-EB powder, which shows an
exponent of nonlinearity, , of 10…25. Figure 3.23 (a) shows the resulting J-E-characteristics for PANI-
EB + graphite composite in a silicone rubber matrix with an  value of approximately 10. Comparing
only the nonlinearity factor of PANI-EB with published data on conventional fillers, summarized in Table
3.2, it appears as PANI-EB has the potential to obtain a stronger nonlinearity than SiC, but weaker
compared to ZnO.

Figure 3.23: (a) J-E-characteristics of silicone rubber + PANI-EB + graphite composite, and (b) PANI-EB powder
[Pelto 2004]

43
TB 794 - Field grading in electrical insulation systems

Table 3.2: Summation of nonlinear conductive properties found in publications

σ0
Filler Matrix % vol α Eb kVmm-1 Ref.
S/m
Carbon black [CB] polyolefins / EPDM 0.2...2.5 ≈ 3×10-12 [Egiziano 1995]
[Mårtensson
SiC/CB EPDM 0.117 ≈4 ≈ 0.22 ≈ 5×10-16
1998]1
0.432/0.08 ≈ 3.1 ≈ 0.027 ≈ 2.7×10-10
SiC/Fe3O4 polybutadiene [Okamoto 2001]1
0.432/0.08 ≈ 1.7 ≈ 0.071 ≈ 2.5×10-6
sealing mastic 6.4 0.18 ≈ 9.2×10-14 [Sonerud 2012]
SiC - 0.7 µm ≈4 ≈ 1.5 ≈ 5.4×10-18
SiC - 3.0 µm ≈ 5.8 0.6 ≈ 3.3×10-17
EPDM 0.4 [Önneby 2001]1
SiC - 9.3 µm ≈ 3.1 < 0.1
SiC - 23 µm ≈ 3.4 < 0.1
0.03 15 3.3 ≈ 3.8×10-16 [Wang 2012a]1
GO VMQ
0.05 17 1.8 ≈ 1.7×10-16 [Wang 2012b]1
PANI-EB+graphite silicone rubber 0.25 ≈ 10 ≈ 10 [Pelto 2004]
ZnO silicone rubber 0.35 22 1.3...1.7 10-12 [Donzel 2011]
0.25 32 0.92
ZnO + boundary formers
[Bi2O2/Sb2O3] + 0.33 31 0.59
LLDPE [Lin 2008]
conductivity enhancers 0.4 22 0.43
[CoO/MnO]
0.45 18 0.26
ZnO epoxy 0.48 ≈ 14 1.23 ≈ 7×10-11 [Secklehner 2017b]
ZnO silicone rubber 0.39 10...11 0.50...0.82 [Yang 2015]
ZnO epoxy 0.42 ≈ 23 0.3 2x10-10 [Matsuoka 2015]
108…1012 
Iriotec 7000®-Series Epoxy, blanket
2…5 surface [Greb 2015]
(ATO) coating
resistance
6.5×10-14
MFF1 (ATO) epoxy, 38 wt%  3.6 [Secklehner 2013]
@30 °C

3×10-14
MFF2 (ATO) epoxy, 38 wt% 3 [Secklehner 2015]
@30 °C

1 Results from fitting plot in publication

3.3.4.2 Graphene oxide


Figure 3.24 highlights the conductive properties in composites filled with graphene oxide (GO). The
interesting percolation behavior of GO in VMQ, in which a low loading (1…2 phr 5) yields materials with
predominantly dielectric behavior, but at higher loading (3…5 phr) a nonlinear field grading-type
material with high nonlinearity exponent (α ≈ 16) arises [Wang 2012a]. The percolation limit for GO in
VMQ was between 2 and 3 phr. Additionally, the relative permittivity was increased from 3 (VMQ) to
7.8 (VMQ + 5 phr GO). These composites thus provide both resistive and capacitive field grading effects.
Additionally, the degree of reduction of GO had also marked effects on the σ-E-characteristic. At higher
filling ratios there is a clear saturation current at high field strengths, which is not properly described
by the simple σ0(1+[E/Eb]α-1) dependance, and fitting results were obtained by taking only the part of
the curve below the saturation current [Wang 2012a] [Wang 2012b].

5 phr = parts per hundred rubber

44
TB 794 - Field grading in electrical insulation systems

Figure 3.24: σ-E-characteristics for graphene oxide (GO) at different loadings in a VMQ matrix [Wang 2012a]

However, the dielectric properties of graphene nanocomposites in combination with other inorganic
particles is intensively studied presently, since they offer possibilities to attain very high permittivity
materials and are prospective FGMs. For example, Wang et al. have studied the synergistic effect of
barium titanate fibers and graphene platelets (GPLs) in silicone rubber composites to obtain very high
dielectric constant materials [Wang 2012c]. The GPLs alone can increase relative permittivity at a very
low filler volume fraction due to their very high polarizability (Figure 3.25 (a), blue curve). However, the
loss increased dramatically when the GPL loading reached the percolation threshold for electrical
conductivity (Figure 3.25 (a), red curve), which can lie at a very low concentration depending on the
extent of exfoliation of the graphite into graphene [Pang 2010]. In another study of the system
PVDF/GO, the relative permittivity has been observed to peek at values of 2000 (Figure 3.25 (b), dotted
curve) [Cui 2011]. Dissipation/conductivity increased at the same time (Figure 3.25 (b), solid curve).

Figure 3.25: (a) Dielectric constant as a function of vol% of GO [Pang 2010]; (b) Dielectric permittivity as a function of
vol% GO [Cui 2011]

3.3.4.3 Tin dioxide


Tin dioxide (SnO2) based ceramics is a metal oxide, which exhibits intrinsic non-ohmic conductivity
similar to zinc oxide where the non-ohmic conduction is a grain-boundary phenomenon [Glot 2011].
The oxide has lately been intensively studied as a potential replacement of ZnO in varistors and surge
arresters. Several advantages such as higher thermal conductivity and activation field as well as lower
leakage current have been listed compared to ZnO and observed in Figure 3.26 [Metz 2007]. The
conductive properties are, however, strongly dependent on the sintering process, and many variations
of both ZnO and SnO2 can be made. An issue seen in SnO2 based varistors is the joule heating due to
a rather high grain resistivity, (1.2…40) Ohmcm, which is likely to come from the porosity of SnO2,
resulting in poor particle contact [Glot 2011, Glot 2010].

45
TB 794 - Field grading in electrical insulation systems

As for ZnO, the conductive properties of SnO 2 are highly dependent on the type and amount of dopants
used. SnO2 on its own does not sinter into a ceramic, wherefore sintering aids must be added [Moulson
2003]. The addition of CuO, a sintering aid, to SnO+Co3O4-Nb2O5-Cr2O3 has shown to affect the electrical
properties in a beneficial way until the CuO on itself forms conductive percolation paths (between
(0.5…2) mol% CuO), which can be seen from the J-E-chart in Figure 3.27, where the content of CuO is
increased from 0 to 8 mol%. Nonlinearity exponents  up to 75 are reported, and the other
compositions’ nonlinear parameters are listed in Table 3.2 [Gaponov 2011].

Figure 3.26: E-J-characteristics for a ZnO and SnO2 powder [Metz 2007]

Figure 3.27: J-E-characteristics for SnO+Co3O4-Nb2O5-Cr2O3 powder with various amounts of CuO, in
mol.%: 0 (1), 0.1 (2), 0.5 (3), 2 (4), 8 (5) [Gaponov 2011]

3.3.4.4 Bridging particles: Antimony doped tin oxide


For a microvaristor based field grading material to function properly, one requirement is to obtain
percolation. Due to the high filling ratio and large spherical geometry of the varistor particles, the
composites can suffer in mechanical strength. The mechanical strength can be increased by replacing
an amount of varistor particles by smaller conducting particles that will bridge the nonlinear properties
in the microvaristor particles and reduce the total filler ratio. In cable applications, carbon black (CB)
has been the dominating choice of conducting filler for conducting composites. In field grading materials
carbon black can work to improve the mechanical properties of the composites and to electrically bridge
the electro active particles [Jonsson 2007].
Electronic conductive particles (ECP) are a type of inorganic conducting particles that have been used
widely in where electric dissipative coatings are needed, i.e. electronics packaging and dissipative floors.
The particle consists of an inert core, i.e. TiO 2 or mica, coated with antimony doped tin oxide (ATO)
that can be of spherical shape or flake shaped with a dimension of a few to some tens of micrometers,
see Figure 3.28 [Merck 2012] [Rüger 2012]. This type of filler has been suggested to be used in FGM
composites with the additional benefit of improving the ability to withstand the local heat development
in a conducting path of particles compared to carbon black [Jonsson 2007]. This type of fillers offers

46
TB 794 - Field grading in electrical insulation systems

temperature stability at several hundred degrees Celsius, is independent of humidity and has a high
resistance to solvents and acids [Merck 2012] [Rüger 2012].

Figure 3.28: Flake shaped particle; mica covered by ATO [Merck 2012] [Rüger 2012]

3.3.4.5 Anisotropy
The usage of filler materials with a high aspect ratio (like the mica particles covered in 3.3.4.4) can lead
to anisotropic effects [Secklehner 2013]. For example, the filler particles might orient themselves in
direction of the flow field during the casting process or arrange themselves along the walls of the mold.
This may cause a significant increase of the conductivity in the tangential direction on or close to the
surface of an insulator. While this behavior can be useful and intended in special cases it may also cause
unpredictable and unwanted effects in terms of anisotropic electrical properties. Therefore, care should
be taken when employing filler particles with high aspect ratios. In doubt, special directional
measurements of the electric properties should be conducted to make sure that no detrimental
anisotropic effects will occur.
On the other hand, one may intentionally choose an anisotropic material in order to achieve special
effects or benefits in the production process. As an example, a compound could be exposed to an
electric field during the curing process, when the filler particles are still able to move in the insulating
matrix. The particles will then arrange along the electric field (force) lines, leading to percolating chains
of fillers in the field direction even if the overall filler content is lower than typically necessary to achieve
percolation. This effect has been utilized to create “compound varistors” with a reduced amount of filler
particles [Ishibe 2014]. This production technology could advantageously be transferred to field grading
materials as well.

3.4 Measurement metrics of field grading materials and systems


3.4.1 Introduction
In general, field grading materials used as bulk materials or coatings are based on an organic polymer
matrix filled with one or more inorganic particulate fillers and fall within the family of composite
materials. This section details the primary metrics and the measurements that are required to assess
the properties of these composite materials. Figure 3.29 summarizes the necessary metrics to
characterize a composite material.
The metrics to characterize a composite material usually depends on its composition, structure and
processability. Measurements and metrics can be derived from the main components of a composite
material, namely the polymer matrix and the particulate filler compounded with the polymer matrix to
yield the composite material.
The next sections detail the characteristics and corresponding measurements that can be carried on the
polymer matrix, the particulate filler, the composite material, and the specific electrical characterization
of nonlinear field grading composite materials. At the end of the section, a brief review of the operating
stresses at component and system levels is made.

47
TB 794 - Field grading in electrical insulation systems

Figure 3.29: Summary of composite materials components and corresponding metrics

3.4.2 Characterization of the polymer matrix


The main metrics for the polymer matrix can be defined in two main areas. First are the physical
properties of the uncured polymer (rheology) as a function of time and temperature (viscosity, reaction
parameters) but also purity, which can be controlled by a number of chemical analysis techniques (ICP-
MS, FTIR). Second are the intrinsic physical properties of the cured polymer, including electrical
properties (permittivity, tan , bulk and surface resistivity with their E-field and temperature
dependencies, breakdown strength, etc.), thermal properties such as glass transition temperature (Tg)
or melting point, and finally mechanical properties (tensile, flexural and compressive strength). A
summary of the polymer matrix’s main measurements metrics is shown in Figure 3.30.

Figure 3.30: Polymer matrix measurements metrics

3.4.3 Characterization of the particulate filler


The main characterization metrics for the particulate filler can be defined by morphology such as shape
and aspect ratio, size and particle size distribution, which can be characterized by various measurements
techniques (SEM, Laser diffraction). Purity can also be assessed by SEM, ICP-MS or FTIR). Surface
chemistry can be assessed by FTIR and other lesser known techniques like Zeta potential (electrokinetic
potential in colloidal dispersions). Particulate fillers’ intrinsic physical properties include the usual range
of electrical, thermal and mechanical properties. A summary of particulate fillers main measurements
metrics is shown in Figure 3.31.

48
TB 794 - Field grading in electrical insulation systems

Other critical aspects of particulate fillers that need to be carefully considered in relation to the polymer
matrix are:
- Particulate reaction functionality and its potential modification of the polymer reaction
stoichiometry
- Particulate surface chemistry and surface functionalization
- Particulate degree of aggregation and agglomeration in precursor liquids and polymer resins
- Water absorption characteristics
- Exothermic curing and its control under isothermal cure conditions in large sections –
dimensional effects/shrinkage (thermosetting matrices)
- Optimum cure cycle with pre-, prime- and post-cure phase

Figure 3.31: Particulate filler measurement metrics

Typically, the characterization of a ZnO microvaristor powder requires measurement of the electrical
properties of both the microvaristor powder and the same powder compounded in a matrix polymer,
e.g. silicone rubber. The AC test setups, seen in Figure 3.32, are designed with guard electrodes allowing
testing under uniform electric field distribution in the sample, to measure current as a function of voltage
[Ahmad 2015]. The same characterization can be performed under DC conditions, as highlighted in
3.4.5.2.
Initially, the voltage and current exhibit a linear relationship, with the current having a predominantly
capacitive nature. Above a certain level of applied voltage, a clear increase of the resistive component
starts to be observed; this is characterized by a sharp nonlinear increase of the current around the
instant of peak voltage. Such increased resistive current component will cause power loss and generate
heat in the test sample which, in turn, can cause an increase in magnitude of the resistive component
due to the negative thermal coefficient of the ZnO resistance. Hence, measurements at this level need
to be taken quickly, and careful consideration should be given to the risk of degradation of the material
beyond such conduction limits.
A possible technique to discriminate the capacitive and resistive components of the total current, based
on point-on-wave analysis of voltage and current waveforms, was reported in [Haddad 1991].

49
TB 794 - Field grading in electrical insulation systems

Figure 3.32: Test setups for characterization of ZnO microvaristor powder (left) and compound (right) [Ahmad 2015]

Figure 3.33 depicts the voltage and current (V-I) characteristic for microvaristor powder and the
compounded material, showing the peak values for each level of applied voltage. As expected, higher
leakage current is expected in the microvaristor powder compared with the mixed compound with the
polymer matrix, which is displaying good insulating properties. Exceptions from this behavior are
addressed in 3.4.5.2.2 and Figure 3.40.

Figure 3.33: Measured V-I characteristics for ZnO microvaristor powder and same compounded in a silicone rubber
[Ahmad 2015]

Figure 3.34 shows the resistivity-electric field characteristics for both the ZnO microvaristor powder and
the compounded material. As can be seen from the figure, the resistivity falls suddenly soon after the
knee of conduction is reached.

Figure 3.34: Measured resistivity dependance on applied electric field for ZnO microvaristor powder and same
compounded in a silicone rubber [Ahmad 2015]

50
TB 794 - Field grading in electrical insulation systems

3.4.4 Characterization of the composite material


The main characterization metrics for the resulting composite material can be outlined as follows and
summarized in Figure 3.35. They can be classified in four main properties grouping:
Physical characterization metrics (prior to polymerisation)
- Viscosity as a function of temperature and time
- Colloidal stability/shelf life by way of filler sedimentation tests
- Pot life during processing prior to gelling and curing phases
Dielectric, electrical and endurance characterization metrics
- Permittivity and tan  as a function of electric field and temperature
- Low frequency dielectric response (to inform on ionic conduction) by way of dielectric
spectroscopy
- Bulk DC conduction as a function of field strength and temperature – under dry and fixed RH
condition
- Dielectric response and effects of water exposure
- Bulk and surface DC conduction/resistivity as a function of electric field, temperature and
relative humidity
- AC or DC electrical breakdown strength by way of short-term rapid voltage rise for
comparative performance (preferably with thin films < 0.5 mm)
- Space charge under poling DC fields and reverse polarity
- Surface arcing/discharge erosion resistance
- Voltage endurance and onset of partial discharge (PD) activity
- Long-term performance by way of electro-thermo-mechanical accelerated ageing tests
Thermal characterization metrics
- Glass transition/softening temperature of the polymer matrix before and after the addition of
particulate fillers
- Thermal conductivity as a function of temperature up to the glass transition temperature
- Coefficient of Thermal Expansion (CTE) – as a function of temperature from -40C up to the
glass transition/softening temperature
- Thermal decomposition stability and long term weight loss
- Oxidative stability as measured by short term Thermo-Gravimetric Analysis (TGA)
Mechanical characterization metrics
- Tensile, shear and flexural modulus
- Tensile strength as a function of temperature
- Yield strength or stress for nonlinear deformation
- Brittle fracture stress and strain
- Impact fracture energy
- Fracture toughness

51
TB 794 - Field grading in electrical insulation systems

Figure 3.35: Composite sample measurement metrics

Next section deals with the specific fundamental principles of characterization of field grading materials
and associated measurement considerations.

3.4.5 Electrical characterization of nonlinear field grading composite materials


The electrical characterization of field grading materials consists of two main elements: choosing
appropriate models and termini to describe the electrical characteristics, and being able to measure
these characteristics in the first place. The measurements can be of interest for ready-to-use field
grading materials, which are often the result of some form of compounding an organic polymer matrix
and filler materials, and for the filler materials itself, as would be convenient for a filler manufacturer.
In the latter case it should be emphasized that the characteristic was obtained on a raw filler sample.
In the following section, appropriate models and parameters for characterizing materials with electrically
nonlinear conductive properties are introduced and methods for measurements on filler materials as
well as compounds made out of them are described.
The focus lies exclusively on the conductivity of the materials, as it is of major importance for nonlinear
materials with varying conductivity used in nonlinear resistive field grading. For capacitively graded
devices the relevant information is given in 2.2.4.2. It should be noted that also materials with nonlinear
permittivities exist and may be used for field grading purposes. However, they will not be further
addressed here.

3.4.5.1 Descriptive models and terminology for nonlinear resistive


materials
As has been shown in 3.3, the different nonlinear field grading materials are characterized using varying
approaches, as there is no standard available. For future work a commonly agreed modelling approach
or at least a shared nomenclature would be helpful. Therefore some proposals on how to describe such
materials are made in the following.
An exemplary σ-E-characteristic of a nonlinear resistive field grading material is shown in Figure 3.36.
From the standpoint of an insulation system designer conductivity in dependance of electric field
strength is the most valuable information on such a material.

52
TB 794 - Field grading in electrical insulation systems

Figure 3.36: Exemplary σ-E-characteristic of a nonlinear field grading material; explanation of terms in 3.4.5.1.1
[Secklehner 2017a]

Conductivity of nonlinear field grading materials can vary strongly in dependence of the field strength.
Especially the shape of the σ-E-characteristic may vary, e.g. in steepness or absolute conductivity. To
compare different materials the typically applied terms are described in 3.4.5.1.1, and a generalized
approach to describe this kind of curve is presented in 3.4.5.1.2, based on the example given in
Figure 3.36.

3.4.5.1.1 Terminology
Base conductivity σ0
Many materials show a rather flat shape of their σ-E-characteristic for low field strengths. This is often
the range of field strength and conductivity that is chosen as continuous operating point of the materials
in the later insulation system. This field strength range may be referred to as the leakage current region
and is marked with  in Figure 3.36. Conductivity in this range is described using the term “base
conductivity”, σ0.

Switching field strength


Nonlinear materials typically exhibit a sharp increase of the conductivity above a certain field strength,
which is called the switching field strength E1 (in the literature sometimes Eb ). While many describing
models make use of the switching field strength to describe the most abrupt change in the σ-E-slope,
the term sometimes also refers to the field strength required to achieve a certain current density in the
field grading material (compare [Christen 2010] [Secklehner 2017a]). Therefore, this value should be
handled with care when found in datasheets or in the literature.
In the following the switching field strength E1 will be used only to describe the point of the most abrupt
change in the -E-characteristic, which is also a general recommendation how to handle this parameter.
In the context of Figure 3.36 the parameter Eb is simply used as a reference point to determine (together
with Ea) the slope of the characteristic curve.

Degree of nonlinearity
The region of the σ-E -characteristic above the switching field strength is called the “breakdown region”
or “nonlinear region” and marked with  in Figure 3.36. The high degree of nonlinearity in this region
is often described using the exponent of nonlinearity , which can be derived using the following
relationship:

53
TB 794 - Field grading in electrical insulation systems

lg (𝐽b ⁄𝐽a )
𝛼=( )
lg (𝐸b ⁄𝐸a )
Equation 3.4

where 𝐽b , 𝐸b , and 𝐽a , 𝐸a are pairs of values of current density and field strength in the nonlinear region.
Since the highly nonlinear region can often be described sufficiently using a single tangent of the σ-E -
characteristic, the nonlinearity can also be expressed using the slope of nonlinearity 𝑚, which can be
directly derived from the σ-E-characteristic using Equation 3.5 [Secklehner 2017a] [Hussain 2017].
𝜎b
lg
𝜎a
𝑚=
𝐸b − 𝐸a
Equation 3.5

Saturation field strength


Typically, nonlinear materials do show a saturation of the field dependent conductivity. This can be of
interest when calculating the field in the event of overvoltages and can be characterized by the
“saturation field strength”, Esat or E2 and the “saturation conductivity”, 𝜎sat .

Temperature dependance
Figure 3.36 does not highlight the fact that nonlinear resistive materials also show a distinct temperature
dependent behavior. This can be accounted for by specifying the activation energy Wa as described in
3.1.1. It should be noted, however, that this temperature dependency varies with different field
strengths and is typically most pronounced below the switching field strength.

Permittivity
Also, the permittivity of such material is of interest and should be stated if known. While permittivity
often is measured via dielectric spectroscopy for a large range of frequencies at a fixed voltage, it seems
possible that the value may also depend on the applied field. Further research on this phenomenon
would therefore be useful. Additionally, permittivity is often temperature dependent.

3.4.5.1.2 Descriptive models of the -E-characteristic


Generalized Three-Tangent-Model
To describe the -E-characteristic over the full range of field strength a generalized model using three
tangents has proven feasible and will be described in this section. This model has been proposed in
[Blatt 2015] and was further developed in [Secklehner 2017a] [Secklehner 2017b] [Hussain 2017].
Generally, there are two representations of the Three-Tangent-Model as shown in Figure 3.37.

54
TB 794 - Field grading in electrical insulation systems

Figure 3.37: Generalized Three-Tangent-Models with constant base conductivity σ0 (blue) and with slightly nonlinear
base conductivity σ0 (red) [M. Secklehner]

The σ-E-characteristic is divided into three regions as described in 3.4.5.1.1 and 3.4.5.1.2. The
mathematical description of this model is given in Equation 3.6 [Secklehner 2017b] [Hussain 2017]:
𝑁1
√1 + 10𝑚𝑁1(𝐸−𝐸1)
𝜎(𝐸) = 𝜎0 ⋅ 𝑁2
√1 + 10𝑚𝑁2(𝐸−𝐸2)
Equation 3.6

N1 and N2 are shaping parameters and can be set to 1 in the first step of design. Equation 3.6 is sufficient
for materials, which are independent of field strength in region . If the investigated material shows a
nonlinear dependency on field strength even in this region Equation 3.7 should be used:
𝑁1
√1 + 10(𝑚−𝑚0 )𝑁1(𝐸−𝐸1 )
𝜎(𝐸) = 𝜎0 ⋅ 𝑁2 ⋅ 10𝑚0 𝐸
√1 + 10𝑚𝑁2(𝐸−𝐸2 )
Equation 3.7

With the help of parameter m0 it is possible to describe a slight nonlinearity in region  as shown in
Figure 3.37. When setting m0 = 0 Equation 3.7 simplifies to Equation 3.6.
Further models that are often encountered in the literature are variations of the power law model that
will be described in the following.

Power-Law
To show how nonlinear field grading materials are able to distribute the field stress over the full length
of an insulating structure it is sufficient to describe only the nonlinear part of the characteristic curve of
the material. For this approach a simple power law as given by Equation 3.8 has proven to be feasible
[Rhyner 1997]:
𝐸 𝛼−1
𝜎(𝐸) = 𝜎0 ⋅ ( )
𝐸1
Equation 3.8

55
TB 794 - Field grading in electrical insulation systems

Figure 3.38: Field-dependent conductivity modelled by a simple power-law

As can be seen in Figure 3.38 this modeling approach does neither include the base conductivity, nor
the saturation of the conductivity. Therefore, it is interesting for the sake of theoretical derivation or
algebraic solving of exemplary problems but not for an overall description of a nonlinear material.

Power-Law with switching field strength


Also often encountered in the literature is a power law that includes a switching field strength, as can
be seen in Equation 3.9. Using this approach it is possible to implement the characteristic curve in
simulations that are meant to validate the performance of a field grading material in stationary as well
as in transient overvoltage operation. However, modeling the base conductivity as a constant value is
still a simplification.
𝐸 𝛼−1
𝜎(𝐸) = 𝜎0 ⋅ (1 + )
𝐸1
Equation 3.9

Power-Law with switching field strength and saturation


The power law including a switching field strength can further be modified to also include saturation of
conductivity (compare Equation 3.10). This model yields results that are comparable to the Three-
Tangent-Model and can, therefore, be useful for simulations over the full field strength range of interest.
𝐸 𝛼−1
1+( )
𝐸1
𝜎(𝐸) = 𝜎0 ⋅ 𝛼−1
𝐸
1+( )
𝐸sat

Equation 3.10

tanh-model
For sake of completeness it should be noted that there are several more models in use to describe
nonlinear materials, which will not all be covered in detail here. In the materials science community
often tanh-equations are used, examples of the literature are [Vojta 1996] und [Zhao 2007]. One
example is given in Equation 3.11.
𝜎sat
𝜎(𝐸) = 𝜎0 + ⋅ (1 + tanh 𝑠(𝐸 − 𝐸sat ))
2
Equation 3.11

This equation yields results comparable to the Three-Tangent-Model, but makes use of the parameter
s. This parameter combines the information of the switching field strength E1 and the degree of
nonlinearity and, therefore, makes this modeling approach less convenient for practical use.

56
TB 794 - Field grading in electrical insulation systems

3.4.5.2 Measurements necessary for the characterization of nonlinear


field grading materials
To obtain useful characteristic curves for the design and simulation of insulation systems it is important
to measure the electrical characteristics over the full range of field strength of interest. For the designer
of such system the electrical characteristics of the resulting field grading material is of interest, while
the manufacturer typically measures the characteristics of the unprocessed filler material. The following
section includes recommendations for both cases, since much of the methodology is transferrable, and
highlights the differences if present.
Another major difference is the question if the material targets AC or DC applications, since the
measurements in the leakage current region should be conducted correspondingly.

3.4.5.2.1 Sample preparation and conditioning


For measurements in the leakage current region the specimen should be contacted using guarded
electrodes. Depending on the matrix material the electrodes might be made out of conductive varnish,
like silver loaded paint or graphite spray. For applicable electrode materials and arrangements the
standards for conductivity measurements (IEC 62631-3-1 [IEC 2016] and IEC 62631-3-2 [IEC 2015])
can be consulted. Measurements in the nonlinear and saturation region can be conducted with only two
electrodes since stray currents are of lower significance here.
For measuring conductivity of filler materials in powder form a test cell as described in [Secklehner
2017b] may be used, see Figure 3.39.

Figure 3.39: Left: test cell for the powder measurements (left); Right: test cell containing microvaristor powder, the
grey bolt on the side acts as a plug to enable filling with silicone oil [Secklehner 2017b]

Insulating fluids may be applied to fill the air gaps in the powder specimen, which may avoid partial
discharges when measuring at higher field strength. However, some insulating fluids have turned out
to change the measured conductivity significantly. This should be controlled by comparative
measurements at low field strength. If no insulating fluid is added the powder can be compacted by
means of a shaking table with defined vertical oscillation and application time duration. This practice
reduces deviations due to different compactness and is, therefore, recommended [Secklehner 2017b].
Contact weight or pressure applied to the powder specimen can have further impact and should be
noted in reports or data sheets. On the whole, an exact documentation of the many parameters possibly
influencing conductivity measurements on filler powder is definitely required, which unfortunately is not
common practice today and makes comparison of fillers or prediction of the final product properties
rather difficult if not even impossible.
For precise measurements in the low field region, which are particularly important for the design of DC
insulation systems, it is recommended to subject the specimen to a conditioning period. This means
that specimen and electrodes should be exposed for sufficiently long time to the surrounding conditions
of the later measurement to allow for thermal and electrical settling. The current driven by the specimen
in a short circuit loop during conditioning should be recorded and serves as a hint for sufficient
conditioning time [Secklehner2017b].

3.4.5.2.2 Conductivity measurements


Low field region
Conductivity in the low field region should be measured using the kind of voltage (AC, DC) for which
the application is intended, since the characteristic curves typically differ for AC and DC measurements

57
TB 794 - Field grading in electrical insulation systems

[Debus 2015]. For AC measurements measurement over a couple of seconds is usually sufficient, yet
there may exist a small time dependance of the measurement. It has proven useful to not only construct
a characteristic curve from the peak values of current and voltage, as it is typically done in the varistor-
literature, but to preserve the information about the conductivity as well as the permittivity. This can
be done either by deconstructing the total current waveform into a capacitive and a resistive part [Debus
2015], [Blatt 2015b], by means of the “point on wave method” [Ahmad 2017], or by giving the D-E-
relationship in the form of a hysteresis curve [Christen 2010]. The method used for the separation into
capacitive and resistive part of the total measured current should be noted when showing a
characteristic curve.
The measurement time for a pair of σ-E-values of a DC measurement depends on the material to be
measured. It is recommended to conduct at least one measurement over a longer period of time to
identify a suitable time interval to allow for polarization to take place. A common recommendation is to
use intervals of 10 minutes [Secklehner 2017b]. The measurement time should be documented for
reproducibility.
Generally, the voltage shape of the measurement and the time interval between voltage application and
measurement should be noted in the report or data sheet. Further, it has been shown that the
environmental conditions, particularly humidity and temperature, can strongly influence the results
[Secklehner 2017b]. They should, therefore, be either well controlled or at least be recorded and
documented.
The measurement methodology does not differ between specimens in the form of pure powders or filled
compounds, aside from limitations in the applicable field strength. However, since the behavior of the
filler might be altered by the addition of a polymer matrix, it should be noted and clearly stated if the
measurement was obtained on a raw filler or a compound material.

Nonlinear and saturation region


Since conductivity of nonlinear field grading materials above the switching field strength can be quite
high, power losses may lead to thermal runaway of the samples when measured with DC or AC.
Therefore, conductivity in this region should be determined by means of short impulses, which is also
a suitable replication of the stress in real systems. Amplitude and impulse duration should be adjusted
to not energetically overstress the material. Also the rise time should be slow enough to allow for a
resistive current to flow, meaning that the current peak should be synchronous to the voltage peak and
distinctly exceed the capacitive current component (compare Figure 3.40, right side). Under this
condition the peak values of voltage and current can be used to determine a pair of values for the σ-E-
characteristic. Adding the information gained by the impulse measurements allows a more
comprehensive insight into the field grading material, which can be seen in Figure 3.40, left side.
Another effect can be seen in Figure 3.40: The conductivity of the composite is one order of magnitude
higher than that of the filler. This unexpected effect has also been reported e.g. in [Donzel 2011]
[Secklehner 2019]. So far there is no commonly accepted explanation for it. It has been assumed that
it might be caused by an increased contact pressure from the shrinking matrix material during the curing
process, or by chemical reactions with the matrix material. Unfortunately, due to such unpredictable
effects a precise forecast of the electrical properties of a composite with nonlinear filler just from the
electrical characteristics of the filler itself is impossible, and the filler as well as the final composite have
both to be independently characterized as a consequence.

58
TB 794 - Field grading in electrical insulation systems

Figure 3.40: left: σ-E-characteristic over the full field strength range of interest, the pairs of values above 1.5 kV/mm
were obtained by impulse measurements, right: current response for voltage impulses of different amplitudes, ranging
from merely capacitive (upper diagram) to virtually merely conductive (lower diagram) [M. Secklehner]

3.4.5.3 Design example of a tailored field grading material


This chapter demonstrates how to develop a desired specific σ-E-characteristic as shown in Figure 3.36.
The goal is to find all parameters necessary to describe conductivity according to Equation 3.6. The
object of investigation is a generic HVDC cable joint, whose principle design is shown in Figure 3.41.
More details on this example are given in [Hussain 2017].

Figure 3.41: Generic 320 kV HVDC cable joint with (nonlinear) resistive field grading; 1…conductor, 2…XLPE
insulation, 3…conductive VMQ shield, 4…metallic connector, 5…joint VMQ insulation body, 6…nonlinear
semiconducting grading layer, 7… outer sheath of joint and outer semicon of cable [Hussain 2017]

Main focus here is the design of the nonlinear semiconducting field grading layer (Figure 3.41 (6)).
According to Figure 3.36 the first step is to find the allowed boundaries for the base conductivity, σ0.
The aim is to achieve “decoupling” between the cable insulation (2) and the joint insulation (5). It was
found in [Secklehner 2017a] that the lower limit of the base conductivity can be determined by Equation
3.12
𝜎0 ≥ 100 ⋅ max{𝜎adjacent,𝑖 },
Equation 3.12

where σadjacent,i are the conductivity values of materials directly adjacent to the field grading material.
The upper limit of base conductivity comes from thermal constraints of the involved insulating materials
and can be determined with stationary thermal simulations as described in [Secklehner 2017a]. Next
step is to define a minimum desired switching field strength E1 indicating the transition to the high
nonlinear region. As a rule of thumb, E1 must be at least as high as the highest occurring field strength
during stationary operation. For a cable joint this value can be determined by dividing the applied DC
voltage by the distance of high voltage (HV) and ground (GND), shown as “d” in Figure 3.42.

59
TB 794 - Field grading in electrical insulation systems

Figure 3.42: Determination of minimum required switching field strength E1 [Hussain 2017]

In the transient region  the conductivity values of the field grading material must satisfy Equation
3.13 so that the system stays in a resistive flow field distribution:
𝜏
𝜎≥
ε0 𝜀r
Equation 3.13

The relaxation time constant τ is set to the value of the fastest transient time values, such as switching
(SI) and lightning impulse (LI) voltages. In [Secklehner 2017a] the permittivity of the field grading
material was set to εr = 10, and τSI = 250 µs as well as τLI = 1.2 µs. The obtained minimum conductivity
values σSI and σLI where determined according to Equation 3.13 and drawn into the diagram (Figure
3.43). The associated field strength during switching and lightning impulse are calculated according to
[Cigre 2012]. As mentioned, with Equation 3.13 the minimum necessary conductivity during transient
operations is determined, but there are also upper limits. These limits are due to the temperature rise
during transient overvoltage stress and the required ability of the material to remain thermally stable
afterwards. They can be determined by fully coupled electrical and thermal simulations as described in
[Hussain 2017].
With all determined parameters it is now possible to find a σ-E-characteristic that fulfils the
requirements. One possible solution is shown in Figure 3.43.

Figure 3.43: Nonlinear σ-E-characteristic for the field grading material with all design parameters according to [Hussain
2017], based on [Christen 2010]

In [Hussain 2017] the saturation field strength E2 was just set to a value 10% higher than the maximum
field strength during a standard lightning impulse ELI. Preferably, E2 should be at least equal to ELI. With
the help of the calculated lower limits of conductivity during transient operation it is now possible to
calculate the desired slope of nonlinearity m. All derived parameters are summarized in Table 3.3.

60
TB 794 - Field grading in electrical insulation systems

Table 3.3: Summary of all parameters of Equation 3.6 [Hussain 2017]

S
Base conductivity 𝜎0 ∈ [10−12 ; 5 ⋅ 10−9 ]
m
kV
Switching field strength 𝐸1 = 0.7
mm
mm
Slope of nonlinearity 𝑚 ∈ [4.6; 5.4]
kV
kV
Saturation field strength 𝐸2 = 1.1 ⋅ 𝐸LI = 2.3
mm

Finally, the behaviour of a linear and a nonlinear field grading material are compared in Figure 3.44 in
case of a standard lightning impulse. The field strength at the time instant of the impulse peak is
evaluated at the tangential interface between cable insulation and joint insulation (see Figure 3.42). It
is obvious that the field grading material with constant conductivity is not able to suppress the electric
field strength at the connector and the outer sheath (blue curve). Reason is that the relaxation time
constant according to Equation 3.13 is τ ≈ 0.9 s, which is excessively high compared to the time constant
τLI, and the field distribution is capacitive rather than resistive. In contrast, the previously designed
nonlinear material behaves as desired and limits the tangential field strength (red curve) to acceptable
values.

Figure 3.44: Comparison of the field strength at the boundary layer during a standard lightning impulse 1.2/50 at the
time instant of the impulse peak; red: nonlinear field grading layer according to Figure 3.43 and Table 3.3; blue: linear
field grading layer of conductivity 10-10 S/m [Hussain 2017]

3.4.6 Operating stresses at component/equipment levels


Characterization and metrics in relation to components and equipment containing stress grading
materials are usually dealt with under the relevant international apparatus standards they are used
under. However, for guidance, the following general guidelines for the validation and qualification of
components and system containing field-grading systems are given hereunder and should be carefully
considered given the range of operating stress they might be subjected to.
Electrical tests are usually focused on proving tests under AC and/or DC stresses, usual as withstand
tests, under wet and dry conditions as appropriate as well as a range of transient voltage withstand
tests, which can be listed as follows, usually as pass/fail tests:
- AC and/or DC short term voltage withstand test, usually for a period of 1 minute
- Capacitance, dielectric losses (tan) tests can be undertaken, usually at power frequencies under
AC depending on the component/equipment standard
- Partial Discharge characterization, usually before and after the short-term voltage withstand

61
TB 794 - Field grading in electrical insulation systems

- Lightning impulse voltage withstand tests (LI)


- Switching impulse voltage withstand tests (SI)
- Other types of transient voltage withstand tests can be performed depending on the type of
application ranging from temporary, slow, fast all the way to very fast front overvoltage tests
(TOV, SFO, FFO, VFFO)
Mechanical tests are also performed, but once again they are component/system dependent but usually
contain a mixture of tensile, flexural and compression tests.
Thermal tests include glass transition measurements (Tg) to validate the integrity of the composite
materials used in the component/equipment. Temperature rise tests are, however,
components/equipment specific thermal tests to validate withstand to a specific operational temperature
profile seen by the component/equipment.
Endurance tests involve the combination of usually two or more stresses applied to the component over
a period of time and are component/equipment dependent with the aim of at least one parameter
providing a level of time acceleration to attempt to replicate in a representative manner and in a
relatively short test duration the stresses that the component or equipment needs to be subjected to
over its lifetime.
Figure 3.45 shows a summary of the possible tests and measurements to be considered at the
component/system level.

Figure 3.45: Summary of component/system measurement matrix

62
TB 794 - Field grading in electrical insulation systems

4. Simulation of field grading systems


4.1 Purpose and requirements
Field grading systems are electrically described by the Maxwell Equations. The solution of these
equations gives the spatially distributed electrical voltage and the related electric field strength,
respectively. For basic configurations, such as bushings or cable terminations, as well as simplified
geometries or material properties, an analytical solution is possible [Thienpont 1964] [Kimura 1986]
[Kogan 1995]. However, for nonlinear material properties or more complex field grading systems, an
analytical solution cannot be derived anymore. Therefore, the two approaches presented below are
often used for precise assessment of a field grading system.

4.2 Transient Network Analysis approach


A TNA or distributed network model describes a real geometry by means of an equivalent electrical
circuit consisting of concentrated or lumped electrical elements, such as resistors, inductors and
capacitors, derived from the physically existing resistances, inductances and capacitances. In general,
a thermal analysis is also possible by use of thermal resistance and capacitance elements.
The challenge is to build an equivalent circuit model representing the physical behavior of the system
in a sufficiently accurate way. In the Figure 4.1 a typical distributed network model representing the
slot exit area of a generator winding with a resistive coating (R s) as field grading is shown.
Based on such network models all relevant currents and voltages can easily be derived in the time- or
frequency domain by means of modern calculation tools, such as “PSpice” or “Proteus”. Normally, a
solution in the frequency domain is sufficient and convenient, although evaluation in the time domain
is often used in case of non-sinusoidal electrical excitations or nonlinear material properties.

Figure 4.1: Distributed network model of a typical end-turn configuration of a high


voltage coil [Kogan 1995]

In special cases or for defined boundary conditions it is possible to provide a differential equation system
describing the network model mathematically. Like for the distributed network model presented above,
it is possible to obtain analytical solutions [Kogan 1995]. Therefore, resistance, inductance and
capacitance are transferred to their corresponding values per unit length. After that, the partial
differential equations can be formed and analytically solved [Thienpont 1964] [Xie 1986] [Rhyner 1997].
Basically, it is also possible to consider coupling, i.e. temperature or field depending material properties,
by introduction of iteration cycles.
There are several disadvantages linked to the TNA approach. The accuracy of the results depends on
the accurate determination of the lumped element values for resistance, inductance and capacitance.
In addition, representing complex systems by means of distributed elements may introduce errors. It
might be difficult to precisely calculate resulting field strength distributions from the voltage solution.
Due to these disadvantages as well as the increasing power of computers and developments in
numerical field simulation techniques nowadays the TNA approach is not often used any more.

63
TB 794 - Field grading in electrical insulation systems

4.3 Numerical field simulation approaches


Numerical field simulation offers the possibility to derive every physical parameter in complex and
spatially extended geometries. This is done by solving the related partial differential equations, e.g.
Maxwell Equations for electrical or Heat Equations for thermal models, with the help of numerical
techniques, such as the finite element method (FEM) or the finite difference method (FDM).
In a first step the geometry of the electrical system to be analyzed is transferred into a computer model,
see Figure 4.2 as an example. It is reasonable to simplify the geometry, especially details without an
expected major impact on the results, to save computation time and power.

Figure 4.2: Transferred model of a 3 m long overhead line insulator geometry


(left) before and after meshing (right) [Weida 2011]

In the next step describing partial differential equations are assigned, material parameters are set and
boundary conditions are defined. In most software tools for numerical simulations a database with all
necessary equations are predefined. In Figure 4.3 the related electrical and thermal boundary conditions
are highlighted (in green) on the model of a generator bar equipped with nonlinear field grading layers.
Accuracy of the simulation results strongly depends on precise values of material resistivity and

Figure 4.3: Illustration of assigned electrical and thermal boundary conditions (in green colour)
on a model of a generator bar [Staubach 2018a]

permittivity for electrical analysis and thermal conductivity, density and heat capacitance for thermal
calculations; nonlinearity and anisotropy have to be considered as well. In most cases electrical
potentials or charges and temperatures, convection or joule heating are defined as boundary conditions.
In a third step the model is divided into small elements on which the partial differential equations are
numerically solved afterwards. Basically, an increase of the degree of discretization will increase the
accuracy of the solution, a decrease of discretization will decrease calculation time and power.

64
TB 794 - Field grading in electrical insulation systems

In most simulation tools solvers for different modes of analysis are pre-defined. In case of field grading
applications solvers for electrostatic (DC-application), electro quasi-static (EQS; low-frequency AC-
application) or transient (surge arrester, converter fed drive) analysis are required. Typically, for EQS
analysis, the solution is given in the frequency range, i.e. as a harmonic, complex solution with real and
imaginary part [Weida 2008] [Staubach 2010]. If nonlinearity of field grading material has to be
considered, e.g. for SiC or ZnO, transient analysis may be necessary to regard resulting harmonics of
higher order. Although harmonic solutions in the frequency domain are normally derived only for linear
material characteristics, it might be useful to conduct this also for nonlinear-material behavior if the
influence of harmonic components is assessed to be negligible [Weida 2009] [Staubach 2012a]. By
doing so a lot of calculation time can be saved, which may be important for following optimization
routines [Weida 2011] [Staubach 2013].

Figure 4.4: Meshed geometry of a generator bar involute with field grading system (OCP and ECP) (left), and detailed
view on the meshed thin field grading layers (right) [Staubach 2018a]

In special cases it is reasonable to additionally include the temperature dependence of the material via
coupling of electrical and thermal simulation models [Staubach 2011].
In a last step – the post-processing – the results are visualized and can be further evaluated. With
regard to field grading common visualizations are contour or path plots of electrical potential,
temperature or electric field strength, especially the tangential component along boundary surfaces.
The maximum field strength occurring in the field grading system is used to access the electrical
efficiency of the grading system, whereas the hot spot temperature must not exceed a defined value to
prevent significant thermal aging or thermal runaway [Staubach 2018a].

Figure 4.5: Visualization of electrical field strength distribution as a colour plot (left) and along a defined
path (right) of a generator bar [Staubach 2018a]

Figure 4.5 shows the result of the electric field distribution in a colour plot and as a path plot along a
defined path. Figure 4.6 shows the same kind of visualization for the temperature. Both results are
determined by means of a coupled electrical-thermal simulation model for a nonlinear resistive field
grading coating.

65
TB 794 - Field grading in electrical insulation systems

Figure 4.6: Visualization of temperature distribution as a colour plot (left) and along a defined
path (right) of a generator bar [Staubach 2018a]

As a further example, Figure 4.7 gives the result of an electric field distribution in a colour plot and as
a path plot along a defined path of an overhead line insulator.

Figure 4.7: Visualization of electrical field strength distribution as a colour plot and along a
defined path of an overhead line insulator [Yang 2018]

4.4 Verification of simulation results


4.4.1 Measurement of potential and temperature distributions
All calculation models, i.e. analytical approach, network model or numerical simulation, need to be
validated and assessed afterwards. Non-acceptable deviations between measurement and calculation
results can be caused by several influencing factors, such as inadequate simplification of the physical
geometry, erroneous material properties and boundary conditions or the calculation method itself. By
means of measurement results iterative adjustments of the simulation model are possible, or corrective
factors can be introduced.

66
TB 794 - Field grading in electrical insulation systems

In case of field grading systems two physical parameters are of main interest to quantify their
effectiveness. To assess the electrical integrity of an apparatus the spatially distributed electric field
strength along the insulation surface is relevant. Thermal deterioration evaluation is enabled by means
of temperature distribution determination and quantification of hot spots. It is desirable to measure the
field strength distribution along the grading interface with only low interference caused by the
measuring device itself. Therefore, a very small measurement setup, consisting of dielectric materials
with low relative permittivity, is needed to guarantee high local resolution and low field distortion.
Basically, there are two common measurement principles to determine the field strength distribution.
An electrostatic voltmeter allows measuring the potential distribution φx along the grading surface
[Schmerling 2009] [Sharifi 2010]. A corresponding test setup is shown in Figure 4.8, which is a
laboratory setup to measure the voltage distribution along the field grading surface via an electrostatic
voltmeter. After receiving the spatial voltage distribution at defined locations, the magnitude of the
electric field strength is calculated by means of the difference quotient equation approach |Δ φx/Δx|.
Disadvantage of this technique is the low accuracy of local field strength enhancement detection. The
calculation of the field strength magnitude strongly depends on the step size between adjacent
measuring points. In addition, the electro-static voltmeter’s probe consists of metallic elements and
distorts the field strength distribution accordingly.

Figure 4.8: Measurement setup with voltage source (1), capacitive divider (2+3), test sample (4), grounding (5), contact
electrode (6), electrostatic voltmeter (7) and power supply of the voltmeter’s light source (8) [Schmerling 2009]

A direct, directional measurement is enabled by means of an electro-optical miniature probe. The


Pockels effect in an anisotropic crystal, e.g. LiNbO3, Bi12SiO20 or Bi12TiO20, is used for measurements of
electric field strength. The electro-optical effect in a suitable polarimetric setup is analyzed via the
polarization ellipsoid formalism together with the Jones calculus and illustrated in the Poincaré-sphere
[Merte 2007]. A small potential-free E-field probe without any metallic electrodes can be built and
calibrated using this effect [Staubach 2015], see Figure 4.9. Basically, an electro-optical miniature probe
allows determination of AC, DC and transient fields.

Figure 4.9: Principle configuration of a Pockels probe (left) [Merte 2007] and an implemented device
(right) [Staubach 2015]

67
TB 794 - Field grading in electrical insulation systems

The signal output from the measurement unit is a voltage signal and is recorded by means of a digital
storage oscilloscope for further evaluation. The magnitude of the voltage signal is directly proportional
to the electric field strength and can be derived after obtaining a calibration curve for a reference field
strength [Staubach 2015]. Figure 4.10 shows the measuring setup up with the test sample, Pockels
sensor and displacement unit. The Pockels sensor is mounted underneath a wooden stick and can be
continuously moved in axial direction to measure the spatially distributed field strength component in
axial direction close to the insulation surface.

Figure 4.10: Measurement setup for direct field strength determination via an optical Pockels sensor [Staubach
2015]

Because of the time-domain resolution of the voltage signal generated by the measurement device,
further assessment possibilities are enabled. Especially for field grading systems with nonlinear material
properties the distortion of the original sinusoidal voltage signal as well as the phase shift between the
applied high voltage signal and the Pockels sensor signal can be used for efficiency analysis, see Figure
4.11. In the presented plot the distortion is caused by the nonlinear specific resistance characteristic of
the used semi-conductive material resulting in harmonics of higher order.

Figure 4.11: Time-domain Pockels sensor voltage distortion (top) and the derived field strength distribution along the
field grading surface (ECP) (bottom) [Staubach 2015]

68
TB 794 - Field grading in electrical insulation systems

Most common measuring method for temperature mapping is infrared thermography by means of
thermographic cameras. Especially for field grading systems with linear or nonlinear material properties
this technique is used for assessment.

Figure 4.12: Transient temperature mapping on a field grading system with nonlinear resistive behaviour [Staubach
2018b]

The plots presented in Figure 4.12 show a transient temperature mapping of a field grading system
with nonlinear resistive behavior. The hotspot area with the highest thermal stress is identified. Due to
the transient measurement calculation of the thermal time constant is possible. Advantage of this
method is that no direct contact with the surface is necessary.
In principle other temperature measurement technologies are available, e.g. sensors based on
luminescence decay time. But these methods require contact to the surface, causing distortion of the
temperature profile and thus of the field grading itself.
Disadvantage of all measuring devices is the fact that they only provide temperatures of the surface
unless no constructive measures are conducted. With regards to infrared thermography adequate values
for the emission coefficients are needed to prevent erroneous temperature measurements.

4.4.2 Measurement of partial discharge (PD) inception voltages


In terms of field grading applications, the efficiency or well-function of an apparatus can be assessed
by its inception voltage of partial discharges (PD) at the grading surface, i.e. corona activity.
Furthermore, the corresponding electric field strength for inception of PD is a good measure for
evaluation of simulation results regarding the design of these apparatus, such as numerical simulation
models based on FEM. While the inception voltage is individual for each electrical configuration and may
vary a lot between the different applications, the related electric field strength is more or less design
independent (except for very small distances and radii) and follows the physics of gas discharge, i.e.
the Paschen’s Law or the streamer criterion. However, due to the nature of the surface, the
environmental conditions and the type of gaseous surrounding medium, there is some dependency on
these parameters. In general, the breakdown field strength in a gas surrounding a tangentially stressed

69
TB 794 - Field grading in electrical insulation systems

surface, like for field grading applications, is lower compared to a pure breakdown in the gas described
by the Paschen’s Law or the streamer criterion (see also 2.2.3).
Basically, there are two possibilities to measure the inception voltage in the field grading area. If this
area is directly accessible, determination of PD inception is possible by means of optical measurement
devices, such as a UV camera (also often called corona camera or finder). Voltage is applied and slowly
increased until PD is observed in the grading area, compare Figure 4.13 (left). If a UV camera is not
available a “black-out test” (i.e. switching off the light for a visual inspection) might also be an option
[Klamt 2006]. This method, however, is not that precise because sensitivity of the naked eye is lower.
Additionally, there exist some empirical equations to estimate PD-inception voltage, see 2.2.3 and
[Thienpont 1964] [Hinrichsen 2011] [Werdelmann 2009]. However, these equations are only valid for
ideal and simplified configurations and need to be experimentally verified.
Second possibility is to use a classical partial discharge measurement system to obtain the PD inception
voltage and related phase resolved partial discharge (PRPD) patterns, see Figure 4.13 (right). The
voltage is slowly increased, and the discharge activity is measured. Starting of PD in the stress grading
area will result in a strong increase of the measurement signal. In addition, the PRPD-pattern will change
[IEC 2000] [IEC 2006].

Figure 4.13: Measurement of PD inception voltage in the field grading area of a generator stator bar with an UV-
camera (left) and by means of a classical partial discharge measurement system (right) [C. Staubach; own
measurements]

With regard to PD inception voltage and the PRPD patterns in the field grading area a lot of experience
is needed for an appropriate evaluation. It becomes even more difficult in case of large spatially
extended apparatus due to damping and transit time effects or superposition with other PD sources.
To determine the PD inception field strength on boundary surfaces it might be reasonable to conduct
basic investigations on simplified geometries, which are easy to analyze by analytical approaches or
numerical field simulations. The inception voltage is measured by means of a UV camera or a PD
measurement system as previously explained. The related field strength component tangentially to the
surface in the area where PD occurs can then be calculated for this voltage.
In [Werdelmann 2009] an inception field strength of 0.64 kV/mm is determined for a clean surface in
air. This reference also provides measurement values for different kinds of surface pollution. In [Marusic
1994] [Wheeler 2005] an inception field strength of 0.6 kV/mm is stated, [Baumann 2011] discusses a
value between (0.4…0.5) kV/mm. In a theoretical approach [Thienpont 1964] an equation for inception
voltage of glow and brush discharges at the slot exits of generator stator windings is derived and a field
strength of (0.39…0.41) kV/mm determined.

70
TB 794 - Field grading in electrical insulation systems

5. Applications
5.1 Cable accessories
5.1.1 General
Power cable accessories are predominantly installed on-site after the cable is finally laid out in trenches
or has been buried underground. In order to reduce the time of installation the individual components
need to be easy to install but nevertheless sufficiently robust to withstand all kind of severe stresses
imposed by the elements, mechanical forces and especially electrical and thermal stresses [Peschke
1999].
Irrespective of whether the concerned power cable accessory is a termination or a joint, in order to
connect the cable end to an overhead line or switch yard or to connect two pieces of cable, the outer
semiconducting screen of the cable must be removed for a certain distance along the cable. Stripping
this outer screen still causes a significant electric field enhancement around the cut edge, which will
lead ultimately to a breakdown or flashover due to the creepage stress (see Figure 5.1). This field
enhancement must be handled by suitable field grading systems [Strobl 2001].

Figure 5.1: Breakdown at the cable screen edge without any field control means; photos left and right: [TE on YouTube
https://www.youtube.com/watch?v=JzbhYhlIJv4]

The exposure of the sensitive cable insulation surface represents a risk. If not protected and handled
well during installation of the accessory, it may get contaminated or even damaged [Cigre 2009]
[Higinbotham 2018]. In this respect ease of installation supports a short installation time, which reduces
not only the total cost of an installation but also the risk of premature breakdown of the cable due to
an extended time of exposure. A few decades back, where taping the geometric field grading was
popular, tents with air conditioning and filtered ventilation systems were required to ensure that no
foreign particles were trapped within the applied insulation layers of a stress cone. Nowadays, the time,
for which the cable ends are exposed to the open environment, is limited to a few minutes in case of
medium voltage accessories and up to a few hours for high voltage accessories [Eigner 2013].
Ease of installation comprises also a sufficient range taking capability 6 of the components (i.e. their
capability to adopt to varying geometries), not only because the manufactured cable insulation
diameters vary, but also because stripping tolerances must be accommodated by the accessories. With
a larger range taking capability the risk of accidental misapplication of the accesssory is further reduced.
The reliability of an accessory is determined significantly by the quality of the cable insulation surface
during installation [KEMA 2018] [Higinbotham 2018]. The electrical withstand performance of the
interface between that surface and the cable accessory’s insulation material is influenced by mainly
three interdependent parameters [Lambrecht 1999]:
- the surface roughness of the insulation materials being in contact with each other,
- the softness or flexibility of each, and
- the interfacial pressure between them.
Due to the interdependence it is possible to compensate for a weak point by adjusting the remaining
two parameters accordingly. If e.g. the shore hardness of an insulation or field grading material is
somewhat unfavourable, it is possible to compensate this disadvantage by either improving the surface
roughness and/or by increasing the interfacial pressure. The downside of the latter is, however, that it
impacts the ease of installation negatively, and too high mechanical pressure leads to creeping of the
(XLPE) insulation.

6
IEC 60502-4 and HD 629.1 S3 both use the term “range taking”; IEC definition: range taking accessory =
accessory designed for covering a range of cable cross sections

71
TB 794 - Field grading in electrical insulation systems

Nowadays mainly geometric and refractive field grading concepts are used with cable accessories. Due
to certain restrictions in materials properties and dimensions commerically available refractive field
grading concepts are limited to AC system voltage levels up to 84 kV, respectively. Beyond that
geometric concepts are dominating the market. This situation is expected to change with the wider
introduction of HVDC cable systems.
Refractive field grading concepts have been commercially available for about four decades in heat-
shrink and for about two decades in cold-applied technology [Haverkamp 1984] [Drumm 2016].
Geometric field grading concepts, however, are limited to cold-applied technologies due to the
manufacturing process on one side and the relatively large volume required to enclose the field grading
funnel at higher voltage levels. Cold-applied in a sense also includes taping, which is a technology not
being widely used any more. It required cellulose and graphite papers to be applied on oil impregnated
papers to shape the required field grading funnel. With the change to polymeric insulation cables the
technology was adapted in the beginning and corresponding self-amalgamating rubber tapes were used
instead, before finally prefabricated stress cone bodies made of suitable elastomeric materials captured
the market [Eigner 2013]. Nowadays, EPDM and VMQ are the main materials used to manufacture
stress cone bodies with embedded geometric field grading.
Heat-shrink technology is based on polymeric materials, which after extrusion are cross-linked by either
peroxide additives in the compound or by a subsequent application of an electron beam. Cross-linking
changes the materials properties in a way that the material does not melt any more but transits to a
rubbery state instead, during which its shape can be changed. Cooling to ambient conserves the
modified shape, but after re-heating, i.e. during the installation process, the material will tend to shrink
back to its initial shape after cross-linking. Polyolefines (mainly PE and EVA) are suitable compounds,
which can be cross-linked to achieve the “shape-memory” effect [Drumm 2015].
Nonlinear refractive field grading concepts based on sintered ZnO fillers (microvaristors) are
commercially available in heat-shrink and cold-applied technology [Haverkamp 2000] [Boettcher 2001]
[Donzel 2004]. For higher voltages concepts up to system voltages of 145 kV are available, which are
based on sintered ZnO fillers embedded into rigid epoxy resin housings similar to bushings [Amerpohl
2002] [Showa 2008].
Following the relevant IEC and IEEE standards for MVAC and HVAC cable accessories (for examples see
Table 5.1) the effectiveness of the field grading performance is to be verified at elevated power
frequency voltages as well as transient stresses. Temperature effects are partially considered as well,
although their focus is mainly on the thermo-mechanical stress of the components and their impact on
critical interfaces. The aspects of harmonics [Pathel 2012] and ageing of field grading materials may,
however, not be covered sufficiently by these standards, which apply mainly but not exclusively to
electrically linear materials. For electrically nonlinear materials the common practice to accelerate
degradation by increasing the electrical stress level may have to be re-considered since operating these
materials above the regular operating stress may initiate additional aging mechanisms and failure
modes, not being relevant to the application.
Table 5.1: Selection of common standards for MVAC and HVAC cable accessories and their insulating and field
grading material

IEC 60502-1 Power cables with extruded insulation and their accessories for rated voltages
from 1 kV (Um = 1,2 kV) up to 30 kV (Um = 36 kV) – Part 1: Cables for rated
voltages of 1 kV ((Um = 1,2 kV) and 3 kV (Um = 3,6 kV)
IEC 60502-2 Power cables with extruded insulation and their accessories for rated voltages
from 1 kV (Um = 1,2 kV) up to 30 kV (Um = 36 kV) - Part 2: Cables for rated
voltages from 6 kV (Um = 7,2 kV) up to 30 kV (Um = 36 kV)
IEC 60502-4 Power cables with extruded insulation and their accessories for rated voltages
from 1 kV (Um = 1,2 kV) up to 30 kV (Um = 36 kV) - Part 4: Test requirements on
accessories for cables with rated voltages from 6 kV (Um = 7,2 kV) up to 30 kV (Um
= 36 kV)
IEC 61442 Electric cables - Test methods for accessories for power cables with rated
voltages from 6 kV (Um = 7,2 kV) up to 30 kV (Um = 36 kV)
IEC 60840 Power cables with extruded insulation and their accessories for rated voltages
above 30 kV (Um = 36 kV) up to 150 kV (Um = 170 kV) - Test methods and
requirements

72
TB 794 - Field grading in electrical insulation systems

IEC 62067 Power cables with extruded insulation and their accessories for rated voltages
above 150 kV (Um = 170 kV) up to 500 kV (Um = 550 kV) - Test methods and
requirements
IEC 62271-209 High-voltage switchgear and controlgear - Part 209: Cable connections for gas-
insulated metal-enclosed switchgear for rated voltages above 52 kV - Fluid-filled
and extruded insulation cables - Fluid-filled and dry-type cable-terminations
HD 629.1 Test requirements on accessories for use on power cables of rated voltage from
3,6/6(7,2) kV up to 20,8/36(42) kV - Part 1: Cables with extruded insulation
HD 632 Power cables with extruded insulation and their accessories for rated voltages
above 36 kV (Um = 42 kV) up to 150 kV (Um = 170 kV)
IEEE-48 Test Procedures and Requirements for Alternating-Current Cable Terminations
Used on Shielded Cables Having Laminated Insulation Rated 2.5 kV through 765
kV or Extruded Insulation Rated 2.5 kV through 500 kV
IEEE-404 IEEE Standard for Extruded and Laminated Dielectric Shielded Cable Joints Rated
2.5 kV to 500 kV
IEC 60684 series Flexible insulating sleeving
IEC 60684-3-281 Flexible insulating sleeving - Part 3: Specifications for individual types of sleeving -
Sheet 281: Heat-shrinkable, polyolefin sleeving, semiconductive
IEC 60684-3-282 Flexible insulating sleeving - Part 3: Specifications for individual types of sleeving
– Sheet 282: Heat-shrinkable, polyolefin sleeving – Stress control
HD 631.1 Electric cables - Accessories - Material characterization - Part 1: Fingerprinting
and type tests for resinous compounds
HD 631.2 Electric cables - Accessories - Material characterization - Part 2: Fingerprinting
and type tests for heat shrinkable components for low voltage applications
HD 631.3 Electric cables - Accessories - Material characterization - Part 3: Fingerprinting for
heat shrinkable components for medium voltage applications from 3,6/6 (7,2) kV
up to 20,8/36 (42) kV
HD 631.4 Electrical cables - Accessories - Material characterization - Part 4: Fingerprinting
for cold shrinkable components for low and medium voltage applications up to
20,8/36(42) kV

For practical considerations and applications in HV equipment, field grading systems need to meet
specific electric stress requirements. These requirements depend on many factors. Variables that need
to be taken into consideration include material dielectric properties like breakdown strengths of the bulk
of the material and interface surfaces, frequencies and wave shapes of applied electrical stresses and
design margins.
Industry standards are typically designed to test electrical apparatus across a wide spectrum of critical
variables. For example IEEE standards for cable terminations and joints have testing requirements that
are up to four times higher than the nominal operation stress that apparatus will be exposed to in the
field. These requirements are determined by a combination of both, the theoretical calculations and
practical field service experience.
One critical variable is the dielectric breakdown strength of the insulating material. This information is
published for commercially available dieletric materials, and it is a starting point for proper selection of
insulating materials to be used in specific application.
Another important factor is the engineering margin that designers use for different applications and
types of equipment. Design margins are typically derived from the customized testing at the R&D stage
and from data pertaining to the historical performance of the equipment in the field.
As an example, dielectric breakdown strength of a commercially available grade of Liquid Silicone Rubber
(LSR) used for molding parts for MV and HV applications is in the range of (25…30) kV/mm (as measured
per IEC 60243). For this material the designer can choose the range of nominal operating AC stress in
HV applications of (4…8) kV/mm, which puts the design margin in the range of 3 to 7 times. However,
these values can vary a lot depending on the material, application and designers’ experience.

73
TB 794 - Field grading in electrical insulation systems

As it can be seen, the margin level is a critical component for determining the operating stresses, and
it needs to be developed not just for industrial frequencies (50 or 60 Hz) but also for other types of
waves that the equipment would be exposed to in the field. Most notably lightning impulse (LI) and
switching impulse (SI) voltages are the wave shapes that are widely used in the industry, and electrical
apparatus is evaluated and subjected to those voltages.
An example of how industry standards are taking electrical stress wave shapes into consideration can
be seen in the case of insulated power cables and cable accessories. Table 5.2 below provides the ratios
between LI and AC withstand voltage for different voltage classes.
Table 5.2: Calculated LI/AC ratios for maximum voltages and corresponding electrical stresses for power cables and
cable accessories acc. to IEEE and IEC standards

Voltage Class Ratio LI/AC


IEEE 48, IEEE 404 IEC 60840, IEC 62067
60 to 69 kV 2.92 3.61
110 to 115 kV 2.75 3.44
132 to 138 kV 2.71 3.42
150 to 161 kV 2.68 3.44
220 to 230 kV 2.63 3.30
330 to 345 kV 2.17 2.80
500 kV 1.78 2.67

It can be seen that the LI/AC ratio is not constant as it may be expected but inversely propotional to
the voltage class. This is related not just to allowable operating stresses under different voltage shapes
but also to the equipment ratings in the networks in terms of what overvoltage events are practically
possible.
In this example targeted electrical stresses in LI situation are between 2 and 3.6 times higher than in
AC fields, depending on voltage class and industry specification.
It is worth noting that in case of laminar insulated systems like oil-impregnated paper insulation,
electrical grading is focused on the longitudinal component of electrical stress in between the layers, as
breakdown strength along the layers can be multiple times lower than that perpendicular to the layer.
By applying this principle in case of oil-impregnated cables, the ideal field grading is derived as a double-
logarithmic curve where the tangential component of the electrical stress is constant at each layer of
the insulation, see Figure 5.2 and Equation 5.1 (according to “Short formula” [Short 1949]). Here,
operating electrical stress of the grading system is set by the LI field and the component of the stress
along the layers, as opposed to the magnitude of electrical stress in the bulk of the dielectric.

Figure 5.2: Geometry for Equation 5.1

74
TB 794 - Field grading in electrical insulation systems

𝑦
ln 𝐷c
𝑈
𝑔 = ln ( 𝐷2i )
𝑥 ln
𝐷c

Equation 5.1

with: g…longitudinal electric field stress along the stress control cone profile, U….applied voltage to
ground, Dc…conductor diameter, Di…outer diameter of cable insulation, y…actual radial distance to
center axis, x…actual horizontal distance to end of joint insulation.

5.1.2 Accessories for MVAC and HVAC extruded cables


In the Cigre Technical Brochure 89 [Cigre 1995] the different types of accessories designs available for
use on HV cables with extruded insulation for AC transmission voltages of 60 kV and above are well
described.

5.1.2.1 Terminations
In simple terms field grading of cable ends exposed to air require two aspects to be considered: Firstly,
the withstand performance of the termination material itself specifically close to the cut edge of the
outer cable screen. Secondly, but at least as important, the withstand performance of the termination’s
outer surface, which is exposed to air, and the elements including all kinds of contamination layers.
With the geometric field grading concept the stress distribution is influenced by a single, funnel shaped
electrode located near the end of the cable’s outer screen layer. The stress distribution between this
electrode and the high voltage potential is determined by the accessory’s insulation materials properties
in combination with the properties of the contamination layer at the accessory’s outer surface.
In case of a refractive field grading concept the area of influence is extended practically to the full length
of the field grading layer. This is beneficial on one side but there is one disadvantage linked with this,
which needs to be taken into account: The insulation between the field grading layer and the outer,
mostly contaminated and sometimes wet surface is typcially stressed at higher levels compared to
corresponding geometric field grading concepts. Reason for this is the higher imposed stress distribution
along the cable end in combination with the low insulation thickness usually found with refractive field
grading concepts. The advantage of the low wall thickness offered by refractive field grading systems
is that the terminations are slim, compact and easy to install. The application is however, limited to a
system voltage of 84 kV.
Elastomeric cold-applied terminations are available up to a system voltage level of 145 kV. Both push-
on and cold-shrink type accessories are available.
Figure 5.3, Figure 5.4, Figure 5.5, Figure 5.6 and Figure 5.7 illustrate some configurations of MVAC
extruded cable terminations.

Figure 5.3: Heat shrink type MV termination with refractive field grading [TE]

75
TB 794 - Field grading in electrical insulation systems

Figure 5.4: Cold-applied MV termination with geometric field grading [NKT]

Figure 5.5: Cold-applied MV termination with refractive field grading [Südkabel]

Figure 5.6: Heat-shrink 84 kV termination with 2 layers of refractive field grading [TE]

Terminations with nonlinear refractive field grading were introduced first for MVAC applications in 1999
based on heat-shrink technology. Sintered ZnO microvaristor powder is embedded into a hot-melt
compound and co-extruded inside a heat-shrink tube, which serves as the outer protection of the
termination.

Figure 5.7: Heat-shrink MV termination with integrated nonlinear refractive field grading layer based on a hot-melt
compound and embedded ZnO microvaristors [TE]

Other development studies [Donzel 2004] [Gramespacher 2003a] [Gramespacher 2003b] have shown
that cold-applied ZnO microvaristor-based nonlinear-resistive field grading tubes can be used up to a
voltage level of 84 kV. A prototype of a flexible cold shrink termination with elastomeric ZnO
microvaristor filled refractive field grading is given in Figure 5.8.

76
TB 794 - Field grading in electrical insulation systems

Figure 5.8: Prototype of a dry, cold-shrink 84 kV HVAC cable termination with integrated refractive field grading based
on nonlinear ZnO microvaristors (dark layer extending to the left) [Donzel 2004]

Terminations for HVAC & EHVAC extruded cables are traditionally made with a rubber stress cone
slipped over the prepared cable and housed in a supporting hollow insulator (porcelain or composite),
which is filled with an insulating oil, gel or SF6 gas (see Figure 5.9).

Figure 5.9: Cold-applied 145 kV liquid-filled termination with composite housing; the composite housing can be also
filled with gel or SF6 gas [TE]

Due to maintenance and installation problems with oil filled cable terminations, dry type cable
terminations were developed [Streit 2011] [Amerpohl 2002]. In this case, the prefabricated cable stress
cone is plugged/compressed into the cavity of an epoxy body. An example is shown in Figure 5.10.
An additional ZnO microvaristor nonlinear resistive field grading layer can be applied at the external
surface between the hollow insulator (wet type) [Boettcher 2001] or the epoxy body (dry type)
[Amerpohl 2002] and the outer silicone rubber housing to control the high electric field stresses at the
termination vicinity and to allow a slim construction.

Figure 5.10: Epoxy based 145 kV HVAC dry type outdoor cable termination with an additional outer nonlinear resistive
field grading layer [Amerpohl 2002]

77
TB 794 - Field grading in electrical insulation systems

5.1.2.2 Joints
Field grading of joints is from a technical aspect the most demanding, because the insulation distances
including the critical interfaces are smaller than for terminations. On the other side only internal fields
within a relatively simple geometry are to be considered. In addition to the field enhancement at the
screen cut suitable field grading is also required at the connector. For applications, which require the
cable screens of the two cables to be kept separated, additional field grading components may be
required on the outer screen of the joint body. Relevant applications are cross-bonded cable systems
and split single point grounded cable systems.
Commercially available products again incorporate either geometric or refractive field grading concepts.
The shape of the geometric field grading funnel is, compared with the funnel used on terminations,
typically shorter but opens wider at the same time. Besides the field grading funnel a Faraday cage is
used to cover the connector area, where air filled voids at the end of the cable insulation cannot be
avoided.
Joints with refractive field grading are available up to AC system voltages of 72.5 kV. At higher system
voltages geometric field grading is used exclusively. Figure 5.11, Figure 5.12, Figure 5.13, Figure 5.14
and Figure 5.15 show typical examples of MVAC and HVAC extruded cable joints.
With increasing wall thicknesses of cold-applied joint bodies, the mechnical forces required to install the
bodies on the cable ends increases significantly. This requires hoist chains and pulling tools to be used.
Another way of reducing the forces is to split the joint body into several, typically three pieces. Doing
so not only reduces the required installation forces but allows more space for the connector in the
center, and it provides the possibility to join cables with significantly different insulation diameters. The
later is often the case when an aluminium cable shall be jointed with a copper cable having the same
ampacity.
Another way to ease installation is to use pre-expanded joint bodies and to put them onto a hold-out
system which is removed during the installation process. Elastomers for these kind of accessories need
to provide a low tension set at high expansion ratios over the whole shelf-life in order to ensure a
suffiently high interfacial pressure when installed on a cable from the lower end of the application range.
The requirements on the mechanical stability of hold-out systems are quite high and increase with
increasing voltage levels. Nevertheless, cold-shrink joints are available up to a system voltage level of
245 kV.

Figure 5.11: Typical MVAC and 72.5 kV AC heat shrink joint (inline version) with refractive field grading. For HVAC
applications an additional Faraday cage may be applied on the outer surface of the refractive field grading sleeving
[TE]

Figure 5.12: Typical MVAC cold-applied joint (inline version) with geometric field grading [Südkabel]

78
TB 794 - Field grading in electrical insulation systems

Figure 5.13: Typical HVAC prefabricated one piece cable joint (inline version) with field grading funnels at the end and
a Faraday cage covering the connector in the center [Brugg, Prysmian]

Figure 5.14: Typical HVAC prefabricated three piece cable joint with field grading funnels at the end and a Faraday
cage over the field grading funnels in the center [nkt]

Figure 5.15: HVAC Prefabricated composite cable joint. Two rubber stress cones are compressed over the cable core
and into a central cast epoxy insulator [Prysmian]

5.1.3 Accessories for HVDC extruded cables


In HVDC extruded cable accessories, the electric field is controlled through a combination of a resistive
and a geometrical grading. The resistive field grading covers the HVDC operating stresses whereas the
geometric field grading is typically provided to take care of transient stresses. The resistive field grading
material’s properties can be either linear or nonlinear, but the latter is preferred.

5.1.3.1 Linear resistive electric field control


To work satisfactorily, a purely linear resistive field control layer should possess a specific defined
resistivity, which is several orders of magnitude lower than that of the cable insulation but high enough
to avoid excessive power losses (see also 3.4.5.3). Such field control layer consists in general of a rubber
matrix: silicone rubber (VMQ) or ethylene propylene diene rubber (EPDM), filled with carbon black
particles.

79
TB 794 - Field grading in electrical insulation systems

The possibility of the use of this type of resistive field control in HVDC extruded cable terminations is
mentioned in the patent US7674979B2 [Nexans 2010] for operating voltages of the order of 150 kV and
more. The proposed resistive insert has an electric resistivity, which is several orders of magnitude
lower than that of the cable insulation for a given applied electric field (5 kV/mm to 50 kV/mm) and a
given temperature (20 °C to 90 °C). It can lie in the range (1011…1016) Ωcm, e.g. close to 1012 Ωcm
(recommended in [Nexans 2010]), whereas the resistivity of the cable insulation is about 1016 Ωcm
approximately.
However, the HVDC cable accessories products (joints and terminations), which are so far available in
the market, are not provided with linear resistive grading layer. Indeed, although the finite element
method (FEM) calculation of the electric field distribution shows that the enhanced electric field at the
cut edge can be lowered in the presence of the resistive field grading material layer compared to the
case without having it, this reduction is not always sufficient under the application of very high operating
voltages and the required test voltage levels.

5.1.3.2 Nonlinear resistive electric field control


To date commercial DC cable accessories for voltage levels up to 320 kV utilize nonlinear resistive layers
to control the electric field under DC operation and which are able to withstand an appropriate electric
field level when subjected to transient voltages (lightning and switching impulses) [Ericsson 2003]
[Jeroense 2008] [Dodds 2010] [Worzck 2013].
According to latest new developments [Gustafsson 2015] [Jeroense 2018], the use of this nonlinear
resistive field grading layer has been extended to 525 kV and 640 kV cable accessories where the
complete HVDC cable system is reported to have successfully passed the recommended CIGRE TB 496
[Cigre 2012] qualification tests. Indeed, to verify that the HVDC cable system, which comprises the
cable and its accessories (joints and terminations), works properly, two types of tests are generally
accepted as proof of a proper function. These tests are referred as short-term (one month) type test
(TT) and long-term (one year) prequalification test (PQT) that should both be successfully passed. The
voltage during the type test is 1.85 x Uo, where Uo is the rated DC voltage between the conductor and
core screen, for which the cable system is designed. The voltage at the prequalification test is 1.45 x
Uo. The maximum conductor temperature used during the test is 70 °C simulating the cable under load.
The technology behind the DC cable accessories development is well highlighted and explained in
[Sorqvist 2009].
The cable joint is shown in Figure 5.16. It consists of four different material layers. An inner deflector
made of a semi-conducting rubber electrically shields the connector. Through the whole length of the
joint body there is a continuous layer of field grading material, which even covers the inner deflector.
A layer of insulating rubber outside the field grading material represents the actual joint insulation.
Finally, an outermost layer of the joint body is made up of a semi-conducting rubber screen. The design
must be simple enough to facilitate production and to keep the total production cost at a reasonable
level. It should be noted that this design has lately been successfully upgraded to the 525 kV
[Gustafsson 2015] and 640 kV [Jeroense 2018] HVDC level. However, this technology is still emerging.
The DC cable termination is a combination of a conventional AC termination based on geometric field
control using a stress relief cone and a continuous layer of field grading adapter (FGA), which is placed
on the cable after the cable outer semiconducting screen is removed, connecting the live electrode to
ground. It is illustrated in Figure 5.17 and Figure 5.18 for operating voltage levels of 320 kV and 150
kV respectively. For 320 kV, the cable termination is housed in an oil filled porcelain body (Figure 5.17),
whereas it is flexible and completely oil free for 150 kV.

80
TB 794 - Field grading in electrical insulation systems

Figure 5.16: 320 kV DC cable joint [Sorqvist 2009] in the top and [Ericsson 2003] in the bottom

Figure 5.17: 320 kV DC cable termination [Sorqvist 2009]

81
TB 794 - Field grading in electrical insulation systems

Figure 5.18: 150 kV DC cable termination [Ericsson 2003]

Similar to the 150 kV and 320 kV DC cable termination, the new developed 525 kV and 640 kV HVDC
extruded cable termination shown in Figure 5.19 is also composed of a stress cone and a nonlinear
resistive field grading adapter whereas housed in an SF6 gas insulated composite insulator. The know-
how from the 800 kV HVDC bushing technology is also utilized for a better control of the electric field
around the termination region on the insulator part [Gustafsson 2015] [Jeroense 2018].

Figure 5.19: 525 kV and 640 kV HVDC extruded cable termination using stress cone and field grading adapter
[Gustafsson 2015] [Jeroense 2018]

For basic considerations on dimensioning of HVDC field grading systems refer to 3.4.5. The suitable
composition of the field grading layer used in prefabricated accessories of the commercially available
DC extruded cables are selected from the patented ones. Indeed, extensive research and development
of field grading materials took place in the early 1990’s resulting in a great number of patents (examples:
[ABB 2004] [ABB 2010]). In general, the nonlinear resistive field grading materials are composed of an
EPDM rubber matrix and contain one or two fillers such as: SiC with CB, SiC with semiconducting flake,
ZnO microvaristor, surface treated nano-ZnO [Wang 2006]. The amount of the filler should be above
the percolation threshold and typically requires 30 % to 40 % by volume.
As mentioned above, besides the development of the suitable nonlinear resistive field grading
properties, the optimisation of the whole cable accessory geometry must be also considered to ensure
a proper level of the electric field distribution at the critical locations such as the interfaces between the
different layers of the accessory and the triple points where the different layers meet and also to control

82
TB 794 - Field grading in electrical insulation systems

the electric field under transient voltages (lightning or switching impulses). A double control of the
electric field under DC voltage (nonlinear resistive layer) and under voltage impulses (geometrical) is
necessary. Detailed information about the accessory design optimisation and electric field calculation
examples are given in [Sorqvist 2009] and the patent [ABB 2014], as well as in terms of a general
approach in 3.4.5.
According to another cable accessories manufacturer, existing medium voltage terminations with
nonlinear field grading were adapted and modified in 2004 for terminating 111 kV and 150 kV DC
precipitator cables. To improve the withstand performance related with fast transients and reflections
of impulses, which are caused by frequently occuring steep electrical breakdowns within HV
precipitators, the nonlinear field grading layer for the DC field was combined with an additional refractive
field grading layer. Figure 5.20 shows such a heat shrink termination with nonlinear ZnO microvaristors
embedded into a hot-melt compound.

Figure 5.20: Heat-shrink type HVDC termination with nonlinear refractive field grading (150 kV precipitator cable
termination) [TE]

5.1.3.3 Purely geometrical electric field control using improved DC


rubber materials
According to another HV cable company and in in the frame of their recent development works for the
qualification of the 320 kV HVDC extruded cable system, it has been demonstrated that the electric field
in the accessories (joints and terminations) can be well controlled by the use of improved new rubber
insulation materials properties with suitable DC conductivity and less space charge formation by taking
into account the accessory design optimisation [Paupardin 2015]. An example of the developed HVDC
320 kV cable accessories is shown in Figure 5.21.

Figure 5.21: 320kV HVDC (a) prefabricated cable joint and (b) stress cone cable termination with an EPDM rubber
material [Paupardin 2015]

5.2 Plug-in systems / separable connectors


5.2.1 General
Separable connectors are mainly used for the connection of cables to transformers or GIS. They
introduce high flexibility into the system configuration without the necessity of oil or gas work and give
the possibility to fully test the transformer or switchgear in the factory. They can be mounted on a

83
TB 794 - Field grading in electrical insulation systems

broad range of different cable types. Further, it is possible to easily exchange cable connections with
pluggable bushings or surge arresters, which gives additional flexibility for the network configuration.
Today, outer cone and inner cone systems are available.
The electric field inside the separable connector is typically graded geometrically with deflectors made
from conductive materials such as carbon black filled elastomers. Alternatively, conductive varnish or
coatings are used. In contrast to joints and terminations, where the electric strength of the interface
between the cable insulation and the stress cone is decisive, separable connectors have an additional
interface between the stress cone and the socket or bushing, respectively. Both interfaces have to
withstand the applied voltage stress. For that purpose a sufficiently high contact pressure between the
layers has to be ensured.

5.2.2 Outer and inner cone systems


Outer cone separable connectors are widely used for connections in the medium voltage range.
Compared to inner cone systems these can be cascaded, which provides higher flexibility and is
therefore often seen as an important advantage. The L- and T-shape versions are most commonly used
due to easier handling, but straight connections are available, too. The insulation and, in case of
screened connectors, the outer conductive screen of those plugs are mainly made of EPDM and silicone
rubber. The electrical contact is made either by a bolted or a sliding device. Separable connectors with
sliding contacts are called plug-in type separable connectors. For further design details see [Cigre 2019].
Electric field stress within unscreened plug-in connectors is typically much lower than within screened
connectors. However, they require a wider clearance between the individual phases. Products are
therefore available only up to 24 kV. For higher voltage ranges up to 72.5 kV and compact switchgear
with narrow cable cabinets the use of screened separable connectors is virtually mandatory.

Figure 5.22: Schematic of the outer cone system (left) and inner cone system (right);
acc. to [Cigre 2015]

Figure 5.22 (left) shows a schematic of outer cone systems, Figure 5.23 different types of screened and
unscreened outer cone connectors and a transformer connection realized with the outer cone plug-in
system. For the reduction of the electric field in the cable–stress cone interface, refractive and resistive
field grading are possible besides the geometrical grading.

84
TB 794 - Field grading in electrical insulation systems

Figure 5.23: Screened and unscreened medium voltage outer cone plugs; left top: [Nexans]; left bottom: [TE]
with and without surge arrester; right: transformer connection realized by outer cone plugs

In contrast to the outer cone systems, inner cone systems are available for voltages up to 550 kV (AC).
Due to their metal encapsulation they are completely screened. The necessary contact pressure between
the elastomeric stress cone and the resin counterpart is maintained by a compression spring element.
The electrical contact is provided by a sliding contact. Figure 5.22 (right) shows a schematic, Figure
5.24 depicts an inner cone plug-in connector with socket and a GIS connected with the inner cone plug-
in system.

85
TB 794 - Field grading in electrical insulation systems

Figure 5.24: Inner cone plug system and its application in a GIS [Pfisterer]

5.2.3 Further examples of application


Due to their flexibility plug-in systems have also been utilized for pluggable joints and terminations. An
example is shown in Figure 5.25. Especially for the latter devices integration of microvaristor filled field
grading layers provides means for reducing the electric field strength in air below critical values
[Amerpohl 2002].

Figure 5.25: Left: pluggable joint [Pfisterer]; right: pluggable dry outdoor termination [nkt/TE]

5.3 Transformers
Transformers are playing a major role in the electrical power systems over all voltage levels, from the
low and medium voltage up to the highest voltages. Especially at power plants or in substations they
are crucial equipment, and a reliable function is inevitable. Nevertheless, there are transformer failures
– approximately one third of them is caused by insulation breakdown, mostly due to excessive
electrostatic stress and ageing. So the transformer insulation system is one of the most critical
transformer components. Therefore, this section is dedicated to the transformer insulation and the
grading techniques used in it. As the insulation system of a transformer is one of the most complex
ones and there are a lot of differences in the design methods, varying from one manufacturer to another,
only general information about the design can be given here.
One up-to-date example of the complexity of field grading in transformers shall be given in advance:
Today, new natural ester based fluids are becoming more and more attractive, and occasionally the end
user retrofits existing transformers with a new fluid. However, having a new dielectric would make an
impact on the field distribution, namely, comparing with mineral oils the ester oils have a higher
dielectric permittivity: r = 3.2…3.5 for esters vs. r = 2.2…2.5 for mineral oil. So, when used in
conjunction with new pluggable bushings, ester oils move the higher electrical stress into the epoxy
zone of these bushings while other, more critical areas are relieved. The same effects are expected to
be experienced in the other transformer's zones when mineral oils are substituted with esters. Such

86
TB 794 - Field grading in electrical insulation systems

sometimes unexpected effects on internal field grading have to be regarded when changing the overall
insulation system of a transformer.

5.3.1 General
Fundamentally, the transformer main insulation comprises a large number of elements. Figure 5.26
illustrates a schematic of the insulation structure of a core-type transformer, which consists primarily of
cellulose based paper barriers and insulating oil(s), where this very effective combination has proven to
exhibit outstanding dielectric and thermal properties.

Figure 5.26: Insulation system of a classic high voltage power transformer with core, low voltage winding
LV, high voltage winding HV and compensating winding CW (schematic representation from inside
towards outside) [Küchler 2017]

In general, the oil-barrier insulation system used in power transformers was introduced by the
transformer industry during 1960’s and is still state of the art. This pressboard barrier technology is
based on liquid insulation, which has to ensure basic electric strength by filling and impregnating the
whole transformer’s insulation structure, while the solid insulation is used to form barrier systems
dividing highly stressed oil channels into smaller gaps. The subdivision of long oil gaps effectively
reduces the hazard of long particles and fibre bridges formation in the field stressed zones. It leads into
permitting higher field stresses / less space and hence lower cost, when compared with full oil insulation,
Figure 5.27. Highest electric strength of the oil can be achieved if barriers are arranged rectangular to
the electric field lines. This means that beside the properties of the liquid insulating oil, the performance
of the whole insulating system is determined by the design and optimization of barrier system geometry.

87
TB 794 - Field grading in electrical insulation systems

Figure 5.27: Weidmann Design Curves, i.e. lines of low probability for AC tests discharge inception in oil gaps
under different conditions [Dahinden 1998] [Ziomek 2011]

As to the barrier functions the following conclusions can be drawn from the Weidmann Design Curves
(Figure 5.27):
(a) Free metallic electrodes should be covered with solid insulation.
(b) Electric strength of wide oil gaps can be improved by sub-division into a number of narrow
gaps. Thereby the insulating volume and the cost, respectively, can be reduced significantly.
(c) By subdivision of wide oil gaps it can be expected that the high statistical scatter of discharge
inception for the long gap is reduced to the statistical scatter of the narrow gaps.
(d) Gaps with very non-uniform fields can be subdivided into smaller gaps with different widths.
In common understanding "barriers" are used to stop something that is propagating. Therefore, barriers
could be seen as barriers against further propagation of a discharge streamer in oil. It should be pointed
out, though, that the main purpose of barriers is to avoid ignition of discharges rather than to stop
propagation of discharges.
In case of quasi non-uniform configurations, the function of the barriers is to subdivide the longer oil
gaps in order to increase electric strength. It was shown that minimum volume can be achieved if barrier
volume and barrier thickness are as small as possible. In case of AC application, the barriers do not
contribute to insulation dominantly, see next chapter. This means that oil gaps should be made as
narrow as possible. For DC application also the barrier system itself has to be regarded carefully - refer
to 5.3.3. Table 5.3 gives an overview on principle design differences for AC and DC application, which
will be explained in the next chapters.
Table 5.3 : Electric field intensity in composite dielectrics and under various stresses

Conditions Oil Solid (PB) Oil-PB insul. system


𝑼 𝑼
AC Voltage 𝑬𝒐𝒊𝒍 = 𝜺𝒓
𝟏
𝑬𝒑𝒃 = 𝜺𝒓
𝟐
𝒅𝟏 + 𝒅𝟐 𝒅𝟐 + 𝒅𝟏
𝜺𝒓 𝜺𝒓 𝟏
𝟐

𝑼 𝑼
DC Voltage 𝑬𝒐𝒊𝒍 = 𝝆𝟐 𝑬𝒑𝒃 = 𝝆𝟏
𝒅𝟏 + 𝒅𝟐 𝒅𝟐 + 𝒅𝟏
𝝆𝟏 𝝆𝟐

𝑼 𝟐𝑼 𝑼 𝟐𝑼
Polarity 𝑬𝒐𝒊𝒍 = 𝝆𝟐 − 𝜺𝒓 𝑬𝒑𝒃 = 𝝆𝟏 − 𝜺𝒓 𝟐
𝒅𝟏 + 𝒅𝟐 𝒅𝟏 + 𝟏 𝒅𝟐 𝒅𝟐 + 𝒅𝟏 𝒅𝟐 + 𝒅𝟏
reversal (PR) 𝝆𝟏 𝜺𝒓𝟐 𝝆𝟐 𝜺𝒓𝟏

During the operation of a converter transformer the oil-pressboard combined insulation near the valve
side is exposed to a composite AC-DC voltage. As a result of such superposition of the electric fields,
the electric field strength (E0) in the transformer oil and the electric field strength (Ep) in the pressboard
can be defined as follows [Chen 2018]:
𝜌1 𝜀𝑟2
𝐸0 = 𝑈 + 𝑈 sin(𝜔𝑡)
𝜌1 𝑑1 + 𝜌2 𝑑2 dc 𝜀𝑟2 𝑑1 + 𝜀𝑟1 𝑑2 ac

88
TB 794 - Field grading in electrical insulation systems

Equation 5.2

𝜌2 𝜀𝑟1
𝐸p = 𝑈 + 𝑈 sin(𝜔𝑡)
𝜌1 𝑑1 + 𝜌2 𝑑2 dc 𝜀𝑟2 𝑑1 + 𝜀𝑟1 𝑑2 ac
Equation 5.3

Under the DC as well as the AC-DC voltage, the breakdown strength of the pressboard increases with
the increase in the thickness. However, it decreases under the AC voltage. As shown in Figure 5.28,
the electric field strength of the transformer oil is inversely proportional to the DC content, whereas it is
directly proportional in the pressboard.

Figure 5.28: Electric field strength as a function of the DC content of a combined AC-DC voltage [Chen 2018]

5.3.2 Design criteria for AC and impulse voltages


For homogeneous or cylindrical electric field distributions, the equations given in 2.2.4.1 can be used.
But often the insulating structures are highly non-uniform, and software, based on finite element
method (FEM) or boundary element method (BEM), must be used to calculate the electric field. There
are different possible criteria for the analysis of the calculation results – maximum field criterion and
cumulative field criterion. In [Capuder 2013] it is shown that it is necessary to analyse both methods as
well as to apply both criteria on the results to conclude, which method and criterion combination gives
adequate results.
Due to the ratio of permittivities, e.g. pressboard: εr,PB = 4…5, and oil: εr,Oil = 2.2, the electrical stress
in oil under AC excitation will always be higher than that in the solid insulation, depending on the
thickness of the layers. Therefore, the designer focuses on mastering the oil duct stresses. In addition,
the electric strength of the oil is lower than that of the solid insulation, which makes the oil gaps forming
"weak links" in series to stronger barriers (see Table 5.3).
Actually for designing reliable power transformer insulation, the following parameters have to be taken
into account:
(a) kind of electric stresses (AC, DC, impulse);
(b) stress duration;
(c) stressed volumes;
(d) surface stress ("creep" condition);
(e) non-uniform fields – depending on complexity, 3-dimensional simulations may be required;
(f) oil quality: water, gas and particles content;
(g) manufacturing technology and quality of the transformer factory: geometry tolerance,
cleanliness, climate, drying, etc.
To define electric strength, it is a common practice to use the criterion of 50% breakdown values.
However, this is not the case with the oil gaps due to the high scatter. For proper dimensioning of the
insulation, the transformer industry uses design curves, often named "oil curves" or "reference curves".

89
TB 794 - Field grading in electrical insulation systems

As an example, the industry adopted curves are depicted in Figure 5.27: the "Kappeler curves” and/or
“Weidmann Design curves”, both giving limits of permissible field strength during one minute AC voltage
withstand tests.

5.3.3 Design criteria for DC voltages


As shown in Table 5.3, the design of the insulation system of transformers used for HVDC applications
is quite different to that of transformers for AC applications. But typically the HVDC converter
transformer is subjected to a DC voltage insulation stress as well as AC voltage stress. The AC voltage
stress is predominantly in the insulating oil and defined by the geometry and permittivity of the materials
(refer to previous chapter), whilst the DC stress is governed by the resistivity of the insulating materials
which, in turn, varies with operating conditions. Especially the conductivity of insulating materials is
strongly affected by temperature, field strength, ageing condition etc. As pointed out in [Carlson 1996],
the windings in HVDC transformers are not basically different from those in conventional transformers,
so the conventional insulation materials used in normal power transformers such as paper and
pressboard, together with mineral oil have proven to be suitable for converter transformers, too. The
actual properties of the insulation materials will, however, result in quite different stress distributions
for AC and steady state DC.
The DC potentials will create extra stresses in the transformer in addition to the normal AC stresses. In
a steady state condition, the DC stresses are primarily governed by the conductivities of the individual
materials in the insulation arrangement (for the theoretical background see 2.2). So the following facts
have to be regarded:
- The ratio of conductivities for cellulose and oil may vary in the range from 10 to 1000, depending
on several factors like oil quality, moisture content, temperature, ageing, etc.
- The ratio of relative permittivity for cellulose and oil is much smaller and close to two. This means
that in a steady state condition – with resistive voltage distribution – almost all stress from the DC
potential will be within the solid insulation.
- On the other hand, the AC and as well a transient DC voltage (e.g. polarity reversal) will result in
a capacitive voltage distribution with approximately double the stress across the oil as compared
to solid insulation. DC voltage results in resistive and AC voltage in capacitive distribution.

Figure 5.29: Dielectric stress in a barrier system for AC and DC voltages; ratio of permittivities 2:1 and ratio of
conductivities 10:1, steady state conditions [Dahinden 1998]

The differences in stress patterns must also be considered in the build-up of the insulation systems. In
comparison between oil and cellulose (paper and pressboard), the insulation withstand strength for
cellulose is considerably higher than for oil. In AC insulation systems the amount of solid insulation is
kept low. In DC insulation systems on the other hand there is a considerable increase in solid insulation
in order to avoid critical stress for steady state conditions. The number of barriers and their thickness
are increased. Also the interconnection between the valve side windings and their bushings require
additional paper covering. Figure 5.29 shows the difference between AC and DC field distribution within
an insulating system.
In addition to the stress pattern caused by the steady state DC voltage and the normal AC voltage there
are two transient phenomena for DC voltage to be considered. One is the normal start-up of the

90
TB 794 - Field grading in electrical insulation systems

converter when full DC potentials from the bridges are developed almost instantaneously, the other is
the so-called polarity reversal (PR).
After a sudden change in DC voltage, the field distribution will be capacitive, and it will change into a
resistive distribution over time [Okubo 2015]. The time constant for the transition from a capacitive to
a resistive distribution can be several hours, depending on the insulating system (see Figure 5.30 and
Figure 5.31).

Figure 5.30: Electric field distributions under HVDC polarity reversal conditions from +100 kV to –100 kV,
for parallel PB lap-model, (σoil/σPB = 1000) [Okubo 2015]

Figure 5.31: Dependency of generating stress in oil on conductivity ratio σoil/σPB (polarity reversal from
DC +100 kV to – 100 kV) [Okubo 2015]

In order to reach steady state conditions, the interface between the solid and the liquid insulation must
be charged by conduction processes in the insulating liquid. Electric conduction in oil is based on ion
transportation [Cigre 2016] [Okubo 2012]. These processes are causing problems in HVDC design
because it is very difficult to define material parameters, which are realistic for test and operational
conditions. Field analyses should work out scenarios with highest possible operational stresses, with
lowest possible test stresses and with highest possible test stresses. This is only possible by means of
sophisticated numerical field analysis and advanced HVDC design experience.
The design principles presented in this chapter reflect the current state of the art in transformer
insulation system design. Nevertheless it has to be mentioned that the basic research on the dielectric
behavior under DC stress especially of liquid insulating materials and the interfaces to solid insulation is
still ongoing. The focus of the research activities is on the physical mechanisms contributing to oil
conductivity, i.e. charge carrier generation, transport and recombination as well as the formation of
space charges and the interaction of liquid and solid barrier interfaces (surface charges) [Fu 2015]
[Backhaus 2014] [Gabler 2019]. For analyzing the DC behavior of the materials, methods like the pulsed
electro-acoustic (PEA) method or measurements using the Kerr-effect of dielectric liquids allow the
direct investigation of materials under DC stress [Wu 2016] [Öftering 2019] [Gabler 2019]. The research
leads to a significantly new understanding of the insulation systems, and hence it may be that in future

91
TB 794 - Field grading in electrical insulation systems

new design rules and methods for modelling of the behavior of insulation systems under DC stress will
be introduced.

5.3.4 Special field grading applications in transformers


5.3.4.1 Composite graded insulation
One very effective grading methodology was presented/described in [Mori 1982] and [Okuyama 1984]
where the insulation is used in a way that allows transformer overall downsizing. In the UHV transformer
insulation was arranged as shown in Figure 5.32. It is called a composite graded insulation and is
characterized by a combination of multi-barrier insulation and dielectric shields, which moderates the
electric stress in the oil gap using press board, corner rings and crepe papers near the electrodes. It is
undesirable for insulation characteristics to increase the electric stress by a simple reduction of the
dimensions. This result shows that the new structure offers reductions in insulation dimensions.
As shown in Figure 5.33, in a conventional insulation structure for which gaps are close to the electrode,

Figure 5.32: Conventional insulation versus composite graded insulation; after [Mori 1982] and
[Okuyama 1984]

the electric stress in the oil gaps is very high. In the new insulation structure, the stress is reduced to
that or less of a conventional type by moving the oil gap away from the electrode through use of thicker
insulation for the electrode, especially for the stress ring of end insulation.

Figure 5.33: Electrical stress distribution at end insulation (conventional versus composite
graded insulation); after [Mori 1982] and [Okuyama 1984]

5.3.4.2 Dry-type transformer


As a potential replacement of conventional power transformers the “Dryformer” (Figure 5.34) was
presented in [Anderson 2000]. This new transformer uses a graded insulation system, which is based
on high voltage cross-linked polyethylene (XLPE) cable technology, where these cylindrical conductors
used in “Dryformer’s” windings distribute the electric field evenly, as depicted in Figure 5.34 (right).

92
TB 794 - Field grading in electrical insulation systems

In order to minimize the insulation used, a cable winding may consist of a few lengths of HV cables with
different insulation-thickness. With modern insulation materials and production techniques, 500 kV
cables accommodating field stresses up to 15 kV/mm can be manufactured today, and in the future,
even higher field stresses will be introduced. Conventional rectangular oil/cellulose coil insulation does
not allow such high fields. For “Dryformer” with cable winding, the outer semiconducting layers are
grounded, and the insulation of the cable-windings must function as the main insulation in conventional
transformers as described in [Leijon 1999] [Leijon 2001].

Figure 5.34: Left: “Dryformer’s” windings arrangement; right: “Dryformer’s” and conventional transformers’
insulation systems [Andersson 2000]

5.4 Bushings
High voltage bushings are used to guide a conductor that is on HV potential through a grounded wall,
e.g. the tank of a transformer. Figure 5.35 shows different types of bushings.

Figure 5.35: Ungraded and graded bushings with typical orders of magnitude of partial discharge inception; a)
ungraded bushing, b) and d) geometrically graded bushing, c) capacitively (capacitance) graded bushing [Küchler
2017]

Ungraded bushings can only be used for rather low voltages because of the low inception voltage of
surface discharges upon the insulation, Figure 5.35 a) and 2.4.3. For medium voltages geometrically
graded bushings are typically used, Figure 5.35 b). For voltages above approx. 70 kV geometrically and
capacitance graded bushings are used, Figure 5.35 c) and d). Those concepts are explained in the
following.

93
TB 794 - Field grading in electrical insulation systems

5.4.1 Basics of field grading within bushings


5.4.1.1 Ungraded bushings
If there is no field grading, the ground potential is directly put upon the insulation of the conductor and
ends up with a sharp edge, Figure 5.35 a). Therefore, the inception voltage for surface discharges is
rather low. An elongation of the distance between the ground potential and the end of the insulation
will not increase the inception voltage, so that this type of bushing can only be used for rather low
voltages and shall not be further considered here.

5.4.1.2 Geometrically graded bushings


The ungraded bushing can be improved by geometrical field grading. This means that the edge of the
ground electrode is rounded in order to minimize the electric field strength there, Figure 5.35 b) and
d). For dimensioning of the electrode geometry, usually simulation tools are used. The goal is to keep
the electric field stress much smaller than the inception voltages for partial discharges of the insulating
materials used. In gas insulated systems (Figure 5.35 d) geometrical grading can be used. Special care
must be taken if solid insulation is used and so-called triple points are created, like in a cable termination,
Figure 5.36. In such a case, the triple point must be filled with an insulating material (e.g. oil, SF 6, or
silicone) that has sufficient electric strength. The geometrical field grading will lead to a short axial field
grading length, i.e. the electric field will be focused close to the rounded electrode, and a longer distance
between the electrodes has no influence on the maximum field at the round electrode.

Figure 5.36: Triple point between the contour of the electrode and the surface of the insulating material [Küchler 2017]

5.4.1.3 Capacitance graded bushings (Condenser bushings)


As the ungraded and the geometrically graded bushings have several disadvantages, especially in the
case of solid insulation material, the most common method for field grading in bushings is the use of
capacitive grading foils inside the solid dielectric. These can be aluminium foils that are put inside the
dielectric while it is wound of Kraft 7 or crepe paper. After the winding process the dielectric is dried and
impregnated. Different types of capacitance graded bushings can be found:
 The RBP (resin bonded paper) bushing is not produced any more. The insulating material was
pre-impregnated paper that was wound upon the conductor of the bushing and cured after
fabrication by a heat cycle.
 The OIP (oil impregnated paper) bushing consists of a dielectric that is made of Kraft paper, and
after drying it is impregnated with dry insulating oil.
 The RIP (resin impregnated paper) bushing dielectric is made of crepe paper, and after drying it
is impregnated with resin that is hardened afterwards with a special heat cycle.
 The RIS (resin impregnated synthetics) bushing has a dielectric made of synthetic fabrics,
impregnated with resin.

7
Kraft paper is paper or paperboard (cardboard) produced from chemical pulp in the “kraft process” [Wikipedia].
It is a smooth paper in contrast to crepe paper.

94
TB 794 - Field grading in electrical insulation systems

Details of the structure of an RIP and an RIS design are shown in Figure 5.37, while in Figure 5.38 a
comparison of water absorption during storage in water is shown, which demonstrates one benefit of
the RIS design.

Figure 5.37: Active component of RIP (left) and RIS (right) design [HSP 2015]

Figure 5.39 shows a complete cut of a 420 kV RIP type bushing with some 50 grading foils. Usually the

Figure 5.38: Water absorption (storage in water) of RIP and RIS design [HSP 2015]

innermost grading foil is connected to the high voltage conductor, and the outermost grading foil is
connected to ground. This ground connection is made outside of the dielectric at the so-called
measurement tap, because the ground connection can be removed when the bushing is taken out of
operation and dielectric measurements for condition assessment shall be performed.

Figure 5.39: Cut through a 420 kV RIP type bushing with the implemented aluminium grading foils (Photo by M. H. Zink)

The grading foils are on floating potential, and they take different potentials during operation. At AC
voltage, the capacitances between the foils determine the field grading. At DC voltage, the resistances
between the foils are responsible for field grading. In both cases the field grading is the same as long
as for DC application the temperature inside the dielectric is kept homogeneous. As the circumference
of the foils increases from the inside to the outside, their length must decrease in order to achieve

95
TB 794 - Field grading in electrical insulation systems

almost constant capacitances/resistances between neighbouring foils. Hence the outermost foil is the
shortest and the innermost is the longest one. This way a field grading area is generated at both ends
of the bushing, see Figure 5.40 and 2.2.2.2.

Figure 5.40: Schematic view of an HV bushing with grading foils; acc. to [Zink 2013]

As the grading foils take different potentials during operation, the equipotential lines are guided by the
geometry and length of the grading foils. So the grading foils have two major tasks: On the one hand
they grade the electric field in radial direction in order to achieve almost constant radial electric field
strength inside the dielectric. On the other hand they guide the electric field also in axial direction by
defining the equipotential points at the surface of the dielectric and the bushing’s outer envelope. Details
about the design of the grading foils can be found in literature, e.g. [Küchler 2017].
The diameter of the dielectric is defined by the electric strength of the insulating material. The length
of the grading area, and hence the field strength at the surface of the dielectric or the envelope has to
be chosen considering the surrounding media. If one side of the bushing is used under open air (like
the left side of Figure 5.40) the outer envelope is made from a porcelain or a composite (silicone/fibre
glass reinforced plastic) insulator, and the grading area must be of sufficient length in order to achieve
sufficiently low electrical surface stress. If the outer insulating material of the bushing is transformer oil
or SF6, the grading area can be much shorter because the insulating material is electrically stronger,
e.g. right side of Figure 5.40.
As the dielectric of a bushing has to insulate high voltages, it must be of sufficient radial thickness. This
causes also a high thermal insulation and prevents the heat generated inside the conductor of the
bushing to be transmitted to the outside. Hence the dielectric must withstand also high operating
temperatures of more than 100 °C. The temperature is the highest at the inner radius of the dielectric,
and a temperature gradient arises from the inside to the outside.
While for AC operation this temperature gradient does not influence the electric field, because
permittivity is not temperature dependent, for DC operation the temperature gradient causes a field
migration inside the dielectric. This means in the warmer regions with the higher electrical conductivity
field strength is reduced and increased at the colder outer regions with lower conductivity [Wirth 2015].

5.4.1.4 New approaches


Recent developments of compact high-voltage bushings make use of layers of nonlinear field grading
material, for instance microvaristors, directly applied to the outer surface of a hollow core insulator as
an interface between the fibre glass reinforced plastic tube and the weather sheds. This allows for very
compact designs with simplified internal geometric grading measures. An implemented example of such
a “slim bushing” is given in 5.6.2.

96
TB 794 - Field grading in electrical insulation systems

5.5 Rotating electrical machines


5.5.1 General
In rotating electrical machines field grading is applied to suppress local field enhancement at the stator
slot end. This field grading system is often called “end corona protection” (ECP), “stress grading” or
“stress suppression”. Figure 5.42 (left) highlights the location of the ECP, on the right hand side an
illustration of the principal electrical insulation in this electrical apparatus is given.

Figure 5.42: Left: picture of a generator end-winding (EGS = ECP = end corona protection) [Staubach 2010]; right:
schematic view of the field grading area in a rotating electrical machine (ECP = end corona protection, OCP = outer
corona protection) [Klamt 2006]

In the slot portion, i.e. the laminated core, the installed stator bars are covered with a semi-conductive

coating to prevent PD activity between the main insulation surface and the grounded laminated core.
Figure 5.41: Resulting electrical potential grading by means of ECP along the bar
surface

This outer corona protection (OCP) ends outside of the slot. To suppress PD activity at the end of the
OCP field grading is mandatory for machines beginning with a rated voltage of approximately 6 kV
[Kaufhold 2008]. In case of an appropriate electrical characteristic of the applied ECP, a smooth
electrical potential distribution along the bar surface is obtained, see Figure 5.41.
There are different techniques for field grading systems used as ECP, e.g. capacitive, linear resistive,
nonlinear resistive (single or multiple layers) or double-shielded configurations [Kimura 1985] [Staubach
2012b]. Figure 5.43 gives an overview of different kind of resistive ECP-configuration setups.
Nowadays a defined number of painted or wrapped layers with nonlinear resistance characteristic
realized by SiC are commonly used on rotating electrical machines [Weidner 2011] [Klamt 2006] (see
also 3.3.3.1). However, due to some special limitations and drawbacks linked to SiC approaches other
nonlinear filler concepts have been developed and applied [Donzel 2016] [Lang 2013] [Staubach 2013]
(see also 3.3.3.2, 3.3.4.4).

97
TB 794 - Field grading in electrical insulation systems

Figure 5.43: Overview of different setups for field grading in a generator [Staubach 2012b]; references in the figure
refer to the reference list of [Staubach 2012b]

Design features, i.e. amount of overlapping with OCP, number and length of the layers, painted or
wrapped version and resistance characteristic, are strongly dependent on the actual stator winding
design. In general, the ECP is dimensioned for high voltage testing conducted during manufacturing of
the machine, where the field grading system is exposed to its highest electric field stress and short-time
thermal stress [Staubach 2012b]. Besides this the ECP must also work efficiently during operation, when
the electrical stress is comparatively low. On the other hand the stator winding warms up due to the
ohmic losses from the copper current. Therefore, the ECP is exposed to elevated temperatures, which
may result in accelerated thermal aging or thermal runaway effects, see also 5.5.2.
Non-accurate tailoring of the ECP may lead to excessive heating or flashover during high voltage testing
or to PD deterioration in service. The change of the ECP-resistivity during manufacturing and aging in
operation has also to be considered [Brütsch 2006]. Finally, the scattering of the ECP-characteristic
must be taken into account. Because of the nature of the filler material the resulting surface resistance
of a tape can vary over several decades [Brandes 2008].

Figure 5.44: Left: Tracking failure caused by thermal overstressing [Baumann 2011]; right: indications of PD
deterioration at the ECP-to-OCP overlapping [Stone 2009]

The typical pattern of PD deterioration at the ECP-to-OCP overlapping caused by wrong electrical
characteristic is shown in the left picture of Figure 5.44. The right picture of Figure 5.44 shows a typical
indication of thermal overstressing during lab testing, which may be caused by the thermal runaway
effect.

5.5.2 Direct line connection


In generators of fossil power plants the stator winding insulation is commonly exposed to test voltages
of up to 1.3·(2·U + 1 kV) for 1 min, where U is the generator’s rated voltage (phase-to-phase). This
means the ECP must withstand a voltage level, which is more than 4.5 times higher than occurring at
the most in operation [Malamud 2000] [Staubach 2012b]. For hydro generators this value can be even
higher. Some customers request individual bar testing up to a voltage level of 4·U. This equals almost
seven times the line to ground voltage in operation, which may be the physical limit for proper function
of ECP even if applied for a short period of time only [Lamas 2009].
Besides the occurrence of high electric field stress, the ECP may be stressed thermally. Because of its
nonlinear field dependent characteristic the ECP becomes more conductive when exposed to field
enhancement, and Joule losses would increase. As a result the field grading material will heat up, and
as a consequence the ECP material will become more conductive. If the design of ECP is not carefully
dimensioned the mechanism of heating up und becoming more conductive will become instable and
may lead to a thermal runaway, also-called thermal runaway effect [Baumann 2011]. As a consequence
the ECP will be destroyed thermally (burning) or electrically (flashover).

98
TB 794 - Field grading in electrical insulation systems

Figure 5.45: Typical temperature distribution (taken by an IR camera) in the ECP-region on a


test bar [Brütsch 2006]

Figure 5.45 shows a typical temperature distribution in the ECP-region during testing. Directly at the
end of the OCP, in the figure named “External corona protection”, in the overlapping area with the ECP
a hot-spot is obtained. This is caused by the highest current density mainly due to the capacitive current
[Staubach 2012b].
In normal operation electrical and thermal stresses in the ECP-region are comparatively low. The
temperature increase caused by Joule heating is even negligible. From the thermal point of view only
winding heating itself caused by current flow through the copper bundle has to be considered. This may
also heat up the ECP and change its resistance. Most important in operation is that no PD are present
in any situation, and that thermal deterioration is minimized. In addition, the effect of stray capacitances
towards adjacent winding elements or the grounded laminated core on the ECP has to be taken into
account.
While some qualification or verification work on an ECP can be performed by means of test samples or
bars in the lab, also FEM-models (see 4.3) are commonly used to simulate operating conditions, which
are difficult to realize by lab testing alone [Staubach 2017]. The left picture in Figure 5.46 shows a test
setup for the direct measurement of the electric field strength distribution along the ECP surface with
the help of a Pockels sensor (see 4.4), whereas the plot on the right hand side presents results of an
FEM simulation. On this basis further assessment can be made.
Depending on manufacturing process or bar design tapes or varnishes are nowadays used for field
grading. While each layer of ECP varnish can be electrically adjusted by using different type or amount
of filler particles, tapes have a defined electrical resistance characteristic, pre-defined by the tape
manufacturer [Roberts 1995]. This may be a disadvantage because each design might need an
individual ECP-characteristic. By painting different lengths of ECP-layers or by wrapping ECP-tape
stepwise, the efficiency of field grading may be increased in addition [Staubach 2012b].

Figure 5.46: left: Direct measurement of the electrical field strength distribution along the ECP surface by a Pockels sensor;
right: Results of an FEM simulation to assess ECP efficiency [Staubach 2017]

99
TB 794 - Field grading in electrical insulation systems

5.5.3 Converter fed drives


In general the above mentioned considerations are also valid for converter fed drives. However, grading
materials that perform satisfyingly under power-frequency (50/60 Hz) may experience a premature
failure when exposed to pulse width modulated (PWM) voltage source converters (VSC) [Sharifi 2010].
Figure 5.48 shows typical resulting voltage signals. The voltage at the motor terminals shows overshoots
(spikes) (Figure 5.48, “B”) due to the very short rise time of the impulses at the converter output and
resulting traveling wave effects on the connection leads. The voltage stress shown here is a kind of
worst-case stress, as modern multi-level converters generate much lower voltage steps as can be seen
in Figure 5.47 [IEC 2017].

Figure 5.48: Left: Trapezoidal test voltage (yellow) and current through the ECP layer (magenta) [Baumann
2011]; right: Resulting voltage signal with sharp spikes at PWM
(A: converter output voltage, B: voltage appearing at the motor terminal) [Wheeler 2005]

Figure 5.47: Waveform representing one complete cycle of the phase to phase voltage at the terminals of a machine
fed from a 3-level converter (left) or an cascaded 10-level converter (right) [IEC 2017]

Grading materials designed for 50/60 Hz applications may not be capable of controlling the stress caused
by high dU/dt on the surface of the winding ends [Espino 2005]. The field grading material will behave
differently when exposed to voltage signals containing a wide band of frequencies. In addition, the
frequency dependent capacitive current caused by the main insulation will affect the field grading
significantly [Staubach 2011]. Figure 5.49 gives an overview on the change in the electrical potential
distribution along the ECP surface dependent on different frequencies. Increasing the frequency will
result in a steeper voltage increase and, subsequently, higher electric field stress.
Thermal hot spots or PD may occur in the grading materials when they are operated with PWM-VSC.
The plots in Figure 5.50 illustrate the influence of the switching frequencies of PWM voltage source
converters. The left picture shows a temperature distribution for 5 kHz, which is comparable to a
generator, where the hot spot occurs at the OCP-to-ECP overlapping. For a higher frequency of 50 kHz
a second hot spot develops directly at the slot exit in the OCP region. This is an indication for field
enhancement in this area and could also be the origin for PD activity. If these effects are not taken into
account, the PD and hot spots potentially result in the failure of these materials, and with time, the
complete failure of the insulation system.

100
TB 794 - Field grading in electrical insulation systems

Figure 5.49: Peak voltage along the stress-graded end-winding calculated for different frequencies (left: medium
conductivity tape, right: high conductivity tape) [Wheeler 2005]

5.6 Insulators

Figure 5.50: Temperature distributions outside of the simulated slot for switching frequencies of 5 kHz (left) and
50 kHz (right) [Sharifi 2010]

5.6.1 Line insulators


High-voltage line insulators may require field grading measures mainly for two reasons:
- improved performance under severely polluted conditions (prevention of dry band formation and
arcing);
- suppression of water droplet corona on the surface of non-ceramic insulators.
The first item is important for non-ceramic (polymeric) insulators only, because electrical surface
discharges may cause severe deterioration of such insulators. To avoid water droplet corona the basic
electric field stress on the insulator surface must be reduced to values, which prevent corona at the
triple zones “water droplet – insulator surface – surrounding air”. These triple zones are the origin of
partly severe electric field enhancement. Factors of up to eight compared to the basic tangential field
stress have been reported in the literature, e.g. [Feier-Iova 2009]. Besides strong radio interferences
surface corona may cause ageing and deterioration of the surface (loss of hydrophobicity, tracking and
erosion) [Gubanski 2005] and in severe cases even brittle fracture of the insulator shaft near the end
fittings [Kuhl 2001].

101
TB 794 - Field grading in electrical insulation systems

In general, line insulators suffer from an uneven axial electric potential distribution, caused by the
affecting stray capacitances to ground. This effect is the more pronounced the longer the insulator is,
because its self-capacitance decreases with length, whilst the stray capacitances would increase. Figure
5.51 explains the situation with the help of a distributed parameter equivalent circuit.

Figure 5.51: Electrical equivalent circuit of a line insulator (left) and resulting axial voltage distribution (right) under
variation of the electrical parameters: dC ... self-capacitances of the insulator, dCE ... stray capacitances to ground, dCL
... stray capacitances to the high-voltage terminal, Ua = Ux/Utotal ... partial voltage between two nodes referred to total
voltage, a = x/l … partial distance referred to total length [Beyer 1986]

It is obvious that the line sided end of the insulator can be exposed to extremely high voltage and thus
electric field values if no countermeasures are taken. But it is just here – at the junction between
insulator shaft and end fittings – where high electric fields are most dangerous (see Figure 5.52).
Electrical discharges, in combination with moisture, cause formation of acids, which may in consequence
lead to brittle fracture (Figure 5.53).

Figure 5.52: Corona activity at line sided insulator ends [IEEE 2008]

Figure 5.53: Examples of brittle fracture on composite line insulators [INMR]

Typical countermeasures are corona rings, which electrically shield the critical areas of the insulator, or,
in other words, increase the stray capacitances dC L in Figure 5.51. These corona rings additionally act
as arcing horns to protect the insulator from the power arc in case of a flashover. An implementation
example is shown in Figure 5.54.

102
TB 794 - Field grading in electrical insulation systems

Figure 5.54: Non ceramic line insulator with corona ring [IEEE 2008]

Theoretically, the overall axial potential distribution of an insulator can be improved by resistive grading
in form of a conductive surface layer. It has turned out, however, that notable effects can be achieved
only if the resulting surface leakage current is very high, in the range of several milli-amperes, which
results in high power losses and potential thermal stability problems. Nevertheless, resistive grading
has occasionally been applied, especially in order to improve insulator performance under severe
pollution conditions. Basically, this feature is covered by appropriate choice of creepage distance, shed
profile and material (in case of non-ceramic insulators). But a certain limited conductivity is assumed to
support these measures. Different approaches are necessary for ceramic and for non-ceramic insulators.
For ceramic insulators the historical development of weakly conductive glazing is addressed in 3.2.1.
The intended and expected benefit has been a possibly improved operating performance under severely
polluted environmental conditions due to following mechanisms being effective:
- The leakage current that permanently flows across the conductive surface causes slight increases
(in the range of 10 K) in surface temperature. Pollution layers thus stay dry even under the
impact of fog and dew.
- There exist always current paths in parallel to dry band zones, which helps avoiding "dry band
arcing".
- In general, a conductive surface layer will help unifying the uneven axial potential distribution
along the insulator and will thus reduce excess field stress at the typically higher stressed
locations of the insulator (in case of line insulators e.g. the conductor-sided end).
There is no general agreement about the benefits of resistive glazing. Effectiveness could be
demonstrated in short term laboratory tests, even though the increase in lightning impulse voltage
withstand was only moderate [Elsässer 2001], as well as in long term tests [Lambeth 1983] [Matsui
1997]. But application of conductive glazing is not without problems. Its production process is difficult
to control, and the resulting properties strongly depend e.g. on the firing process. Repeated firing
improves the yield, though for the price of extraordinarily high production costs. The manufacturers,
therefore, prefer specifying tolerances in conductivity of up to  50 % instead. Aging of the glazing
(resistance increasing with time) was observed [Egiziano 1987] [Gubanski 2005], and corrosion
problems have been reported [Ullrich 2001] [Ullrich 2002]. There is also a high risk of too extensive
heating, because the additional surface current flow should be in the order of one magnitude above the
current through the pollution layer. A basic disadvantage in this context is the positive temperature
coefficient of a semiconductor's conductivity, making the layer potentially thermally unstable, which

Figure 5.55: Non-ceramic line insulator with microvaristor filled silicone rubber
layer; acc. to [Debus 2011b]

103
TB 794 - Field grading in electrical insulation systems

might result in a "thermal runaway". In general, resistive glazing has not yet found extensive
applications.

Figure 5.56: Model of dry zone arcing suppression by a capacitively coupled microvaristor layer; left: dry zone
flashovers due to high field stress underneath the sheds of a conventional insulator; middle: flashover suppression by
a microvaristor layer directly in contact with the insulator surface; right: flashover suppression by a microvaristor
layer covered by a conventional layer with no direct contact to the environment; acc. to [Hinrichsen 2011]

Also for non-ceramic line insulators a similar approach has been investigated and implemented [Debus
2010a] [Debus 2011a] [Debus 2011b] [Debus 2015] [Lapp 2014]. Controlled, nonlinear electric field
dependent conductivity has been achieved by a layer of silicone rubber filled with microvaristors. It has
turned out, however, that the filled silicone rubber does not easily fulfill the requirements on tracking
and erosion resistance necessary for outdoor applications. For this reason a double extrusion process
has been introduced, in which the insulator core is first covered by the functionally filled material and
then by a conventional silicone rubber sheath that protects the layer beneath from environmental impact
(Figure 5.55).
The effects observed in laboratory tests – reduced dry band arcing activities under rain conditions – are
not caused by overall potential distribution improvements (for this purpose the conductivity would have
had to be chosen much higher) but by local, capacitively coupled conductive bypass effects, which
prevent flashovers of dry bands, as is principally shown in Figure 5.56. This may explain why a
microvaristor filled layer can be effective just by capacitive coupling, without any galvanic connection
to electrodes.
A critical parameter for achieving the intended effects is the correct choice of the “switching” field
strength of the applied microvaristors, i.e. the field strength at which current density starts to notably
increase (see 3.3.3.2). Typically, this field strength is in the range of 1000 V/mm for cable accessory
applications. This value has turned out to be too high for insulator applications. Switching field strengths
down to 500 V/mm were investigated on the above line insulators, and possibly even lower values might
be necessary in this special application. Further research on this new technology will, therefore, be
necessary. Different nonlinear electric characteristics of the microvaristor filled silicone rubber systems
have to be developed and investigated, and long term tests under natural environmental conditions
have to be performed [Seifert 2008] [Seifert 2011] [Schulte 2013].

5.6.2 Hollow core insulators


Microvaristors have also been applied to hollow core insulators for use in high-voltage bushings. This
approach was proposed and theoretically and experimentally validated for the first time on a hollow
core insulator for a 145 kV cable termination [Boettcher 2001]. Later on [Donzel 2007a] [Donzel 2011]
performed further validations by field simulations. Benefit is a possible remarkable reduction in size
(diameter) of the hollow core insulators. The first known implementation for an extra-high-voltage
system dates from 2014 [Ye 2014]. A microvaristor filled epoxy resin layer was applied directly to the
core of the hollow core insulator of a 550 kV ungraded bushing, limited to the region around the internal
shield, and was afterwards covered by the silicone rubber sheath (Figure 5.57). By this measure the

104
TB 794 - Field grading in electrical insulation systems

internal insulator diameter could be reduced by 30 %. This compact (“slim”) design was theoretically
investigated and a prototype successfully tested in the high-voltage laboratory.

Figure 5.57: Conventional geometrically graded bushing (left) and a microvaristor controlled, non-graded “slim”
bushing (middle); details of microvaristor application (right) [Ye 2014]

Figure 5.58 shows a direct comparison between a compact and a conventional bushing.

Figure 5.58: Comparison of a compact and a conventional bushing of the same


technical ratings [Ye 2015]

5.6.3 GIS insulators


5.6.3.1 Conventional GIS insulators
Different insulator geometries are applied in GIS (gas insulated switchgear), depending on different
particular electrical and mechanical requirements. Figure 5.59 shows some basic examples.

Figure 5.59: Principle sketches of GIS insulator designs; a) 1-phase disc-shaped closed insulator; b) 3-phase disc-
shaped closed insulator; c) 1-phase cone-shaped closed insulator; d) 1-phase cone-shaped open insulator [Tenzer
2015]

105
TB 794 - Field grading in electrical insulation systems

Up to 170 kV system voltage three-phase encapsulation is quite common, at least for the bus bars, and,
therefore, three conductors must be led through the insulator. For the system voltages above 170 kV
typically one-phase encapsulation is used, and the insulator becomes part of a nearly ideal coaxially
cylindrical system, in which electric field strength would decrease inversely proportional to the radius
from its maximum value at the inner conductor to its minimum at the outer conductor.
In case of adequately chosen permittivity and geometry a disc-shaped insulator may have a nearly
constant tangential electric field stress along its surface. However, the creepage distance might be too
short. This can be avoided by choosing a cone shape, where again by a carefully chosen permittivity
and geometry nearly constant and at the same time very low tangential electric field stress can be
achieved. Both cases are shown in Figure 5.61.

Figure 5.61: Equipotential lines of a disc-shaped (left) and a cone-shaped (right) GIS insulator [Tenzer 2015; acc. to
[Mosch 1979]]

In GIL (gas insulated lines) additionally basically cylindrical support insulators are applied, which can

Figure 5.60: Cylindrical support insulators of a GIL [Tenzer 2015; acc. to [Koch 2012]]

slide along the surface of the inner conductor (Figure 5.60). Their axial electric field distribution is
basically the same as the radial field distribution in a disc- or cone-shaped insulator.

5.6.3.2 Functionally graded approaches


By application of -field grading material with a graded permittivity distribution, the electric field
distribution around the solid insulator in power apparatus can be improved [Kurimoto 2002] [Kurimoto
2010] [Ju 2009]. According to the critical points of electrical insulation in GIS, the merits of the application
of -FGM to GIS spacer can be listed as follows:
- relaxation of concentrated electric field stress, like at the triple junction;
- relaxation of electric field stress at the embedded electrode in the solid insulator;
- relaxation of electric field stress at the interface gas – solid insulator;
- simplification of spacer and electrode configurations in gas/solid composite insulation system.

106
TB 794 - Field grading in electrical insulation systems

Supposing the application of -field grading material to GIS spacers, the fabrication and simulation
techniques for the post-type GLP-FGM in 2.2.4.2 have been extended to the coaxial disk-type GLP-
FGM. Figure 5.62 [Ishiguro 2014] [Hayakawa 2014] shows the appearance and permittivity distribution
of coaxial disk-type GLP-FGM with TiO2 and SiO2. The fabricated GLP-FGM exhibited a relative
permittivity (r) distribution from r = 8.2 at the HV side to r = 5.3 at the GND side. The fabrication result
of permittivity distribution agreed well with the simulation result.

60

20

Central
conductor

Figure 5.62: Appearance and permittivity distribution of a coaxial disk-type GLP-FGM [Ishiguro 2014] [Hayakawa
2014]

Figure 5.63 [Hayakawa 2014] shows the electric field distribution on a GIS spacer surface with the
coaxial disk-type GLP-FGM of Figure 5.62 and a conventional spacer with a uniform r = 6. The
maximum electric field strength at the HV side (surface length x = 0 mm) was reduced by 12 % for the
fabricated coaxial disk-type GLP-FGM spacer, compared with that for the conventional spacer. An even
higher value of maximum strength reduction by 27 % has been obtained by simulation with GLP-FGM
with SrTiO3 (r = 332) and SiO2 [Ishiguro 2014].

Figure 5.63: Electric field distribution on GIS spacer surface with coaxial disk-type GLP-FGM [Hayakawa 2014]

-field grading material can be fabricated by not only the centrifugal force application, but also e.g. by
3D printing [Liu 2016] or by flexible mixture casting for GIS spacers and other power apparatus.
Japanese GIS manufacturers have investigated the feasibility of -field grading material to actual GIS
spacers within a cooperative research project ‘Leading Research for the Development of Innovative
Functional Insulating Materials’, which has been undertaken by four universities and three companies
in Japan under the funding of NEDO (“New Energy and Industrial Technology Development
Organization”) during 2015 to 2016 [Ohki 2015]. A size reduction of GIS spacers by (20…30) % has
been expected by -field grading materials from the viewpoint of electric field analyses, and the target
specifications of -field grading materials for GIS miniaturization have been discussed.
Electric field grading and control can also be expected by adding a moderately conductive thin layer on
the GIS spacer surface. The conductivity can be graded locally on the spacer surface, which can be
referred to as -field grading material [Hayakawa 2014]. The electric field grading techniques using -

107
TB 794 - Field grading in electrical insulation systems

and -field grading materials can be expected to contribute to compactness, simplification, lower losses,
lower cost, etc. of power apparatus, compared with conventional power apparatus.

5.6.3.3 HVDC gas insulated systems (GIS)


HVDC gas insulated systems (GIS) have two principal advantages: installations can be made far more
compact than the traditional air insulated systems (AIS), and they have a significantly lower sensitivity
to ambient factors. The most obvious cost-saving potential can be found on offshore converter platforms
where the air clearance required for AIS would lead to a large and heavy offshore structure. By using
DC GIS, where the live parts are encapsulated in a pressurized and grounded enclosure, the need for
air clearance around the components is completely eliminated, and the volumetric space of the
switchgear component can be drastically reduced, typically by 70 to 90 %.
However, a direct application of existing AC GIS for DC use is typically not possible due to the space
charges build-up phenomenon on the insulating spacers’ surfaces [Winter 2011] [Winter 2014]. These
accumulated charges locally enhance the electric field, especially at DC polarity reversal leading to a
premature failure of the system [Cigre 2012]. To avoid this problem advanced functionally filled epoxy
resin with adjusted electrical conductivity can be used.
Nonlinear conductive filled epoxy resin with standard commercial ZnO microvaristors, where the
switching field strength Eb 8 of the composite is about 1.2 kV/mm, can be used in GIS spacer insulators
under moderate applied DC electric field stress. However, for compact GIS where the operating applied
electric field is very high, Eb must reach higher values, and hence further development of the
microvaristor filler type is necessary. Trials on using a special mixture of small spherical (d50  20 m)
and large amount of small fragments (d50  3 m) of ground particles of ZnO microvaristor in epoxy
resin have shown that a high switching field strength of Eb  3.5 kV/mm can be reached (see Figure
5.64), however, the DC conductivity of the composite was reported to be unstable over time [Tenzer
2013a]. Further improvements of the manufacturing process are in progress to solve the dc instability
problem.

Figure 5.64: Characteristic DC E-J-characteristic of microvaristor filled epoxy resin insulators with “standard”
microvaristors and of the microvaristor filled epoxy resin filled with “Type A”, which is a mixture of spherical and
ground particles microvaristors [Tenzer 2013a]

A new approach with another type of filler, namely MFF (“MFF” = Minatec® functional fillers; since 2015
named lriotec 7000®) [Rüger 2012] [Greb 2015]), has been recently investigated and has shown very
promising results [Tenzer 2013b] [Winter 2014], refer also to 2.2.4.2. The MFF filler provides, as in the
case of ZnO microvaristors filler, a defined intrinsic nonlinear conductivity that can be adjusted to fit the
intended application requirement. A resulting field distribution in a HVDC GIS spacer is shown in Figure
2.19. Simulations have shown that less charges tend to accumulate in the volume and along the surface,
when using a material made out of MFF filler. Also the maximum field strengths in steady state
conditions can be reduced.

8 Eb is defined here as the field strength, at which current density JDC = 1 µA/cm² (see also 3.4.5.1.1).

108
TB 794 - Field grading in electrical insulation systems

6. Summary and conclusions


Field grading in electrical insulation systems has evolved over the years from the application of simple
geometries and homogeneous dielectrics with linear properties to complex systems that make use of
designed composites made of selected matrix materials and fillers with nonlinear properties of the
permittivity or the conductivity and deliberately modified distribution of the filler particles to further
optimize the utilization of the material. The latter is currently in the introductory phase, and promising
results have already been achieved. The trend towards HVDC electric power transmission represents a
technical challenge since DC stress requires in almost all cases to include the temperature in the
technical evaluation of the field grading systems, which is dominated by the conductivities of the
involved materials. This was not the case to that extent in the past due to the fact that permittivity, the
dominant parameter for field grading in AC applications, is hardly affected by temperature whereas
conductivity is. Additionally, transient phenomena are much more difficult to handle in DC systems
because the time constants of the transitions between resistive and capacitive electric field distributions
may take extremely high values of up to several weeks or even months.
Although it would have been most likely beneficial for some end users of field grading materials to base
their requirements on standards, no attempt was undertaken so far to standardize neither field grading
materials nor their components. A reason may be that the general requirements are too diversified to
establish a single standard being useful to a reasonable number of users.
The application of highly nonlinear resistive field grading systems in MVAC and HVAC apparatus is limited
to some extent by the fact that relevant standards usually require to perform AC withstand tests at
comparably high levels in order to provide safety margins based on historic data. With the integration
of nonlinear field grading systems into modern designs this may cause overloading and overheating or
even local destruction of the field grading during testing. Future new standards or new editions of
existing standards may have to take this into account.
The purpose of this Technical Brochure is to introduce the theoretical concepts as well as the various
practical implementations of electric field grading, which is a basic and essential measure of field control
in many apparatuses of the electric power transmission and distribution systems, such as cables and
their accessories, plug-in-systems (connectors), bushings, rotating electrical machines, transformers or
gas insulated systems.
Common uncertainties and misunderstandings of the terms geometric, capacitive, refractive, resistive,
bulk, interface or surface, linear and nonlinear electric field grading are smoothed out in Chapter 2, in
which a detailed theoretical overview of the different approaches is given. As a novel and still emerging
technology, field grading by functionally graded materials with non-homogeneous, spatially distributed
electrical properties (permittivity, conductivity) is also discussed, even though apparatus that would
make use of them are not yet commercially available. Anyway, deployable solutions have been published
how to achieve functional grading by centrifugalizing filler powders of different mass densities and
permittivities in the polymer matrix during the curing process.
Unitary materials with suitable electrical properties, e.g. very high permittivity or well controlled
conductivity, could be directly applied as field grading materials. However, these are typically ceramics,
which have mostly unfavorable mechanical properties and, therefore, cannot directly be used as bulk
insulating material. Field grading materials are thus compounds in most cases, and detailed knowledge
about the different electrical, mechanical, chemical properties of matrix materials and fillers is
indispensable. Most of the matrix materials are polymers, which are presented and discussed in detail
in Section 3.2. In the majority of cases, the following types of polymers are applied:
 thermoplastic polymers – e.g. polyethylene (PE), polyvinyl chloride (PVC), polypropylene (PP), or
ethylene-vinyl acetate (EVA);
 thermoset polymers – e.g. epoxy resins (EP), polyurethane (PUR), or cross-linked polyethylene
(XLPE);
 elastomers – e.g. silicone rubber (VMQ) (subdivided into high temperature vulcanizing (HTV), room
temperature vulcanizing (RTV) and liquid (LSR) silicone rubbers), ethylene propylene rubber (EPM),
or ethylene propylene diene rubber (EPDM).

109
TB 794 - Field grading in electrical insulation systems

Similar attention has to be paid to the fillers, which are used to tune either the permittivity or the
conductivity of a material. They are discussed in Section 3.3. The most widely used fillers are carbon
blacks, mainly in the form of powders. Percolation theory must be well understood in order to achieve
the desired electrical properties, which are very different for filler concentrations either below or above
the percolation threshold. It must also be considered that carbon particles tend to agglomerate in the
manufacturing process, thus causing inhomogeneities. A promising alternative for future developments
with good electrical conductivity are carbon nanotubes. These are microscopic tubular structures
(molecular nanotubes) consisting of carbon atoms. They are commercially available but have not
reached a broad range of application yet. Further development is still required to reliably control the
manufacturing process and to reproducibly achieve the intended results of the resulting product.
In many cases, especially in cable accessories or in rotating electrical machines, nonlinear field grading
systems have turned out to be electrically very effective and at the same time space saving. Field
grading can be achieved by very thin layers with nonlinear electric properties and are unrivaled in many
applications. As fillers for nonlinear field grading materials either silicon carbide (SiC) particles or zinc
oxide (ZnO) microvaristors are used in commercially available products. A new approach with antimony
doped tin oxide is still under investigation. SiC has been used for quite a long time, but has some
shortcomings such as only weak nonlinearity (which, however, is sometimes desirable), poor
reproducibility of the electric properties and a mechanical hardness that causes strong mechanical wear
of the manufacturing tools. These drawbacks are overcome by the use of ZnO microvaristors, which
have been applied mainly to cable accessories since around the year 2000. The theory of both types of
nonlinear fillers is basically discussed, and Section 3.3 concludes with background information about
some quite new approaches: polyanilines (most often in the form of PANI-EB), graphene oxides, and
tin oxide (SnO2) as semiconductive coating of mica particles. Finally, potential anisotropy (either
intended or unintended) is addressed, which may be a concern with filler particles of high aspect ratios.
It can cause wanted as well as unwanted effects, which should be kept in mind.
Section 3.4 informs about measurement metrics of field grading materials and systems. Matrix materials,
fillers and the resulting composites necessitate different and independent characterization approaches.
As an example, the polymer matrix requires consideration of the physical properties of the uncured
polymer (rheology), as well as of the intrinsic physical properties of the cured polymer. The latter
includes electrical properties (permittivity, tan , bulk and surface resistivity with their E-field and
temperature dependencies, breakdown strength, etc.), thermal properties (glass transition temperature,
melting point, etc.), and mechanical properties (tensile, flexural and compressive strength, etc.).
Detailed lists of all parameters of interest are presented, which give an idea about the manifold
challenges in the design process of a new field grading system.
The application of nonlinear field grading systems is comparatively new but has gained increasing
importance with the currently emerging DC applications (DC cable accessories, DC GIS). The electrical
characterization of nonlinear field grading materials gives rise to two major challenges: choosing
appropriate models and termini to describe the electrical characteristics, and measuring these
characteristics in a well-defined and reproducible way. So far there are no standardized approaches for
neither of these aspects, and this Technical Brochure, therefore, makes proposals and gives
recommendations for future standardization. Without such standards, nonlinear grading systems are
extremely difficult to develop, as typically all manufacturers of fillers or of final composites may choose
different approaches of characterizing their products. Today, it is hardly possible to communicate the
requirements on a filler in order to develop a tailored grading system. Too many parameters are defined
and determined in different ways by different manufacturers. One section (3.4.5) is, therefore, devoted
to this aspect, on the example of a nonlinear DC field grading system. A suitable terminology is
proposed, and mathematical descriptions of the nonlinearity (conductivity as a function of the applied
electric field) from the literature are presented. Not less important are clear rules for reproducibly
characterizing the filler properties and those of the final composite. One current finding is that at present
it is still not possible to forecast the electrical properties of a compound just by electrically characterizing
the filler. The resulting composite may have surprisingly different electrical properties. Systematically
designing “tailored functionality” of nonlinear field grading systems thus remains a future goal and still
requires further research. At the end of section 3.4.5, a design example for a DC cable joint is given in
greater detail.
By the end of Chapter 3, a comprehensive understanding of the various possible field grading systems
(matrix materials and fillers) should have been achieved, which provides an important background for
making decisions on appropriate field grading systems in individual applications. With the knowledge of

110
TB 794 - Field grading in electrical insulation systems

the specific properties and the different options, optimal approaches adapted to the intended
functionality can be chosen.
In Chapter 4, current approaches of modeling and simulating electric field grading systems are
presented. Though Transient Network Analysis (TNA) simulations are possible and often applied even
for thermo-electrically coupled calculations, the rapid progress in electric field simulation software
packages and consequently reduced need for expensive hardware enables not only highly-specialized
users to perform nonlinear fully coupled 3D field simulations, which typically yield realistic and detailed
results on electric field and temperature distributions. Nevertheless, experimental validation of the
simulation approaches and results is still indispensable. Measuring the electric field and potential
distributions along a field grading layer is quite challenging because this needs to be performed
capacitively, without direct contact to the surface. A very effective approach is the use of Pockels
sensors, which can give results of high accuracy in the time domain, and thus being applicable to DC,
AC as well as transient field stressed systems. An example on the investigation of the end corona
protection of stator bars for rotating electric machines is given in this chapter. Pockels sensors might
not be generally available. But anyway field grading systems, especially nonlinear ones, require
evaluation of thermal heating as a result of the electric field distribution and the generated power losses.
Thus, in many cases, thermal imaging is a good choice in order to assess the overall efficiency of a field
grading system, as long as it is externally accessible. Another, indirect means to validate field simulation
results is to measure partial discharge inception voltages, which are good indicators of excessively high
electric field stress. On externally accessible systems this can be done optically by UV sensitive “corona
cameras”. But more common and always applicable are “classical” electrical measurements of apparent
charges and the evaluation of charge patterns.
Besides explaining the theoretical backgrounds and approaches of electric field grading, this Technical
Brochure is intended to inform about the most important applications that make use of electric field
grading, and to address the related major challenges. The chosen examples of Section 5 are cable
accessories, plug-in systems and separable connectors, transformers, bushings, rotating electrical
machines, and insulators. Some major statements and conclusions are summarized here.
Following the relevant IEC and IEEE standards for MVAC and HVAC cable accessories, the effectiveness
of the field grading performance is to be verified at elevated power frequency voltages as well as
transient stresses. Temperature effects are partially considered as well, although their focus is mainly
on the thermo-mechanical stress of the components and their impact on critical interfaces. The aspects
of harmonics and ageing of field grading materials may, however, not be covered sufficiently by these
standards, which apply mainly but not exclusively to electrically linear materials. For electrically
nonlinear materials the common practice to accelerate degradation by increasing the electrical stress
level may have to be re-considered since operating these materials above the regular operating stress
may initiate additional aging mechanisms and failure modes, not being relevant to the application. Most
recent developments of cables and their accessories are for HVDC systems, where the 320 kV level
seems to be state of the art and 550 kV and 640 kV are under development. Main field grading principles
for HVDC cables rely on nonlinear resistive grading, but also geometrical grading is applied, using
improved new rubber insulation materials with suitable DC conductivity and less space charge formation.
Though first type and prequalification tests on HVDC cables of 320 kV and above have been successfully
passed, the technology is still not finally established and needs further development. An important driver
for these activities are plans in Germany and other parts of Europe to implement DC underground
transmission lines of several Gigawatts over several hundreds of kilometers.
Regarding pluggable systems and separable connectors, today outer and inner cone systems are
available. The electric field inside a separable connector is typically graded geometrically with deflectors
made from conductive materials such as carbon black filled elastomers. Alternatively, conductive varnish
or coatings are used. In contrast to joints and terminations, where the electric strength of the interface
between the cable insulation and the stress cone is decisive, separable connectors have an additional,
very critical interface between the stress cone and the socket or bushing, respectively. Both interfaces
have to withstand the applied voltage stress. For that purpose, a sufficiently high contact pressure
between the layers has to be ensured, which belongs to the main design challenges of these devices.
While outer cone separable connectors are widely used for connections in the medium voltage range
and lower transmission voltages (72.5 kV), inner cone systems are available for system voltages up to
550 kV (AC). They require a metal encapsulation and are thus completely screened. Pluggable joints
and terminations with nonlinear microvaristor filled field grading layers are also available to date.

111
TB 794 - Field grading in electrical insulation systems

Oil-barrier insulation systems were introduced into power transformers during the 1960’s and are still a
commonly used solution. This pressboard barrier technology is based on liquid insulation, which has to
ensure basic electric strength by filling and impregnating the whole transformer’s insulation structure,
while the solid insulation is used to form barrier systems dividing highly stressed oil channels into smaller
gaps. The subdivision of long oil gaps effectively reduces the hazard of long particles and fibre bridges
formation in the field stressed zones. It leads to permitting higher field stresses and less space,
respectively, and hence lower cost, when compared with full oil insulation.
Today, new natural ester based fluids are becoming increasingly attractive, and occasionally the end
user wishes to retrofit existing transformers with a new insulation fluid. However, ester fluids have
higher permittivities than mineral oils, and sometimes unexpected effects on internal field grading have
to be regarded when changing the overall insulation system of a transformer.
Special attention is required for converter transformers, as the oil-pressboard combined insulation near
the valve side is exposed to a composite AC-DC voltage. The design of the insulation system of
transformers used for HVDC applications is quite different to that of transformers for AC applications.
The basic research on the dielectric behavior under DC stress especially of liquid insulating materials
and the interfaces to solid insulation is still ongoing. The focus of the research activities is on the physical
mechanisms contributing to oil conductivity, i.e. charge carrier generation, transport and recombination
as well as the formation of space charges and the interaction of liquid and solid barrier interfaces
(surface charges). For analyzing the DC behavior of the materials, methods like the pulsed electro-
acoustic (PEA) method or measurements using the Kerr-effect of dielectric liquids allow the direct
investigation of materials under DC. The research leads to a significantly new understanding of the
insulation systems, and hence it may be that in future new design rules and methods for modelling the
behavior of transformer insulation systems under DC stress will be introduced.
High voltage bushings are used to guide a conductor that is on HV potential through a grounded wall,
e.g. the tank of a transformer. Ungraded bushings can only be used for rather low voltages because of
the low inception voltage of surface discharges upon the insulation. For medium voltages, geometrically
graded bushings are typically used, whereas for voltages above approx. 70 kV geometrically and
capacitance graded bushings are required. Recent developments of compact high-voltage bushings
make use of layers of nonlinear field grading material, for instance microvaristors, directly applied to
the outer surface of a hollow core insulator as an interface between the glass-fibre reinforced plastic
tube and the weather sheds. This allows for very compact designs with simplified internal geometric
grading measures, referred to as “slim bushings” in the literature.
In rotating electrical machines, field grading is applied to suppress local field enhancement at the stator
slot end. This field grading system, often called “end corona protection,” is mandatory for machines
with rated voltages of approximately 6 kV and higher. The mechanical, thermal and electrical operating
conditions of these insulation systems are extremely demanding. Depending on manufacturing process
or bar design, tapes or varnishes are nowadays used for field grading. Typically, moderately nonlinear
SiC fillers are applied. Alternatives have been investigated, e.g. based on ZnO microvaristors or tin
oxide, but they could not replace classical concepts so far. While each layer of varnish can be electrically
adjusted by using different types or amounts of filler particles, tapes have a defined electrical resistance
characteristic, pre-defined by the tape manufacturer. This disadvantage has to be compensated by
painting different lengths of layers or by wrapping different types of tapes stepwise. The resulting
innumerable degrees of freedom require a comprehensive understanding of the system behavior and
many years of design and operating experience.
Converter fed drives require special approaches. Electric field grading materials that perform satisfyingly
under power-frequency (50/60 Hz) may experience a premature failure when exposed to pulse width
modulated (PWM) voltage source converters (VSC). Field grading materials designed for power-
frequency applications may not be capable of controlling the stress caused by high d U/dt on the surface
of the winding ends.
Also the performance of “simple” insulators can principally be improved by electric field grading
measures. For ceramic line insulators, resistive glazing has been applied in order to improve the
performance under pollution, but without greater success. A similar approach was investigated for
composite line insulators by applying a nonlinear, microvaristor based layer, but has not gained major
importance as well. Microvaristor based field grading has also been investigated for composite hollow
core insulators, thus allowing for “slim bushings”, and for DC GIS insulators, here in order to better

112
TB 794 - Field grading in electrical insulation systems

control the field and temperature dependance of conductivity and thus electric field distribution, and to
avoid charge carrier accumulation.
Future developments might be the application of 3D printing, which seems to be a very promising means
of easily creating spatially and functionally graded insulators and HV equipment, and which is currently
under investigation. HVDC insulation and field grading systems will gain in importance, as well as
“smarter” insulation systems that self-adapt their electrical and mechanical properties to the actual
environmental and operational service conditions. One basic problem especially with novel, nonlinear
field grading approaches is the lack of nonlinear fillers available on the market. Applications like cable
accessories, insulation of rotating electrical machines, line and hollow core insulators or high-field GIS
insulators require very different degrees and a wide range of nonlinearity and “switching” field strengths,
i.e. the electric field strength necessary to make the material become more conductive. The market for
such applications is possibly too small or the need for performance improvement too low to drive the
development of such materials for novel applications. A systematic and predictive tailoring of nonlinear
field grading systems is, therefore, not yet possible, and current approaches are often “trial and error”
based with occasionally unexpected and surprising, sometimes wanted, sometimes unwanted results.
Besides the lack of appropriate filler materials, also the lack of standardization is still impeding a
development into this direction. This Technical Brochure shall contribute to improve the situation, by
giving a broad overview about electric field grading concepts, as well as explaining special aspects in
detail. It further gives recommendations for modeling, characterization and standardization especially
for nonlinear field grading systems, which may be used as input for future standardization of
requirements and adequate test procedures.

113
TB 794 - Field grading in electrical insulation systems

7. Terms and definitions


7.1 Symbols and units

Symbol Unit Meaning


- Indicates a complex number/quantity (below symbols)
∂ - Partial derivative (in front of other symbols)
Asurface m² Total area of the surface
𝛼 °; - Geometric angle; Exponent of nonlinearity
C1 Constant (Equ. 3.1)
C2 Constant (Equ. 3.2)
csurface F/m² Specific surface capacitance
Csurface F Total surface capacitance
d m Thickness (of the insulation)
d - Derivative (in front of other symbols)

𝐷 As/m² Electric displacement field vector

Dc m Diameter of the conductor


Di m Diameter of the cable insulation
Dp m Particle diameter
𝛿 ° Loss angle
e - Euler’s number (constant): e = 2.718281828

𝐸⃗ V/m Electric field vector

E V/m Magnitude of the electric field


E1 eV Activation energy for the formation of an ion
E2 eV Activation energy for the mobility of an ion
Eb V/m Switching field strength in the percolated case
Eg eV Bandgap energy
En,max V/m Maximum field strength in layer n
Esat V/m Saturation field strength
𝜀 As/Vm Permittivity
𝜀0 As/Vm Permittivity of vacuum (constant): 𝜀0 = 8.854∙10-12 As/Vm
𝜀𝑟 - Relative permittivity
𝜀𝑟‘ - Real part of the complex relative permittivity
𝜀𝑟‘‘ - Imaginary part of the complex relative permittivity
𝜀𝑟∗ - Complex relative permittivity
FD N Viscous force
g V/m Longitudinal electric field stress alongside the stress cone
(Equ. 5.1)

114
TB 794 - Field grading in electrical insulation systems

G m/s² Acceleration
 - Experimental parameter
I A Electric current
ich A Current in the discharge in Toepler’s configuration
j - Imaginary unit
J A/m² Magnitude of the electric current density
𝐽 A/m² Electric current density vector
⃗⃗𝐽c A/m² Electric current density of the conduction current
⃗⃗⃗
𝐽d A/m² Electric current density of the displacement current

k m²∙kg/(s²∙K) Boltzmann constant: k = 1.38064852∙10-23 m²∙kg/(s²∙K)


K kV/cm-1/2 Factor for surface discharge inception voltage (Section
2.2.3.2)
Kc - Geometric and material factor (Equ. 2.24)
Lg m Grading length
m0 S/V Initial slope of  -E-characteristic (Section 3.4.5.1.2)
m S/V Slope of  -E-characteristic (Section 3.4.5.1.2)
M kg/m³ Mass of a filler particle (Equ. 2.25)
µ (m/s)/(V/m) mobility of charge carriers (Equ. 3.2)
n - Number of the layer
N - Maximum number of layers (Section 2.2.4.1), number of
ions (Section 3.1.1)
N1 - Shaping parameter (Section 3.4.5.1.2)
N2 - Shaping parameter (Section 3.4.5.1.2)
P W Real Power
𝜑 ° Phase angle
π Constant: π = 3.1459…….
𝜎 S/m Electric conductivity
𝜎0 S/m Base conductivity
𝜎𝑠𝑎𝑡 S/m Saturation conductivity
Q W Reactive power
Qsurface As Total surface charge
r m Radius
RL Ω Resistance alongside the insulation
Rs Ω Surface resistance
𝜌f kg/m³ Density of the fluid
𝜌𝑝 kg/m³ Density of the particle
s m Thickness of the insulating material
t s Time

115
TB 794 - Field grading in electrical insulation systems

 s Relaxation time constant


𝜏LI s Relaxation time constant for a lightning impulse (Section
3.4.5.3)
𝜏SI s Relaxation time constant for a switching impulse (Section
3.4.5.3)
tan 𝛿 - Dielectric dissipation factor
tan 𝛿c - Dielectric dissipation factor caused by conduction current
tan 𝛿pol - Dielectric dissipation factor caused by polarization
𝜗 °C Temperature
T K Absolute temperature
Tg °C Glass transition temperature
𝜗conductor °C Temperature of the conductor
𝜗enclosure °C Temperature of the enclosure
usurface V Surface voltage in Toepler’s configuration
U V Voltage
Ua V Partial voltage of an insulator
Ui V Inception voltage
Um V Highest voltage for equipment (phase-to-phase voltage,
r.m.s. value)
vr m/s Relative velocity of a filler particle
vt m/s Terminal velocity of a filler particle
Wa eV Activation energy
𝑊𝑎𝑓 eV Activation energy for a formation of an ion

𝑊𝑎𝑚 eV Activation energy for the mobility of an ion


𝜔 s-1 Angular frequency, angular speed
x m Geometric dimension
y m Geometric dimension
z m Geometric dimension

7.2 Acronyms
Note: Standardized acronyms for plastics, rubbers and lattices can be found in ISO 1043 [ISO 2016]
and ISO 1629 [ISO 2013]. As far as they are used in this Technical Brochure they are also listed here.

Acronym Meaning

AC Alternating current

AIS Air insulated systems

ATO Antimony doped tin oxide

BEM Boundary element method

116
TB 794 - Field grading in electrical insulation systems

BPA Bisphenol A

CB Carbon black

CNT Carbon nanotubes

CTE Coefficient of thermal expansion

DC Direct current

ECP Electronic conductive particles/ End corona protection

EHVAC Extra high voltage AC

EVA Ethylene-Vinyl Acetate

EP Epoxy Resin

EPDM Ethylene-propylene-diene-rubber

EPM Ethylene-propylene-rubber

EQS Electro quasi static

FDM Finite element method

FEM Finite difference method

FFO Fast front overvoltage

-FGM Field grading by grading part that shows a gradually changing permittivity

FGA Field grading adapter

FGM Functionally graded material

FTIR Fourier-transform infrared spectroscopy

GEM General Effective Media

GHP-FGM Gradually Higher Permittivity – Field Grading Materials

GIL Gas insulated lines

GIS Gas insulated switchgear

GLP-FGM Gradually Lower Permittivity – Field Grading Materials

GO Graphene oxide

GPLs Graphene platelets

HTV High temperature vulcanizing silicone rubber

HV High-voltage

HVAC High-voltage AC

HVDC High-voltage DC

ICP Intrinsic conductive polymer

ICP-MS Inductively Coupled Plasma – Mass Spectrometry

LI Lightnig impulse

LSR Liquid silicone rubber

117
TB 794 - Field grading in electrical insulation systems

MFF Minatec® functional fillers

MO Metal-oxide

MVAC Medium voltage AC

OCP Outer corona protection

OIP Oil impregnated paper

PANI-EB Emeraldine base form of Polyaniline

PE Polyethylene

PEVA Poly(Ethylene-Vinyl Acetate)

PD Partial discharge

PMN-PT Lead magnesium niobium-lead titanate

PQT Prequalification test

PR Polarity reversal

PUR Polyurethane

PVC Polyvinylchloride

PVDF Poly(vinylidenfluoride)

PWM Pulse width modulation

RBP Resin bonded paper

RH Relative humidity

RIP Resin impregnated paper

RIS Resin impregnated synthetics

r.m.s. Root mean square

RTV Room temperature vulcanizing silicone rubber

SCLC Space charge limited current

SEM Scanning electron microscopy

-FGM Field grading by grading part that shows a gradually changing conductivity

SFO Slow front overvoltage

SI Switching impulse

TGA Thermogravimetric analyses

TNA Transient network analyses

TOV Transient overvoltage

TT Type test

UV Ultraviolet

VFFO Very fast front overvoltage

VIS Visible light

118
TB 794 - Field grading in electrical insulation systems

VMQ Vinyl methyl silicone rubber

VRH Variable range hopping

VSC Voltage source converter

XLPE Crosslinked polyethylene

119
TB 794 - Field grading in electrical insulation systems

8. References
Acronym Authors, title, publication data Clause

[ABB 2004] Patent WO2004038735A1: Field grading material 5.1.3

[ABB 2010] Patent US20100147556A1: Electric insulation arrangement 5.1.3

[ABB 2014] Patent US20140116746A1: Device for electric field control 5.1.3

[Ahmad 2015] Ahmad, H.; Tsukamoto, N.; Haddad, A.; Griffiths, H.; 3.4.3
Robson, S.: Electrical Characterisation of ZnO microvaristor materials and compounds.
Proceedings 2015 IEEE Conference on Electrical Insulation and Dielectric Phenomena
(CEIDP 2015), ISBN 978-1-4673-7499-6

[Ahmad 2017] Ahmad, H.; Robson, S.; Albano, M.; Haddad, A.: Characteristics of ZnO Microvaristors- 3.4.5
Loaded Grading Polymeric Materials, INSUCON 2017, 16-18 May 2017, Birmingham.

[Aladenize 1995] Aladenize, B.; Assier, J.C.; Galaj, S.; Mirebeau, P.; Janah, H.: Materials for cables based on 3.3.4
polymeric conductors. JICABLE’95, pp. 375-379, 1995.

[Amerpohl 2002] Amerpohl, U.; Kirchner M.; Böttcher B.; Malin G.: Dry Type Outdoor Termination with new 5.1.1
Stress Control Management. CIGRE Session 2002 , report 21-106 5.1.2
5.2.3
[Andersson 2000] Andersson, T.; Forsmark, S.; Jaksts, S.: DryformerTM – a new type of oil-free power 5.3.4
transformer with low environmental impact. ABB POWERFORMER, ABB Review 3/2000

[Backhaus 2014] Backhaus, K.; Speck, J.; Hering, M.; Großmann, S.; Fritsche, R.: Nonlinear dielectric 5.3.3
behaviour of insulating oil under HVDC stress as a result of ion drift. IEEE Int. Conf. High
Volt. Eng. (ICHVE), 2014

[Barber 2009] Barber, P.; Balasubramanian, S; Anguchamy, Y.; Gong, S.; Wibowo, A.; Gao, H.; Ploehn, 3.1.2
H. J.; zur Loye, H.-C.: Polymer Composite and Nanocomposite Dielectric Materials for Pulse
Power Energy Storage. Materials 2009, 2, 1697-1733

[Bärsch 2008] Bärsch, R.; Kindersberger, J.: Nichtlineare dielektrische Funktionseigenschaften von 3.1.1
Dielektrika. ETG-Workshop „Werkstoffe mit nichtlinearen dielektrischen Eigenschaften“, 3.1.2
Stuttgart, 13.3.2008, ETG-Fachbericht 110, VDE-Verlag Berlin, pp. 7-33

[Baumann 2011] Baumann, T.: Electrical Stress Grading Systems for Generators with Rated Voltages above 2.2.3
20 kV ETG-Workshop „Feldsteuernde Isoliersysteme“, Darmstadt, November 2011 4.4.2
5.5.1
5.5.2
5.5.3
[Beyer 1986] Beyer, M.; Boeck, W.; Möller, K.; Zaengl, W.: Hochspannungstechnik, Theoretische und 3.1.2
praktische Grundlagen für die Anwendung. Springer Verlag Berlin/Heidelberg/New York; 5.6.1
1986

[Blatt 2015a] Blatt, S.; Hinrichsen, V.: Mathematical model for numerical simulation of current density in 3.4.5
microvaristor filled insulation materials. IEEE Transactions on Dielectrics and Electrical
Insulation Vol. 22, No. 2; April 2015, pp. 1161-1170. DOI 10.1109/TDEI.2014.004740

[Blatt 2015b] Blatt, S.: Untersuchungen zu einem möglichen Einsatz von Mikrovaristoren in der Isolation 3.4.5
umrichtergespeister Antriebe. Doctoral Thesis, TU Darmstadt, 2015, tuprints.ulb.tu-
darmstadt.de/5181/

[Boettcher 2001] Boettcher, B.; Malin, G.; Strobl, R.: Stress control system for composite insulators based 5.1.1
on ZnO-technology. 6.6.2
Transmission and Distribution Conference and Exposition, 2001 IEEE/PES, Year: 2001,
Volume: 2, Pages: 776 - 780 vol. 2, DOI: 10.1109/TDC.2001.971336

[Böning 1953] Böning, W.: Kleines Lehrbuch der elektrischen Festigkeit. Braun-Verlag, Karlsruhe, 1953 2.2.3

[Brandes 2008] Brandes, H.; Hillmer, T.; Stebler, P.: Halbleitende Lacke und Bänder – für die Anwendung 3.3.1
nützliche Ergebnisse und Eigenschaften. ETG-Workshop „Werkstoffe mit nichtlinearen 3.3.3
dielektrischen Eigenschaften“, Stuttgart, 13.3.2008 5.5.1

[Brinkmann 1975] Brinkmann, C.: Die Isolierstoffe der Elektrotechnik. Berlin, Heidelberg: Springer, 1975 3.2.2

[Brütsch 2006] Brütsch, R.; Hillmer, T.: Corona Protection in Rotating High Voltage Machines. INDUCTICA 5.5.1
2006 Conference, Berlin, pp. 1-6 5.5.2

[Capuder 2013] Capuder, K.; Štih, Ž.; Plišić, G.: Choice of Electric field Calculation Method According to the 5.3.2
Dielectric Design Criteria. Compumag 2013

120
TB 794 - Field grading in electrical insulation systems

[Carlson 1996] Carlson, Å.: Specific requirements on HVDC converter transformers. ICEE, 1996. 5.3.3

[Chen 2018] Chen, Q.; Zhang, J.; Chi, M.; Guo, C.: Breakdown Characteristics of Oil-Pressboard 5.3.1
Insulation under AC-DC Combined Voltage and Its Mathematical Model, Energies 2018, 11,
1319

[Christen 2010] Christen, T; Donzel, L.; Greuter, F.: Nonlinear Resistive Electric Field Grading Part 1: 3.4.5
Theory and Simulation. IEEE Electrical Insulation Magazine, Nov./Dec.- Vol. 26, No. 6,
2010

[Cigre 1995] CIGRE TB 89: Accessories for HV extruded cables. 1995 5.1.2

[Cigre 2009] Cigre TB 379: Update of Service Experience of HV Underground and Submarine Cable 5.1.1
Systems. WG SC B1, February 2011

[Cigre 2012] CIGRE TB 496: Recommendations for testing DC extruded cable systems for power 3.4.5
transmission at a rated voltage up to 500 kV. 2012 5.1.3
5.6.3
[Cigre 2015] Cigré WG B1/B3.33: Feasibility of a Common, Dry Type Plug-In Interface for GIS and 5.2.2
Power Cables above 52 kV. Cigré TB 605, January 2015

[Cigre 2016] Cigre TB 646: HVDC Transformer insulation: Oil Conductivity. JWG A2/D1.41, January 5.3.3
2016

[Cigre 2019] Cigre TB 784 : Standard design of a common, dry type plug-in interface for GIS and power 5.2.2
cables up to 145 kV. JWG B1/B3.49. November 2019

[Cottevielle 1999a] Cottevielle, D.; Le Méhaute, A.; Challioui, C.; Mirebeau, P.; Dernay, J.N.: Industrial 3.3.4
applications of polyaniline. Synthetic Metals, v. 101, pp. 703-704, 1999.

[Cottevielle 1999b] Cottevieille, D.; Cariou, F.; Mirebeau, P.; Demay, J.N.; Parasie, Y.: New potentialities of 3.3.4
conductive polymers for cables. Proceedings of 8th International Conference on Power
Insulated Cables (Jicable 99), Paris, France, pp. 845-848, 1999.

[Cui 2011] Cui, L.; Lu, X.; Chao, D.; Liu, H.; Li, Y.; Wang, C.: Graphene-based composite materials 3.3.4
with high dielectric permittivity via an in situ reduction method. Phys. Status Solidi A 208,
pp. 459–461.

[Dahinden 1998] Dahinden, V.; Schultz, K.; Küchler, A.: Function of Solid Insulation in Transformers. 5.3.1
Transform 98, Munich, 1998 5.3.3

[Dai 2004] Dai, L.: Intelligent Macromolecules for Smart Devices. ISBN 1-85233-510-6, 2004 3.1.1

[Debus 2010a] Debus, J.; Hinrichsen, V.; Seifert, J.; Hagemeister, M.: Investigation of composite 5.6.1
insulators with microvaristor filled silicone rubber components. International Conference on
Solid Dielectrics (ICSD), Potsdam, July, 2010

[Debus 2011a] Debus, J.; Tenzer, M.; Hinrichsen, V.: Verbundlangstabisolatoren mit 5.6.1
mikrovaristorgefüllten Silikonelementen. Burghauser Isolierstoffkolloquium, 9.-10. Juni
2011

[Debus 2011b] Debus, J.; Hinrichsen, V.; Seifert, J.: Improved Performance of Silicone Rubber Composite 5.6.1
Insulators by Micro-Varistor filled Components. XVII International Symposium on High
Voltage Engineering, Hannover, Germany, August 22-26, 2011, paper D-076, ISBN 978-3-
8007-3364-4

[Debus 2015] Debus, J.: Untersuchung der Anwendungsmöglichkeiten mikrovaristorgefüllter 3.4.5


Feldsteuerelemente in der elektrischen Energietechnik. Doctoral Thesis, TU Darmstadt, 5.6.1
2015, tuprints.ulb.tu-darmstadt.de/5129/

[Dodds 2010] Dodds, S.; Railing, B.; Akman, K.; Jacobson, B.; Worzyk, T.; Nilsoon, B.: HVDC VSC (HVDC 5.1.3
light) transmission – operating experiences. Cigré Session 2010, paper B4_203_2010

[Donzel 2004] Donzel, L.; Christen, T.; Kessler, R.; Greuter, F.; Gramespacher, H.: Silicone composites 5.1.1
for HV applications based on microvaristors. Proceeding of IEEE International Conference 5.1.2
on Solid Dielectrics, Toulouse (2004) 403

[Donzel 2007a] Donzel, L.; Hagemeister, M.: Nonlinear field control of high voltage bushings. Proceedings 5.6.2
ISH 2007, Ljubljana (2007), paper T3-73

[Donzel 2011] Donzel, L.; Greuter, F.; Christen, T: Nonlinear Resistive Electric Field Grading Part 2: 3.3.1
Materials and Application. IEEE Electrical Insulation Magazine, March/April - Vol. 27, No. 2, 3.3.3
2011 3.3.4
5.6.2
[Donzel 2016] Donzel, L.; Montenegro-Uratsun, M.; Hagemeister, M.; Rukwid, P.: ZnO stress grading 5.5.1
tape for stator windings for electrical machines located at higher altitudes. Cigré Session
2016, Paris, paper D1-107

121
TB 794 - Field grading in electrical insulation systems

[Drumm 2015] Drumm, F.: Heat Shrink Cable Accessories: A 40 Years Proven Technology for Applications 5.1.1
in Harsh Environments; INMR World Congress, October 18-21, 2015, Munich

[Egiziano 1987] Egiziano, L.; Lupo, G.; Tucci, V.: An approach to the design of protected semiconducting 5.6.1
glazed insulators. 5th ISH, Braunschweig, Germany, Aug. 24 – 28, 1987, paper 52.04

[Egiziano 1995] Egiziano, L.; Tucci, V.; Lupo, G.; Petrarca, C.: Electrical properties of different composite 3.3.4
materials for stress relief in HV cable accessories. IEEE, CEIDP-95,
doi:1109/CEIDP.1995.483588.

[Eigner 2013] Eigner, A.; Semino, S.: 50 years of electrical-stress control in cable accessories. IEEE 5.1.1
Electrical Insulation Magazine, Vol. 29, No.5, September/October 2013

[Elsässer 2001] Elsässer, O.; Messerer, F.; Feser, K.: Flashover Behavior of Semiconducting Glazed 5.6.1
Insulators Under Positive Lightning Impulse Stress at Different Climatic Conditions, 9. Int.
Conf. on Gaseous Dielectrics, Elicott City, Maryland USA, May 21-25, 2001

[Ericsson 2003] A. Ericsson, M. Jeroense, J. Miller, L. Palmkvist, B. Railing, P. Riffon: HVDC Light Cable 5.1.3
Systems - The latest Projects.
Nordic Insulation Symposium (Nordis), Tampere, June 11-13, 2003

[Espino 2005] Espino-Cortes, F.P.; et al.: Effectiveness of Stress Grading Coatings on FormWound Stator 5.5.3
Coil Groundwall Insulation Under Fast Rise Time Pulse Voltages.IEEE Transactions on
Energy Conversion, Vol. 20, No. 4, December 2005

[Feast 1996] Feast, W. J.; Tsibouklis, J.; Pouwer, K. L.; Groenendaal, L.; Meijer, E.W.: Synthesis, 3.3.4
processing and material properties of conjugated polymers, Polymer Vol. 37 No. 22, pp.
5017-5047, 1996

[Feier-Iova 2009] Feier-Iova, S.: The Behaviour of Water Drops on Insulating Surfaces Stressed by Electric 5.6.1
Field, Doctoral Thesis, TU Darmstadt, 2009

[FLAME 2019] The LOEWE project FLAME. www.flame.tu-darmstadt.de/flame/index.en.jsp 2.2.3

[Forrest 1940] Forrest, J.S.; Marshall, C.W.: Improvement relating to HV insulators, Brit. Patent 527357, 3.2.1
1940

[Forrest 1942] Forrest, J.S.: The characteristics and performance in service of HV porcelain insulators, 3.2.1
Journal of IEE, Vol. 89, Part ll, pp. 60-92, 1942.

[Fu 2015] Fu, M.; Lou, B.; Hou, S.; Liao, Y.; Hao, M.; Chen, G.: Space Charge Dynamics in 5.3.3
Pressboard-Oil-Pressboard Multilayer System Under DC Voltages. IEEE Int. Conf. Prop.
Appl. Dielectr. Mater. (ICPADM), 2015

[Fukui 1978] Fukui, H.; Fujimura, T.; Naito, K.; Irie, T.: Studies on various performance of 3.2.1
semiconducting glazed insulators, NGK Review, No. 2, pp. 2341,1978

[Gabler 2019] Gabler, T.; Backhaus, K.; Großmann, S.; Fritsche, R.: Dielectric Modeling of Oil-Paper 5.3.3
Insulation Systems at High DC Voltage Stress Using a Charge Carrier-based Approach.
IEEE Trans. Dielectr. Electr. Insul., Vol. 26, No. 5, 2019

[Gäfvert 2000] Gäfvert, U.; Jakst, A.; Jorendal, G.: Effect of DC steady state and transient stress on 5.3.3
electrical insulation systems of converter transformers. Cigré Session 2000, Paris

[Gaponov 2011] Gaponov, A.V.; Glot, A.B.: Non-ohmic conduction in tin dioxide based ceramics with copper 3.3.4
addition. Semiconductor physics, quantum electronics & optoelectronics, V.14, N1, pp. 71-
76, (2011)

[Glot 2010] Glot, A. B.; Bondarchuk, A.N.; Ivon, A.I.; Fuentes, L.; Aguilar-Martinez, J.A.; Pech-Canul, 3.3.4
M.I.; Pineda-Aguilar, N.: Influence of Joule Heating on Varistor Behavior in SnO2–Co3O4–
Nb2O5–Cr2O3 Ceramics. XIX Int. Conf. on Extractive Metallurgy, Saltillo, Coahuila, Mexico
(May 18-21, 2010, paper 52) p. 572.

[Glot 2011] Glot, A.B.; Lu, Z.Y.; Ivon, A.I.: Electrical properties of SnO2 ceramic varistor withstanding 3.3.4
high current pulses. Superficies y Vacio 24(2) pp. 61-67, 2011

[Gramespacher 2003a] Gramespacher, H.; Christen, T.; Donzel, L.; Greuter, F.: Microvaristor based field grading 5.1.2
elements for HV terminations. JICABLE '03 (International Conference on Insulated Power
Cables), Conference proceedings (Versailles, France, 2003)

[Gramespacher 2003b] Gramespacher, H.; Svahn, J.; Greuter, F.; Christen, T.: Microvaristor based field grading 5.1.2
elements for HV terminations. XIIIth ISH, Delft, August 2003

[Greb 2015] Greb, M.; Bauer, J.: Elektrisch-halbleitfähige Metalloxid-Pigmente mit nichtlinearen 2.2.4
elektrischen Eigenschaften zum Einsatz in der Hochspannungstechnik. RCC-Fachtagung 3.3.1
„Werkstoffe – Forschung und Entwicklung neuer Technologien zur Anwendung in der 3.3.4
elektrischen Energietechnik. Berlin, 20th/21st May, 2015

122
TB 794 - Field grading in electrical insulation systems

[Greuter 1995] Greuter, F.: Electrically active interface in ZnO varistors, Solid State Ionics, Vol. 25 (1995) 3.3.3
67-78

[Gubanski 2005] Gubanski, S.: Modern Outdoor Insulators – Concerns and Challenges. IEEE Electrical 3.2.1
Insulation magazine, Nov./Dec. 2005, Vol. 21, No. 6, pp. 5-11 5.6.1

[Gupta 1990] Gupta, T.K.: Application of zinc oxide varistors. J. Am. Ceram. Soc. Vol 73, 1990, 1817- 3.3.3
1840

[Gustafsson 2015] A. Gustafsson, M. Jeroense, H. Ghorbani, T. Quist, M. Saltzer, A. Farkas: Qualification of an 5.1.3
Extruded HVDC Cable System at 525 kV. 9th International Conference on Insulated Power
Cables, JICABLE 2015

[Haddad 1991] Haddad, A.; Rosado, F.; German, D.M.; Waters, R.T.: Equivalent Circuit For Surge Arrester 3.4.3
Elements At Power Frequency Voltages. 25th Universities Power Engineering Conference
(UPEC), vol.1, pp 361-364, Aberdeen (Scotland), 1991

[Haverkamp 1984] Haverkamp, W.; LeBaut, P.; Heat-shrink cable accessories for plastic cables up to 36kV, 5.1.1
Jicable Proceedings 1984, C-2

[Haverkamp 2000] Haverkamp, W.; Malin, G.; Strobl, R.: Termination system for polymeric distribution cables 5.1.1
based on ceramic stressgrading technology. Power Journal of the South African Institute of
Electrical Engineers, Jan./Feb. 2000, pp. 66-69

[Hayakawa 2014] Hayakawa, N.; Kato, K.; Okubo, H.; Hama, H.; Hoshina, Y.; Rokunohe, T.: Electric Field 2.2.4
Grading Techniques in Power Apparatus Using Functional Materials. CIGRE Paris Session, 5.6.3
D1-309, August 24-29, 2014

[Hayakawa 2016] Hayakawa, N.; Ishiguro, J.; Kojima, H. ; Kato, K.; Okubo, H.: Fabrication and Simulation of 2.2.4
Permittivity Graded Materials for Electric Field Grading of Gas Insulated Power Apparatus.
IEEE Transactions on Dielectric and Electrical Insulation, Vol. 23, No. 1, February 2016, pp.
547-554

[Higinbotham 2018] Higinbotham, W; Higinbotham, K.: Review of Medium Voltage Asset Failure Investigations, 5.1.1
Powertest Conference 2018

[Hinrichsen 2011] Hinrichsen, V.; Küchler, A.: Grundlagen der Feldsteuerung. ETG Workshop "Feldsteuernde 2.2.2
Isoliersysteme", 22. & 23. November 2011, Darmstadt, ISBN 978-3-8007-3390-3 2.2.3
4.4.2
5.6.1
[HP 1992] Hewlett Packard: Application Note 1217-1 “Basics of measuring the dielectric properties of 3.1.2
materials”, Hewlett Packard literature number 5091-3300E, 1992

[HSP 2015] HSP: Resin Impregnated Synthetic - The latest generation of innovative bushings. HSP 5.4.1
Pub. RIS Bushings 09/15-500 TH 102-150429.
https://www.hspkoeln.de/cms/upload/downloads/RIS_Flyer_e.pdf (visited March 2019)

[Hussain 2017] Hussain, R.; Hinrichsen, V.: Simulation of thermal behavior of a 320 kV HVDC cable joint 3.4.5
with nonlinear resistive field grading under impulse voltage stress. CIGRÈ Winnipeg 2017
International Colloquium & Exhibition, 30.09.2017 - 06.10.2017, Winnipeg, Canada

[IEC 2000] IEC 60270:2000: High-voltage test techniques – Partial discharge measurements 4.4.2

[IEC 2006] IEC TS 60034-27:2006: Rotating electrical machines – Part 27: Off-line partial discharge 4.4.2
measurements on the stator winding insulation of rotating electrical machines

[IEC 2015] IEC 62631-3-2: Dielectric and resistive properties of solid insulating materials - Part 3-2: 3.4.5
Determination of resistive properties (DC methods) - Surface resistance and surface
resistivity. 2015

[IEC 2016] IEC 62631-3-1: Dielectric and resistive properties of solid insulating materials - Part 3-1: 3.4.5
Determination of resistive properties (DC methods) - Volume resistance and volume
resistivity - General method. 2016

[IEC 2017] IEC 60034-18-42: Rotating electrical machines - Part 18-42: Partial discharge resistant 5.5.3
electrical insulation systems (Type II) used in rotating electrical machines fed from voltage
converters - Qualification tests. Edition 1.0, 2017-02

[IEEE 2008] IEEE Taskforce on Electric Fields and Composite Insulators: Electric Fields on AC 5.6.1
Composite Transmission Line Insulators. IEEE Transactions on Power Delivery, Vol. 23, No.
2, April 2008, pp. 823-830

[Ishibe 2014] Ishibe, S.; Mori, M.; Kozako, M.; Hikita, M.: A New Concept Varistor With 3.3.4
Epoxy/Microvaristor Composite. IEEE Transactions on Power Delivery, Vol. 29, No. 2, April
2014, pp. 677-682

[Ishiguro 2014] Ishiguro, J.; Kurimoto, M.; Kojima, H.; Kato, K.; Okubo, H.; Hayakawa, N.: Electric Field 5.6.3
Control in Coaxial Disk-Type Solid Insulator by Functionally Graded Materials (FGM). 2014

123
TB 794 - Field grading in electrical insulation systems

IEEE Conference on Electrical Insulation and Dielectric Phenomena (CEIDP), Des Moines,
USA, October 19-22, 2014

[ISO 2013] ISO 1629, Fourth edition, 2013-06-15: Rubber and latices — Nomenclature 7.2

[ISO 2016] ISO 1043-1:2011 + Amd.1:2016: Plastics - Symbols and abbreviated terms - Part 1: Basic 7.2
polymers and their special characteristics

[Jeroense 2008] Jeroense, M.; Gustafsson, A.; Bergkvist, M.: HVDC Light cable system extended to 320 kV. 5.1.3
Cigré Session 2008, paper B1-304

[Jeroense 2018] Jeroense, M.; Bergelin, P.; Quist,T.; Abbasi, A.; Rapp, H.; Wang, L.: Fully qualified 640 kV 5.1.3
underground extruded DC cable system, CIGRE 2018 Session, paper B1-309

[Jonsson 2007] Jonsson, E.; Palmqvist, L.: A field grading material. Patent application EP 1 975 949 A1 3.3.4

[Ju 2009] Ju, H.-J.; Ko, K.-C.; Choi, S.-K.: Optimal Design of a Permittivity Graded Spacer 2.2.4
Configuration in a Gas Insulated Switchgear 5.6.3

[Kaufhold 2008] Kaufhold, M.; Weidner, J.R.; Kielmann, F.; Nutzung nichtlinearer dielektrischer 5.5.1
Werkstoffeigenschaften in Isoliersystemen rotierender elektrischer Maschinen –
Anforderungen, Design, Betriebserfahrungen. ETG-Workshop „Werkstoffe mit nichtlinearen
dielektrischen Eigenschaften“, Stuttgart, 13.3.2008, ETG-Fachbericht 110

[KEMA 2018] KEMA: Power Cable System Testing, Position Paper, June 2018 5.1.1

[Kieffel 2000] Kieffel, Y.; Travers, J.P.; Cotteville, D.: Undoped polyaniline in the high voltage domain: 3.3.4
Non linear behavior and aging effects. Proceedings of the 2000 Conference on Electrical
Insulation and Dielectric Phenomena (CEIDP 2000), Victoria, Canada, pp. 52-56, 2000.

[Kimura 1985] Kimura, K.; Hirabayashi, S.: Improved Potential Grading Methods with Silicon Carbide 5.5.1
Paints for High Voltage Coils. IEEE Trans. Electrical Insul. EI-20, 511, 1985

[Kogan 1995] Kogan, V.; Dawson, F.; Gao, G.; Nindra, B.: Surface corona suppression in high voltage 4.1
stator winding end turns. Proceedings: Electrical Electronics Insulation Conference and 4.2
Electrical Manufacturing & Coil Winding Conference, 1995

[Küchler 2017] Küchler, A.: High Voltage Engineering - Fundamentals - Technology - Applications. 2.1.2
Springer Vieweg 2017, ISBN 978-3-642-11992-7 2.2.2
2.2.3
3.2.2
3.2.3
5.3.1
5.4
5.4.1
[Kuhl 2001] Kuhl, M.: FRP rods for brittle fracture resistant composite insulators. IEEE Trans. Dielect. 5.6.1
Elect. Insul., Vol. 8, pp. 182–190, 2001

[Kurimoto 2002] Kurimoto, M.; Kato, K.; Adachi, H.; Sakuma, S.; Okubo, H.: Fabrication and Experimental 2.2.4
Verification of Permittivity Graded Solid Spacer for GIS. 2002 IEEE Conference on Electrical 5.6.3
Insulation and Dielectric Phenomena, pp. 789-792, 2002

[Kurimoto 2010] Kurimoto, M.; Kato, K.; Hanai, M.; Hoshina, Y.; Takei, M.; Okubo, H.: Application of 2.2.4
Functionally Graded Material for Reducing Electric Field on Electrode and Spacer Interface. 5.6.3
IEEE Trans. on Dielectrics and Electrical Insulation, Vol. 17, No. 1, pp.256-263, 2010

[Lambeth 1983] Lambeth, P.J.: The use of semiconducting glaze insulators. Electra no. 86, pp. 89–106, 5.6.1
1983

[Lambrecht 1999] Lambrecht, J.; Pilling, J.; Bärsch, R.; On the Behaviour of Insulations with Interfaces in 5.1.1
Medium Voltage Cable Accessories – Model Investigations. High Voltage Engineering
Symposium 1999, Conference Publication No. 467

[Lang 2013] Lang, S.; Staubach, C.: Neue Generation einer resistiv-kapazitiven Potentialsteuerung 5.5.1
ermöglicht optimiertes Design des Endenglimmschutzsystems - Teil A: Entwicklung und
Herstellung. Conference: „Grenzflächen in elektrischen Isoliersystemen – Beiträge der 4.
ETG-Fachtagung 12.11.2013 - 13.11.2013“, Dresden, Germany

[Lapp 2014] Field-controlled composite insulator and method for producing the composite insulator. 5.6.1
Patent US 8637769 B2, 2014

[Leijon 1999] Leijon, M.; et al.: Powerformer™ is based on established products and experiences from 5.3.4
T&D. Invited Paper for IEEE T&D Committee , Summer Meeting, Edmonton , Alberta
Canada, July 18-22, 1999

[Leijon 2001] Leijon, M.; Dahlgren, M.; Walfridsson, L.; Ming, L.; Jaksts,A.: A Recent Development in the 5.3.4
Electrical Insulation Systems of Generators and Transformers. IEEE Electrical Insulation
Magazine May/June 2001 — Vol. 17, No. 3, pp. 10-15

124
TB 794 - Field grading in electrical insulation systems

[Libert 1997] Libert, J.; Cornil, J.; dos Santos, D.A.; Bredas, J.L.: From neutral oligoanilines to 3.3.4
polyanilines: A theoretical investigation of the chain-length dependence of the electronic
and optical properties. Phys. Rev B, 56(14), pp. 8638-8650, 1997

[Lin 2008] Lin, C.-C.; Lee, W.-S.; Sun, C.-C.; Whu, W.-H.: A varistor–polymer composite with 3.3.4
nonlinear electrical-thermal switching properties. Ceramics International 34, pp. 131–136

[Liu 2016] Liu, Z.; Li, W.; Wang, Y.; Su, G.; Zhang, G.; Cao, Y.; Li, D.: Topology Optimization and 3D- 5.6.3
Printing Fabrication Feasibility of High Voltage FGM Insulator. 2016 IEEE International 6
Conference on High Voltage Engineering and Application (ICHVE), DOI:
10.1109/ICHVE.2016.7800864

[Malamud 2000] Malamud, R.; Cheremisov, I.: Anti-Corona Protection of the High Voltage Stator Windings 5.5.2
and Semi-Conductive Materials for its Realization. International Symposium on Electrical
Insulation, 2000

[Mårtensson 1998] Mårtensson, E.; Nettelblad, B.; Gäfvert, U.; Palmqvist, L.: Electrical properties of field 3.3.4
grading materials with silicon carbide and carbon black. Proceedings of the International
Conference on Conduction and Breakdown in Solid Dielectrics, 1998

[Mårtensson 2003] Mårtensson, E.: Modelling electrical properties of composite materials. Kungl Tekniska 3.3.3
Högskolan Stockholm, Sweden, 2003

[Marusic 1994] Marusic, B.: Efficiency Evaluation of Semiconducting Stress Control System. Electrical 2.2.3
Insulation Conference EIC, Pittsburgh, PA, USA, 1994 4.4.2

[Matsui 1997] Matsui, S.; et al.: State of the art of semiconducting glazed insulators for transmission 5.6.1
lines in heavily contaminated area. 5th Int. Conf. Properties and Applications of Dielectric
Materials, Seoul, South Korea, 1997, pp. 726–729

[Matsuoka 2015] Matsuoka, N.; Komesu, D.; Mori, M.; Kozako, M.; Hikita, M.; Ishibe, S.: Electric field 3.3.4
grading of gas insulated switchgears with microvaristor composites. 2015 IEEE 11th
International Conference on the Properties and Applications of Dielectric Materials
(ICPADM)

[Matsusaki 2012] Matsusaki, H.; Nakano, T.; Ando, H.: Effects of Second Particles on Nonlinear Resistance 3.4.3
Properties of Microvaristor-Filled Composites. Conference on Electrical Insulation and
Dielectric Phenomena, CEIDP, Montreal, pp183-186, 14-17 October 2012

[Merck 2012] Merck Minatec® CM series, Conductive pigments product brochure. May 2012 3.3.4

[Merte 2007] Merte, R.: Measurements of electric fields with an electro-optic miniature probe. 4.4.1
International Symposium on High Voltage Engineering ISH, Ljubljana, Slovenia, 2007

[Metz 2007] Metz, R.; Morel, J.; Houabes, M.; Panisot, J.; Hassanzadeh, M.: High voltage 3.3.4
characterization of tin oxide varistors. J Mater Sci (2007) 42, pp. 10284-10287

[Mizuno 1999] Mizuno, Y; Naito, K.; Suzuki, Y.; Mori, S.; Nakashima, Y.; Akizuki, M.: Voltage and 3.2.1
Temperature Distribution along Semiconducting Glaze Insulator Strings. IEEE Transactions
on Dielectrics and Electrical Insulation, vol. 6, No. 1, February 1999

[Mori 1982] Mori, E.; Hoshi, M.: Development of EHV Transformers Based on UHV Technique. 5.3.4
HITACHI Review Vol 31 (1982) No 3

[Mosch 1979] Mosch, W.; Hauschild, W.: Hochspannungsisolierungen mit Schwefelhexafluorid. VEB 5.6.3
Verlag Technik, Berlin; Edition of Dr. Alfred Hüthig Verlag, Heidelberg, 1979. ISBN: 3-778-
50540-8.

[Moulson 2003] Electroceramics: Materials, Properties, Applications. 2nd Edition. Edited by A. J. Moulson 3.3.4
and J. M. Herbert. 2003 John Wiley & Sons, Ltd

[Mzenda 2002] Mzenda, V. M.; Goodman, S. A.; Auret, F. D: Conduction models in polyaniline – The effect 3.3.4
of temperature on the current-voltage properties of polyaniline over the temperature range
30 <T(K) >300. Synthetic Metals, vol. 127, pp. 285-289, 2002.

[Naito 1997] Naito, K.; Mizuno, Y.; Murasige, K.; Suzuki, Y.; Mori, S.; Nakashima, Y.; Akizuki, M.: 3.2.1
Voltage and temperature distribution along semiconducting glaze insulator string, NGK
Review: No. 21, Dec. 1997, Overseas Edition, NGK Insulators, Ltd, page 15-18

[Nexans 2010] Synthetic material end for a DC electric cable. Patent US7674979B2, March 9, 2010 5.1.3

[Öftering 2019] Öftering, H.-P.; Rumpelt, P.; Küchler, A.; Jenau, F.; Fritsche, R.: Time-dependent dielectric 5.3.3
behavior of mineral oil under the influence of different DC voltage conditions. 2019 IEEE
20th International Conference on Dielectric Liquids (ICDL)

[Ohki 2015] Ohki, Y.: Some R&D Activity on Polymer Composites in Japan. IEEE Electrical Insulation 5.6.3
Magazine, Vol.31, No.6, November/December 2015, pp.64-65

[Okamoto 2001] Okamoto, T.; Koyama, M.; Inoue, Y.; Takahashi, N.; Nakamura, S.: Percolation 3.3.4
phenomena of field grading materials made of two kinds of filler. Proc. Int. Symp.

125
TB 794 - Field grading in electrical insulation systems

Electrical Insulating Materials, Himeji, Jpn (ISEIM 2001). DOI:


10.1109/ISEIM.2001.973569

[Okubo 2012] Okubo, H.: Enhancement of Electrical Insulation Performance in Power Equipment Based 5.3.3
on Dielectric Material Properties. IEEE Trans. on Dielectrics and Electrical Insulation, Vol.
19, No. 3; June 2012

[Okubo 2015] Okubo, H.; et al.: Electrical Insulation Performance under HVDC and polarity reversal 5.3.3
conditions in oil-immersed DC converter transformers. The 19th ISH, Pilsen, Czech
Republic, 2015

[Okuyama 1984] Okuyama, K.; Hosi, M.; Mori, E.; Kasima, Y.; Kamata, Y.: Development of a Prototype UHV 5.3.4
Transformer and its Application to 500 kV Transformers. IEEE Transactions on Power
Apparatus and Systems, Vol. PAS-103, No. 9, September 1984, pp 2545- 2551

[Önneby 2001] Önneby, C.; Mårtensson, E.; Gäfvert, U.; Gustafsson, A.; Palmqvist, L.: Electrical Properties 3.3.4
of Field Grading Materials Influenced by the Silicon Carbide Grain Size. 7th International
Conference on Solid Dielectrics, Eindhoven, Netherlands, pp. 43-45, 2001

[Pang 2010] Pang, H.; Chen, T.; Zhang, G.; Zeng, B.; Li, Z.-M.: An electrically conducting 3.3.4
polymer/graphene composite with a very low percolation threshold. Mater. Lett. 64, pp.
2226–2229.

[Paupardin 2015] M. L. Paupardin, M. Mammeri, N. Lecourtier: 345 kV DC XLPE Extruded Cable Systems 5.1.3
Development. 9th International Conference on Insulated Power Cables, JICABLE 2015

[Pelto 2004] Pelto, J.; Paajanen, M.; Kannus, K.; Lathi, K.; Harju, P.: Nonlinear DC Voltage-Current 3.3.4
Characteristics of New Polymeric Composite Materials Based on Semiconductive Polyaniline
Emeraldine Base Filler. Proceedings of the 2004 International Conference on Solid
Dielectrics (ICSD 2004), Toulouse, France, 2004

[Peschke 1999] Peschke, E.; von Olshausen, R.; Cable Systems for High and Extra-High Voltage, 5.1.1
Development, Manufacture, Testing, Installation and Operation of Cables and Their
Accessories. Erlangen, Germany, Publicis MCV Verlag, 1999

[Ramakrishnan 1997] Ramakrishnan, S.: Conducting Polymers – From a Laboratory Curiosity to the Market Place. 3.3.4
Resonance, 1997

[Rhyner 1997] Rhyner, J.; Bou-Diab, M.: One-dimensional model for nonlinear stress control in cable 3.4.5
terminations. IEEE Transactions on Dielectrics and Electrical Insulation, Vol.4 1997, 785 4.2

[Roberts 1995] Roberts, A.: Stress Grading for High Voltage Motor and Generator Coils. IEEE Electrical 5.5.2
Insulation Magazine, July/August 1995, Vol. 11, No. 4, pp. 26-31

[Romasanta 2011] Romasanta, L.J.; et al.: Functionalized graphene sheets as effective high dielectric 3.1.2
constant fillers. Nanoscale Research Letters 2011, 6:508

[Rüger 2012] Rüger, R.: Halbleitende Metalloxid-Pigmente mit nichtlinearen elektrischen Eigenschaften. 2.2.4
RCC–Tagungsbericht 2012: Werkstoffe – Forschung und Entwicklung neuer Technologien 3.3.1
zur Anwendung in der elektrischen Energietechnik. Berlin, 3rd/4th May, 2012. www.rcc- 3.3.4
polymertechnik.de 5.6.3

[Salaeh 2012] Salaeh, S.; Nakason, C.: Influence of Modified Natural Rubber and Structure of Carbon 3.1.2
Black on Properties of Natural Rubber Compounds. POLYMER COMPOSITES - 2012

[Schmerling 2009] Schmerling, R.: Entwicklung und Aufbau einer Prüfeinrichtung zur rückwirkungsarmen 4.4.1
Messung der elektrischen Potenzialverteilung entlang der Oberfläche von
Endenglimmschutzanordnungen großer elektrischer Maschinen. Diploma Thesis, TU
Dortmund, 2009

[Schulte 2013] Schulte-Fischedick, J.; Lehretz, F.; Denndörfer, H.; Seifert, J.; Debus, J.; Hinrichsen, V.; 5.6.1
Ye, H.; Clemens, M.; Bornowski, M.: Electric Field Grading Using Insulators with
Microvaristor Filled Silicon Rubber. DOI: 10.1109/ICSD.2013.6619780, 11th IEEE
International Conference on Solid Dielectrics, Bologna, Italy, 30.6.-4.7.2013

[Secklehner 2013] Secklehner, M.; Tenzer, M.; Hinrichsen, V.; lmamovic, D.: Funktionell gefülltes Epoxidharz 2.2.4
für HGü-Anwendungen in kompakten lsoliersystemen. Dresden, 4. ETG-Fachtagung 3.3.1
"Grenzflächen in elektrischen lsoliersystemen“, 2013 3.3.4

[Secklehner 2015] Secklehner, M.; Hinrichsen,V.: Reduktion von Oberflächenladungen in 2.2.4


Gleichspannungsanwendungen mittels elektrisch nichtlinearer Metalloxid-Pigmente. RCC- 3.3.1
Fachtagung „Werkstoffe – Forschung und Entwicklung neuer Technologien zur Anwendung 3.3.4
in der elektrischen Energietechnik. Berlin, 20th/21st May, 2015

[Secklehner 2017a] Secklehner, M.; Hussain, R.; Hinrichsen, V.: Tailoring of new Field Grading Materials for 3.3.1
HVDC Systems. INSUCON 2017, Birmingham, 16th to 18th may, 2017 3.4.5

126
TB 794 - Field grading in electrical insulation systems

[Secklehner 2017b] Secklehner, M. ; Hering, M.; Hinrichsen, V.: Characterization of fillers for HVDC field 3.3.1
grading materials. CIGRÉ Colloquium, 30. September - 6. October 2017, Winnipeg 3.4.5

[Secklehner 2019] Secklehner, M.: Auslegung und Charakterisierung nichtlinearer Feldsteuermaterialien für 3.4.5
kompakte Gleichspannungsisoliersysteme (Design and Characterization of Nonlinear Field
Grading Materials for Compact DC Insulation Systems). Ph.D. Thesis, TU Darmstadt, 2019,
http://tuprints.ulb.tu-darmstadt.de/9074/

[Seen 2012] Seen, F.: Charakterisierung der nichtlinearen Widerstandeigenschaften von 3.3.1
Endenglimmschutzsystemen rotierender Hochspannungsmaschinen. Elektrotechnik & 3.3.3
Informationstechnik, Vol 129, 2012, 306-313

[Seifert 2008] Seifert, J.; Hinrichsen, V.; Debus, J.; Clemens, M.; Weida, D.; Hagemeister, M.: Einsatz 5.6.1
von Feldsteuermaterialien in Hochspannungsisolatoren – Potentiale und Risiken. ETG-
Workshop „Werkstoffe mit nichtlinearen dielektrischen Eigenschaften“, Stuttgart,
13.3.2008, ETG-Fachbericht 110

[Seifert 2011] Seifert, J.: Praxis-Erfahrungen mit feldgesteuerten Verbund-Isolatoren in der Produktion 5.6.1
und Anwendung. ETG-Workshop „Feldsteuernde Isoliersysteme“, Darmstadt, November
2011

[Senn 2012] Senn, F. : Charakterisierung der nichtlinearen Widerstandeigenschaften von 3.3.3


Endenglimmschutzsystemen rotierender Hochspannungsmaschinen. Elektrotechnik und
Informationstechnik, Vol 129, Heft 5 (2012) pp 306-313

[Sharifi 2010] Sharifi-Ghazvini, E.: Analysis of Electrical and Thermal Stresses in the Stress Relief System 4.4.1
of Inverter Fed Medium Voltage Induction Motors. PhD thesis University of Waterloo, ON, 5.5.3
Canada, 2010

[Short 1949] Short, H.D.: A Theoretical and Practical Approach to the Design of High-Voltage Cable 5.1.1
Joints. Transactions of the American Institute of Electrical Engineers (Volume: 68, Issue:
2, July 1949), pp. 1275 – 1283, Print ISSN: 0096-3860

[Showa 2008] Patent EP 2161805B1: Polymer Bushing Insulator and Cable Terminating Connection Part 5.1.1
Using the Polymer Bushing Insulator

[Sonerud 2012] Sonerud, B.; Josefsson, S.; Boyer, L.; Mirebeau, P.; Lefevre, S.; Castellon, J.; Notingher, 3.3.4
P.: Impulse and step voltage measurements on materials with nonlinear V-I characteristic.
Ann. Rep. Conf. IEEE Electrical Insulation and Dielectric Phenomena (CEIDP), Montreal,
QC, Canada, pp. 207-210

[Sonerud 2013] Sonerud, B.; Furuheim, K.-M.; Josefsson, S.; Pelto, J.; Ketonen, M.; Härkki, O.: Conduction 3.3.4
behavior of polyaniline/elastomer composites and the influence of carbon black addition.
Proceedings of the 23rd Nordic Insulation Symposium (NORD-IS 13), Trondheim, Norway,
2013

[Sorqvist 2009] Sorqvist, T.; Christen, T.; Jeroense, M.; Mondiet, V.; Papazyan, R.: HVDC Light Cable 5.1.3
Systems-Highlighting the accessories. 21st Nordic Insulation Symposium (Nordis 2009),
Gothenburg (Sweden), June 15-17, 2009.

[Staubach 2010] Staubach, C.; Kempen, S.; Pohlmann, F.: Calculation of Electric Feld Distribution and 4.3
Temperature Profile of End Corona Protection Systems on Large Rotating Machines by Use 5.5.1
of Fnite Element Model. International Symposium on Electrical Insulation ISEI, San Diego,
CA, USA, 2010

[Staubach 2011] Staubach. C.; Kempen, S.; Pohlmann, F.; Jenau, F.: Computer Aided Design of an End 5.5.3
Corona Protection System for Accelerated Voltage Endurance Testing at Increased Line
Frequency. Electrical Insulation Conference EIC, Annapolis, MD, USA, 2011

[Staubach 2012a] Staubach, C.; Jenau, F.: Comparison of Transient Time-domain and Harmonic Quasi-static 4.3
Solution of Electrical and Thermal Coupled Numerical Stress Grading Calculations for Large
Rotating Machines. International Symposium on Electrical Insulation ISEI, San Juan,
Puerto Rico, USA, 2012

[Staubach 2012b] Staubach, C.: Synthetisch basiertes Entwurfsverfahren zur elektro-thermischen 5.5.1
Optimierung der resistiv-kapazitiven Potentialsteuerung großer, rotierender Maschinen. 5.5.2
Kölner Wissenschaftsverlag, 2012, ISBN 3942720345

[Staubach 2013] Staubach, C.: Neue Generation einer resistiv-kapazitiven Potentialsteuerung ermöglicht 4.3
optimiertes Design des Endenglimmschutzsystems - Teil B: Auslegung und Qualifikation. 4. 5.5.1
ETG-Fachtagung Grenzflächen in elektrischen Isoliersystemen, 2013

[Staubach 2015] Staubach, C.; Merte, R.: Direct Electric field Strength Distribution Determination on 4.4.1
Electrical Apparatus by Means of an Electro-optical Miniature Field Sensor. International
Symposium on High Voltage Engineering ISH, Pilsen, Czech Republic, 2015

127
TB 794 - Field grading in electrical insulation systems

[Staubach 2018a] Staubach, C.; Hildinger, T.; Staubach, A.: Comprehensive electrical and thermal analysis of 4.3
the stress grading system of a large hydro generator. IEEE Electrical Insulation Magazine
Volume: 34, Issue: 1, January-February 2018

[Staubach 2018b] Staubach, C.; Hildinger, T.: Advanced Techniques for Electrical and Thermal Analysis of 4.4.1
the Stress Grading System using the Example of a 20 kV Large Hydro Generator, CIGRE
SCIENCE & ENGINEERING Journal, October 2018

[Stone 2009] Stone, G.C.; Sascic, M.; Dunn, D.; Culbert, I.: Recent problems experienced with motor 5.5.1
and generator windings. IEEE PCIC, 2009

[Streit 2011] P. Streit, K. Gallego, F. Gahungu: Dry termination for high voltage cable system. 8th 5.1.2
International Conference on Insulated Power Cables, JICABLE 2011

[Strobl 2001] Strobl, R.; Haverkamp, W.; Malin, G.; Fitzgerald, F.: Evolution of stress control systems in 5.1.1
medium voltage cable accessories. Transmission and Distribution Conference and
Exposition, 2001 IEEE/PES , vol.2, no., pp.843-848 vol.2, 2001

[Tenzer 2011] Tenzer, M.; Hinrichsen, V.: Investigations on Microvaristors as Functional Fillers for HVDC 2.2.4
Applications. XVII International Symposium on High Voltage Engineering, Hannover, 3.3.1
Germany, August 22-26, 2011, paper E-072, ISBN 978-3-8007-3364-4

[Tenzer 2013a] Tenzer, M.; Secklehner, M.; Hinrichsen, V.: Short and long term behavior of functionally 5.6.3
filled polymeric insulating materials for HVDC insulators in compact gas insulated systems.
23rd Nordic Insulation Symposium, June 9-12 2013, Trondheim, Norway, ISBN 978-82-
321-0274-7

[Tenzer 2013b] Tenzer, M.; Winter, A.; Hinrichsen, V.; Kindersberger, J.; Imamovic, D.: Compact Gas-Solid 5.6.3
Insulating Systems for High-Field-Stress in HVDC Applications. CIGRE Study Committee B3
& Study Committee D1 Colloquium, 9-11 September 2013, Brisbane, Australia

[Tenzer 2013c] Tenzer, M.; Secklehner, M.; Hinrichsen, V.; Imamovic, D.: Funktionell gefülltes Epoxidharz 2.2.4
für HGÜ-Anwendungen in kompakten Isoliersystemen. 4. ETG-Fachtagung "Grenzflächen 3.3.1
in elektrischen Isoliersystemen", 12.-13. November 2013, Dresden, ISBN 978-3-8007-
3557-0

[Tenzer 2015] Tenzer, M.: Funktionell gefüllte Isolierwerkstoffe für Hochfeld-Gleichspannungs- 2.2.4
Isoliersysteme in kompakten gasisolierten Anlagen. Doctoral Thesis, TU Darmstadt, 2015, 3.3.1
ISBN 978-3-8439-2068-1, tuprints.ulb.tu-darmstadt.de/4431/ 5.6.3

[Thienpont 1964] Thienpont, J.; Sie, T. H.: Suppression of Surface Discharges in the Stator Windings of High 2.2.3
Voltage Machines. CIGRE Session 1964 4.1
4.2.
4.4.2
[Toepler 1921] Toepler, M.: Über die physikalischen Grundgesetze der in der Isolatorentechnik 2.2.3
auftretenden elektrischen Gleiterscheinungen. Archiv für Elektrotechnik 10(1921)5/6

[Ullrich 2001] Ullrich, H.; Gubanski, S.M.; Wiener, G.: Natural ageing of semiconducting glazed 5.6.1
insulators. XII International Symposium on High Voltage Engineering, 20.-24. August
2001, Bangalore, India, paper 5-33, pp. 715-718

[Ullrich 2002] Ullrich, H.: Aging and characterisation of semiconducting glazes. Lic. Thesis, Chalmers 5.6.1
Univ. Technol., Göteborg, Sweden, 2002

[Ulrich 2000] Ulrich, R.; Schaper, L.: Comparison of Paraelectric and Ferroelectric Materials for 3.1.2
Applications as Dielectrics in Thin Film Integrated Capacitors. The International Journal of
Microcircuits and Electronic Packaging, Volume 23, Number 2, Second Quarter 2000 (ISSN
1063-1674)

[Vojta 1996] Vojta, A.; Clarke, D.R.: Microstructural origin of current localization and ‘‘puncture’’ failure 3.4.5
in varistor ceramics. Journal of Applied Physics 81, 985 (1997);
https://doi.org/10.1063/1.364226

[Wang 2006] Wang, X.; Herth, S.; Hugener, T.; Siegel, R.W.; Nelson, J.K.; Schadler, L.S.; Hillborg, H.; 5.1.3
Auletta, T.: Nonlinear Electrical Behavior of Treated ZnO-EPDM Nanocomposites. 2006
Annual Report Conference on electrical Insulation and Dielectric Phenomena, Kansas City,
US, pp. 421-424, 2006

[Wang 2012] Wang, Z.; Nelson, J. K.; Miao, J.; Linhardt, R. J.; Schadler, L.S.; Hillborg, H.; Zhao, S.: 3.1.2
Effect of High Aspect Ratio Filler on Dielectric Properties of Polymer Composites: A Study
on Barium Titanate Fibers and Graphene Platelets. IEEE Transactions on Dielectrics and
Electrical Insulation, Vol. 19, No. 3; June 2012

[Wang 2012a] Wang, Z.; Nelson, J. K.; Hillborg, H.; Zhao, S.; Schadler, L.S.: Nonlinear Conductivity and 3.3.4
Dielectric Response of Graphene Oxide Filled Silicone Rubber Nanocomposites. Ann. Rep.
Conf. IEEE Electrical Insulation and Dielectric Phenomena (CEIDP), Montreal, QC, Canada,
pp. 40-43

128
TB 794 - Field grading in electrical insulation systems

[Wang 2012b] Wang, Z.; Nelson, J. K.; Hillborg, H.; Zhao, S.; Schadler, L.S.: Graphene Oxide Filled 3.3.4
Nanocomposite with Novel Electrical and Dielectric Properties. Adv. Mater. 24, pp. 3134–
3137.

[Wang 2012c] Wang Z.; et al.: Effect of High Aspect Ratio Filler on Dielectric Properties of Polymer 3.3.4
Composites: A Study on Barium Titanate Fibers and Graphene Platelets. IEEE Trans.
Dielectr. Electr. Insul. 19, pp. 960-967.

[Warfield 1961] Warfield, R. W.; Petree, M. C.: Electrical Resistivity of Polymers, SPE Transactions, April 3.1.1
1961

[Weida 2008] Weida, D.; Steinmetz, T.; Clemens, M.: Electro-Quasistatic High-Voltage Field Simulations 4.3
of Insulator Structures Covered with Thin resistive Pollution or Nonlinear Grading Material.
Proceedings of the 2008 IEEE International Power Modulators and High Voltage
Conference, Las Vegas, US, pp. 580-583, 2008

[Weida 2009] Weida, D.; Steinmetz, T.; Clemens, M.: Electro-Quasistatic High Voltage Field Simulations 4.3
of Large Scale Insulator Structures Including 2D Models for Nonlinear Field-Grading
Material Layers. IEEE Trans. on Magn., Vol. 45, No. 3, pp. 980-983, March 2009

[Weida 2011] Weida, D.: Optimierung von Applikationen aus der Hochspannungstechnik mit dünnen 4.3
Schichten aus mikrovaristorgefüllten Polymeren mithilfe von nichtlinearen transienten 3D
Simulationen. Dr.-Ing. Thesis, 2011

[Weidner 2011] Weidner, J. R.; Staubach, C.; Schmidt, G.: Elektrische Feldsteuerung bei 5.5.1
Ständerwicklungen von Turbogeneratoren – Anforderungen, Berechnungsverfahren,
Konstruktion, Betriebsverhalten. ETG-Workshop „Feldsteuernde Isoliersysteme“,
Darmstadt, November 2011

[Werdelmann 2009] Werdelmann, P.; Pohlmann, F.; Kempen, S.: Partial Discharge Inception on Tangentially 2.2.3
Stressed Boundary Surfaces for Insulation Materials of Large Turbine Generators. 4.4.2
International Symposium on High Voltage Engineering ISH, Capetown, South Africa, 2009

[Wheeler 2005] Wheeler, J.C.G.: Effects of converter pulses on the electrical insulation in low and medium 2.2.3
voltage motors. IEEE Electrical Insulation Magazine, Volume 21, Issue 2, March-April 2005 4.4.2
5.5.3
[Winter 2011] Winter, A.; Kindersberger, J.; Tenzer, M.; Hinrichsen, V.: Leitfähig eingestellte 3.3.1
Isoliersysteme für Hochfeldbeanspruchung in gasisolierten Anlagen bei 5.6.3
Gleichspannungsbelastung. ETG-Workshop „Feldsteuernde Isoliersysteme“, Darmstadt,
November 2011

[Winter 2014] Winter, A.; Kindersberger, J.; Tenzer, M.; Hinrichsen, V.; Zavattoni, L.; Lesaint, O.; Muhr, 2.2.4
M.; Imamovic, D.: Solid/Gaseous Insulation Systems for Compact HVDC Solutions. Cigré 3.3.1
Session 2014, Report D1_102_2014 5.6.3

[Winter 2015] Winter, A.; Kindersberger, J.; Tenzer, M.; Hinrichsen, V.; Zavattoni, L.; Lesaint, O.; Muhr, 2.2.4
M.; Imamovic, D.: Solid/Gaseous Insulation Systems for Compact HVDC Solutions. Cigre 3.3.1
Science & Engineering, Vol. 1, pp. 123-134, February 2015

[Wirth 2015] Wirth, I.; Reumann, A.; Zink, M.H.; Küchler, A.; Schnitzler, T.; Langens, A.; Berger, F.: 2.2.2
Steady-State and Transient Electrical Potential Distributions in HVDC Bushings Measured 2.2.4
Under Different Thermal Conditions 5.4.1

[Worzck 2013] Worzck, T.: NordBalt-the ultra-long 300kV extruded HVDC cable 280 miles through the 5.1.3
Baltic Sea. ICC Spring Meeting, TNL, April 30, 2013

[Wu 2003] Wu, Y.; Zhao, X.; Li,F.; Fan, Z.: Evaluation of Mixing Rules for Dielectric Constants of 3.1.2
Composite Dielectrics by MC-FEM Calculation on 3D Cubic Lattice. Journal of 3.3.2
Electroceramics, 11, 227–239, 2003

[Wu 2016] Wu, K.; Cheng, C.: Interface Charge Behavior in the Composite Insulation of Oil Immersed 5.3.3
Paper. IEEE Int. Conf. High Volt. Eng. (ICHVE), 2016

[Yang 2015] Yang, X.; He, J.; Hu, J.: Tailoring the nonlinear conducting behavior of silicone composites 3.3.4
by ZnO microvaristor fillers. J. APPL. POLYM. SCI. 2015, DOl: 10.1002/APP. 42645

[Yang 2018] Yang, X.; Zhang, X.; Hu, J.; He, J.: Grading Electric Field in High Voltage Insulation Using 4.3
Composite Materials, EEE Electrical Insulation Magazine Volume: 34, Issue: 1, January-
February 2018

[Ye 2014] Ye, H.; Clemens, M; Schulte-Fischedick, J.; Seifert, J.: Investigation of Electric field 5.6.2
Grading of Bushings with Microvaristor Filled Epoxy Resin. 2014 IEEE International Power
Modulator and High Voltage Conference (2014 IPMHVC), Santa Fe, USA. 2014.

[Ye 2015] Ye, H.: Application of Nonlinear Microvaristor-Filled Materials in High-Voltage Devices and 5.6.2
Algorithmic Optimization of High-Voltage Simulations Based on Surrogate Models. Doctoral
Thesis Universität Wuppertal, 2015, urn:nbn:de:hbz:468-20150511-115707-6 [http://nbn-

129
TB 794 - Field grading in electrical insulation systems

resolving.de/urn/resolver.pl?urn=urn%3Anbn%3Ade%3Ahbz%3A468-20150511-115707-
6]

[Zhao 2007] Zhao, G.; Joshi, R.P.; Lakdawala, V.K.: Electro-thermal Simulation Studies for Pulsed 3.4.5
Voltage Induced Energy Absorption and Potential Failure in Microstructured ZnO Varistors.
IEEE Transactions on Dielectrics and Electrical Insulation Vol. 14, No. 4; August 2007

[Zink 2013] Zink, M.: Zustandsbewertung betriebsgealterter 5.4.1


Hochspannungstransformatordurchführungen mit Öl-Papier-Dielektrikum mittels
dielektrischer Diagnose. Dissertation TU Ilmenau, 2013

[Ziomek 2011] Ziomek, W.; et al.: High Voltage Power Transformer Insulation Design. Electrical 5.3.1
Insulation Conference, Annapolis, Maryland, 5 to 8 June 2011, pp 211-215

[Zulkifli 2012] Zulkifli, A.: Polymeric Dielectric Materials. http://dx.doi.org/10.5772/50638 3.1.2

130
CIGRE
21, rue d'Artois
75008 Paris - FRANCE

© CIGRE

ISBN : 978-2-85873-496-2

You might also like