Sei sulla pagina 1di 15

F

Dielectric Spectroscopy in Time and Frequency Domain for HV Power Equipment, Part I: Theoretical Considerations
Key Words: Dielectrics, polarization effects, dielectric response function, dielectric measurements in time and/or frequency domain.

oday, the catchword in utility companies is condition-based or predictive testing and maintenance, or even life management, as opposed to time-based or preventive maintenance only. The reasons for this evolution are well known: In North America, for instance, their roots can be found in the restructuring and de-regulation of the electric utility industry. Here in Europe, it was the adoption of the Single Market Directive which became the norm in February 1999, whereas it was still the exception many years before [1]. The main issue of this de-regulation of the electricity market, now subdivided into independent power producers, transmission companies, system operators and distribution companies, is to encourage competition while still maintaining basic public policy and service objectives. All partners within this new scheme are thus forced to cut costs in maintenance and operation without endangering a steady supply of electricity to demanding customers. Costs can be reduced, first of all, by a transition from time-based maintenance (TBM) to condition based maintenance (CBM), if the actual conditions of the expensive high voltage components within the electric power transmission systems are reliably known. The application of unscheduled maintenance, a philosophy based on a reactive mode of operation, will not reduce costs. Unscheduled maintenance means that repairs and maintenance will be made only when the equipment, e.g. a transformer or cable, breaks down. But this generally causes a downtime of the electricity supply, so that breakdowns become much more costly than planned maintenance. A CBM based on reliable diagnostic tools should thus be applied today.

Walter S. Zaengl Swiss Federal Institute of Technology (ETH), Zurich, Switzerland

Most dangerous breakdowns are caused by the aging effects of HV insulation systems used within [HV] components, and there is still a lack of appropriate tools to diagnose such systems non-destructively and reliably in the field.
The driving force for the development and application of improved diagnostic methods is the steadily increasing age of expensive HV components. In many parts of the world, the majority of large power transformers were installed in the 1960s and 70s. Also, cable technology changed at this time, and the first generations of PE or XLPE cables are still prone to breakdowns. All these facts are very well known. It is not the aim of this article to discuss the full complexity of all the existing diagnostic techniques that have been successfully applied to different HV components to support CBM. Most of the dangerous breakdowns, however, are caused by the aging effects of HV insulation systems used within these components, and there is still a lack of appropriate tools to diagnose such systems non-destructively and reliably in the field. New methods have been published in the last decade and even before, 5

September/October 2003 Vol. 19, No. 5

0883-7554/03/$17.002003IEEE

for which reliable diagnostics are claimed. Some of these methods are based on changes of the dielectric properties of the insulation. Dielectric properties are dependent on many factors, e.g. on frequency or time, on temperature and chemical composition of an individual dielectric, or on the structure of an insulation system composed of different dielectrics. In electric power engineering, most of these factors, e.g. dielectric dissipation or power factor, are considered during standardized tests for many power apparatus, but the frequency of the test voltage is, in general, only the power frequency for which the equipment is designed. The power factor at a single frequency is, however, sometimes insufficient to qualify even strong changes in the dielectric properties in complex systems. A critical review of publications related to alternative diagnostic methods that are based on different dielectric response methods, as well as discussions with users of such methods, disclose a lack of knowledge of the fundamentals of such alternative methods. Part I of this article introduces the fundamentals of dielectric spectroscopy in the time, as well as in the frequency, domain, and provides a short overview of more recently applied methods which are able to quantify dielectric properties in the laboratory as well as in the field. This contribution will extend the short introduction into dielectric response measurements recently published in EI by Barry H. Ward in his feature article on power transformers [2]. Part II will describe some specific applications of dielectric spectroscopy for HV power equipment, and it will show why some interpretations of the methods will not comply with assumed expectations.

Background of Dielectric Response Every kind of insulation material consists, at an atomic level, of negative and positive charges balancing each other on the microscopic as well as on more macroscopic scales (if no unipolar charge was deposited within the material before by well known charging effects). Macroscopically, some localized bipolar space charge may be present, but even then, an overall charge neutrality exists. As soon as a material is exposed to an electric field (generated by a voltage across electrodes embedded in the insulation), the positive and negative charges become oriented thus forming different kinds of dipoles even on atomic scales. A local charge imbalance is thus induced within the neutral species (atoms or molecules) as the centers of gravity for the equal amounts of positive and negative charges, q, become separated by a small distance d, thus creating a dipole with a dipole moment, p = qd, which is related to the local or microscopic electric field E acting in close vicinity to the species. Thus, the dipole moment can also be written as p = E, where is the polarizability of the species or material under consideration. Note that p, d and E are vectors not marked here. As the distance d will be different for different species as well as their number of dipoles per unit volume,

their polarizability is also different. Due to chemical interactions between dissimilar atoms forming molecules, many molecules will have a stable distance d between the charge centers, thus forming permanent dipoles, which are usually randomly oriented and distributed within the material, as long as no external field is applied. (Note that any kind of permanent polarization, for example, in electrets or ferroelectrics, is not considered here). The macroscopic effect of the polarizability of individual species can be given in a general relation between the macroscopic polarization P and the number of polarized species per unit volume of the material. These relationships are well known, but not treated here. But let us now briefly recall the main mechanisms that produce macroscopic polarization P: G Electronic Polarization is effective in every atom or molecule as the center of gravity of the electrons surrounding the positive atomic cores will be displaced by the electric field E. This effect is extremely fast and thus effective up to optical frequencies. G Ionic (or atomic/molecular) Polarization refers to material containing molecules forming ions that are not separated by low electric fields or low working temperatures. Apart from electronic polarization induced in such molecules by an electric field, elastic displacements of charges (nuclei and electrons) will also occur; i.e. these types of molecules are polar substances, which can be polarized up to infra-red frequencies. G Dipolar (or orientational) Polarization refers to materials containing molecules with permanent dipole moments with orientations statistically distributed due to the action of thermal energy. Under the influence of E, the dipoles will be oriented only partially, so again, a linear dependency of P and E exists. Ionic and dipolar polarization are still quite fast effects and may follow ac frequencies up to MHz or GHz. G Interfacial Polarization is predominantly effective in insulating materials composed of different dielectric materials, such as e.g. oil impregnated paper/cellulose. The mismatch of the product of permittivity times resistivity for the different dielectrics causes, under the influence of an electric field, movable positive and negative charges to become deposited on the (micro- or macro-) interfaces of the different materials, also forming some kinds of dipoles. This phenomena is often very slow and in general active in the power frequency range and below. G Trapping and hopping of charge carriers between localized charge sites may occur also creating polarization. This is also a slow, but strongly temperature dependent, process found mostly in solids, for example in cable materials (PE etc.). For more details the reader may either refer to one of the standard text books, e.g. [3], or to some more generalized reflections on this subject, e.g. [4][7].

IEEE Electrical Insulation Magazine

In summary, the dielectric polarization is the result of a relative shift of positive and negative charges in a material. During all of these processes, the electric field is not able to force the charges to escape from the material, which would cause inherent electric conduction. Measurement methods to perform dielectric spectroscopy (Dielectric Response Methods) are based on fundamental interactions between well known electric quantities: Usually, HV insulation materials (also called dielectrics) are isotropic and usually homogeneous, at least at macroscopic scales. Then, the vectors of the macroscopic polarization P and the electric field E are of equal direction and related by P= 0 E. (1)

sources and sinksof all electric field lines. For time-varying fields, E(t), the so called displacement current must be supplied by the voltage source to maintain the area charge density at the electrodes. This current is governed by dQ/dt, where Q is the total electric charge deposited on each of the electrodes. If the vacuum is now replaced by any kind of isotropic dielectric material, the electric displacement D of eq. (2) increases by its inherent (macroscopic) polarization P as defined in eq. (1): D(t) = 0 E(t) + P(t) =
0

(1 + )E(t).

(3)

Here, is the electric susceptibility of the material, a dimensionless, pure number which is zero for ideal vacuum. Thus, the susceptibility accounts for all kinds of polarization processes within a dielectric. 0 is the permittivity of free space or vacuum (8.85419 10-12 [As/Vm]), a number with units relating the unit for electric field [V/m] to that of electric displacement [As/m2]. This already provides a hint that all polarization processes will induce electric charges at the electrodes, as soon as a voltage is applied. From eq. (1), it follows that the polarization P will change or vanish if the field E is changed or set to zero. In any dielectric ( > 0), a reduction in E will thus lead to a depolarization (or relaxation) process, which will follow with some delay or retardation to the reduction of E. Dielectric properties thus become dynamic events that can be quantified in the time- as well as in the frequency domain.

Dielectric Response/Spectroscopy in Time Domain (TD) In a vacuum-insulated electrode arrangement, the vector of electric displacement (or dielectric flux density or electrical induction) D is exactly proportional to the electric field vector E,
D=
0

The significance of this equation is first of all related to the fact that it separates the two kinds of charge induction. (As both vectors, P and D, are still in parallel to E for isotropic materials, we will from now on in this article avoid the use of bold letters to denote vectors). Of no minor importance is the fact that the time dependency of P(t) will no longer be the same as that of E(t), as the different polarization processes have different time delays with respect to the appearance of E. This delay is obviously caused by the time-dependent behavior of the susceptibility = (t). This time delay may best be understood in the following way: Let us assume that a step-like electric field of magnitude E0 is applied to the dielectric at any time t0 and that this field remains constant for t t0. The dielectric can then be characterized by its time dependent susceptibility (t) as well as by its specific polarization P(t) as a response in the time domain, i.e. the formation and evolution of the different kinds of polarization processes, which develop within extremely short times (as e.g. electronic polarization) as well as in longer (e.g. dipolar polarization) or even much longer (e.g. interfacial polarization) time spans. For t t0, the magnitude of the susceptibility or polarization is zero. Figure 1 shows the development of polarization processes in the time domain, which, according to eq. (1), can be expressed as P(t)/E0 =
0

(2)

(t) 1(t).

(4)

or, if the electric field is generated by a time-varying voltage, D(t) =


0

E(t).

(2a)

Here, 0 is again the permittivity of vacuum. The origin of D and E is usually provided by a voltage source connected to the electrodes under consideration. No time delay will exist between both magnitudes, if the time scales considered still produce electrostatic field conditions. Note, however, that D represents the positive and negative electric charges per unit area induced at the electrode surface, and that these charges are the origin
September/October 2003 Vol. 19, No. 5

Here (t) as well as P(t) represent step response functions. The factor 1(t) is used to indicate the unit step for the electric field E0. In Figure 1 the first part of this function is simplified by an ideal step to account for the very fast polarization processes, an instantaneous polarization, P(t = t0) = P, which includes not only electronic but also other very fast polarization processes. (The index of P is thus related to the frequency domain). This step, at least for large HV power equipment, can be recorded neither in the time- nor in an equivalent frequency domain. As all polarization processes are finite in magnitude and will settle at long times, the polarization finally becomes static, P(t ) = PS. 7

According to Figure 1, the step response of this somewhat simplified polarization can be written as P(t) = P + (PS P) g(t t0), (5)

written as a sum of conduction, vacuum and polarization displacement currents, i.e.: j( t ) = 0 E( t ) + D( t ) t E( t ) P( t ) = 0 E( t ) + 0 + t t

where g(t) is a dimensionless, monotonously increasing function. Therefore, eq. (4) may also be written as P(t) =
0

(8)

+(

) g(t

t0)] E0

(5a)

and with eq. (6) for E(t) = const. j( t ) = 0 E( t ) + 0 [ ( t ) + f ( t )]E( t ) (8a)

or, if the relative permittivity, = 1 + , is introduced: P(t) =


0

[(

1) + (

) g(t

t0)] E0.

(5b)

As known from general circuit theory, it is now possible to compute any other time dependent polarization P(t) for any other time dependent excitation E(t) of a test object, as the special solutions for the step excitation are already known. This can be done by the use of Duhamels Integral or convolution in the time domain. For the quantities in eq. (5a), the result is: P( t ) = 0 E( t ) + 0

f (t )E( )d,

with = 1 + . Equation (8a) and eq. (9) are a first basis for the measurement of the dielectric response function f(t) or for characterizing dielectric materials with the time-domain (TD) method [5]. To do so, a step dc charging voltage of magnitude UC, which must be constant and free of ripple, is suddenly applied to the test object which has been previously carefully discharged. Then the polarization (or absorption, or charging) current ipol(t) through the test object can be recorded according to i pol ( t ) = C 0 U c 0 + ( t ) + f ( t ) 0

(6)

where f(t) is the so called dielectric response function: f(t) = (


S

(9)

) g(t)/t = (

) g(t)/t.

(7)

f(t) defined by eq. (7) is obviously a monotonously decreasing function and inherent to the dielectric being investigated. The polarization P(t) produces the main part of the polarization (or absorption, or charging) current in a test object if the electric field, E(t), is suddenly applied. So far, we have not yet considered any pure dc conductivity 0, which represents the movement of the free charges in the dielectric and which is not involved in polarization. As already postulated by Maxwell in 1891 [8], the field E(t) generates a total current density j(t), which can be

where C0 is the geometric capacitance of the test object, and (t) is the delta function arising from the suddenly applied step voltage at t = t0. (NOTE: The dimension of f(t) is 1/s and its magnitude is tied to C0, the geometric capacitance, which may either be the vacuum capacitance of the electrode system between which the dielectric is sandwiched, or the high frequency capacitance of the dielectric at time t0 at which the current measurement was started.) The transition from eq. (8a) to (9) is easy to perform. The charging current contains three terms: The first one is related to the intrinsic conductivity of the test object and is independent of any polarization process, the last one represents all the activate polarization processes during the voltage application and the middle part with the delta function cannot be recorded in practice due to the large dynamic range of current amplitudes inherent with the very fast polarization processes. A polarization current measurement can usually be stopped if the current becomes either stable due to the dc term or becomes very low. If the test object is now short-circuited at t = tc, the depolarization (or discharging, de-sorption) current idepol can be measured, see Figure 2. According to the superposition principlethe sudden reduction of the voltage UC to zero is regarded as a negative voltage step at time t = tc- and neglecting the second term in eq. (9) which is again a very short current pulsewe get for t (t0 + TC): i depol ( t ) = C 0 U C [ f ( t ) f ( t + T C )]. (10)

Figure 1. Polarization of a dielectric exposed to a step field of magnitude E 0 at t = t 0 .

IEEE Electrical Insulation Magazine

Tc is the time duration when the step voltage was applied. This current is of opposite polarity. The second term in this equation can only be neglected if the charging period Tc was long enough to complete all polarization processes. The depolarization current then becomes directly proportional to the dielectric response function f(t), as the dc conductivity 0 of the dielectric is not involved, but can easily be calculated from the difference between the polarization and depolarization currents. Equation (10) is thus a second basis for the measurement of the dielectric response function f(t) or for characterizing dielectric materials in the time-domain (TD). Note, however, that insufficient charging periods Tc will not force the second term to become zero, i.e. there is a memory effect in the dielectric due to polarization phenomena which have not been completed due to the insufficient charging period Tc. (This effect is demonstrated later in Figure 5). In Figure 2, the sharp current peaks associated with the delta functions (t) in eqs. (9) and (10) are not included for obvious reasons. In practice, the polarization and depolarization currents are conveniently measured with a two electrode technique as illustrated in Figure 3. The two resistors of this circuit represent small protection resistances, which will not influence the recorded currents. If the test object contains an insulation system, which can be subdivided into different sub-systems, the polarization and depolarization currents can then be defined by the selected electrode arrangement and conveniently be sensed at virtual ground potential. The complex insulation system of power transformer test samples embedded in guard ring electrodes are typical examples for such an application.

trical Technology, Switzerland). The test cell was a cylindrical glass vessel in which the electrode arrangement (parallel plate electrodes with a guard ring on the ground electrode, diameter of guarded section 113 mm) was placed. The currents were monitored between the guarded electrode and ground. The pressboard samples (circular disks of 160 mm diameter and 2 mm thick) were carefully dried under vacuum (<1 mbar, 24 hours at 105 C) before impregnation, then exposed to ambient air to reach the desired moisture level (as quantified by the increase of weight), and storedhermetically sealedfor one week before impregnation. This latter procedure is necessary to make the moisture distribution within the samples more uniform. Finally, impregnation with degassed and dried Technol US 3000 mineral oil (moisture content < 5 ppm) was performed. The same oil was used within the test cell. As the results are somewhat dependent on the pressure applied to the samples, a pressure of about 0.9 N/cm2 was always used. More information can be found within the original publication [9]. The computer-controlled instrumentation used for the measurements consisted of a Keithley 617 electrometer, a very well stabilized dc power supply and HV relays for performing the switching events. Figure 4 shows the results for unaged samples only. All measurements started 1 s after step voltage application (ipol) and after short circuit (idepol). From the results it can be seen

Performing Time Domain (TD) Measurements Combined polarization and depolarization current measurements covering some decades in TD are indicated in the literature as PDC measurements. As long as both kinds of currents are measured during time periods starting at not less than about 0.1 to 1 s after the switching events (t0 and tc respectively in Figure 2), and if measurement periods up to hours are made, the measurements can easily be performed by means of standard equipment. As too high electric fields within dielectric materials may cause non-linear effects, the dc power supply indicated in Figure 3 need only be designed for voltages up to about 1000 V . However, the power supply must be free of significant ripple and provide a constant and very stable voltage about 1 ms after each switching event. Switching events can be performed by means of relays or electronics. Electrometers should be able to record currents down to some pA or less. Figure 4 shows typical results for such a PDC measurement procedure, carried out on oil-impregnated pressboard samples with different moisture content (m.c.) in the pressboard (Type T IV of Weidmann ElecSeptember/October 2003 Vol. 19, No. 5

Figure 2. Principle of polarization and depolarization current (PDC) measurements.

Figure 3. Principle of test arrangement for the PDC measuring technique.

that stable polarization currents could be reached only for the higher moisture content after about 200,000 s, i.e. about 56 hours. The slow and more or less continuous decay of the depolarization currents confirms that the material is highly dispersive, i.e. that quite a large distribution of relaxation effects (interfacial polarization due to the tissue structure of the pressboard) exists. Representation of such results on log-log-scale is paramount due to the large dynamic range of the quantities. According to eq. (10) it can be assumed that all four depolarization currents of Figure 4 represent the more or less perfect dielectric response functions f(t) of the pressboard samples with different moisture contents, as the charging time TC was very long. But this may not be true, as even for this very long charging period not all polarization processes may not have been completed. This effect is well known and is called the memory effect, which can also be explained by the equivalent circuit, Figure 9, discussed later. This effect can clearly be seen by the experimental results shown in Figure 5 for the very dry sample from Figure 4 (m.c. 0.2 %). Between each mea-

surement, the sample was carefully discharged for a long time, so that every new measurement cycle started with the same virgin conditions. This can be easily recognized by the polarization currents, which follow the same curve. The depolarization currents, however, become smaller and smaller for every reduction in the charging time. A rule of thumb for the measurements is that the dielectric response functions f(t) or the depolarization current become quite accurate if the charging time is about 5 to 10 times longer than the depolarization current is quantified. Measurements of depolarization currents on oil-impregnated Kraft paper for limited time durations (0.01 to 0.10 s) were performed nearly fifty ago, see [10]. The magnitudes of the currents and the slopes of their decay were dependent on the moisture content, which varied between 0.5 to 7.5 %. Such transient measurements were made with an oscilloscope. The example in Figure 4 as well as the measurements published in [10] demonstrate the complexity of the time dependence of the polarization (charging) and depolarization (discharging) currents of a typical power transformer dielectric. Dependencies of this kind are quite different from the time-domain behavior for several solids discussed in Jonschers publications [5]-[7]. Nevertheless, certain time spans may be simulated by idealized responses [11]; but quite long periods as measured and quantified here should always take the specific response function for the investigated material into account.

Figure 4. Polarization and depolarization currents of unaged oil-impregnated pressboard samples with different moisture content (m.c.).

Dielectric Response/Spectroscopy in Frequency Domain (FD) An analytical transition from time to frequency domain can be made using the Laplace- or Fourier transform by rewriting eqs. (6) and (8). An ideal step response for the total current density of a dielectric response function f(t), considering also instantaneous polarization processes are assumed:
j( t ) = 0 E( t ) + 0 with j(t) j(p); E(t) E(p); E(t) p E(p); f(t) F(p); and considering the convolution of the last term in this equation we get, with p being the Laplace Operator: j( p) = 0 E( p) + 0 pE( p) + 0 pF( p)E( p). (12) dE( t ) d + 0 f ( t )E( )d dt dt 0
t

(11)

As p is the complex frequency i , we can reduce the equation to


Figure 5. Polarization and depolarization currents of a dry pressboard sample as a function of charging duration T C .

j( ) = E()[ 0 + i 0 ( I + F()].

(13)

10

IEEE Electrical Insulation Magazine

Thus F( ) is the Fourier Transform of the dielectric response function f(t) or the complex susceptibility : ( ) = F( ) = ( ) i ( ) = f ( t )exp( it )dt.
0

(14)

Note that the frequency scale is now 0 . Combining eqs. (13) and (14) shows the total current density: j( ) = { 0 + 0 ( ) + i 0 [1 + ( )]}E( ). (15)

The main part of this current has its origin in the complex electric displacement D( ) which is proportional to the complex dielectric permittivity, (), with the relation: D( ) = 0 ( )E( ) where: ( ) = ( ) i ( ) = (1 + ( )) i ( ). (17) (16)

made at only one frequency. As aging effects will change these quantities in quite different and specific frequency ranges, new diagnostic tools will monitor and detect this effect. Equation (14) disclosed the coherency between the time and frequency domains. Thus it is obvious that the complex susceptibility ( ) and its real and imaginary parts can be converted to the dielectric response function f(t) and vice versa [13]. Both domains extend from zero to infinity, but in practice for every conversion, only the measurement results will be available. Possible conversion procedures are not discussed here in detail, but some information is provided at the end of this article. Finally, it should be noted that all dielectric quantities are more or less dependent on temperature. Any comparison or measurement of these quantities must take this into account.

Actual measurements of this dielectric response in the frequency domain are difficult to perform, if the frequency range becomes very large. Usually and at least in electric power engineering, only a single C - tan measurement is made, i.e., at power frequency. Sophisticated laboratory instruments can cover, however, even many decades in frequency [12]. Note, that according to eq. (15) such instruments cannot distinguish between the current contribution of the pure dc conductivity 0 and that of the dielectric loss (). This means that the measured relative dielectric permittivity~ r ( ) is different from the relative permittivity () defined in eqs. (16) and (17). Then the measured relative dielectric permittivity ~ r ( ) is defined from the following relation: j( ) = i 0 ~ r ( )E( ). Therefore: ~ r ( ) = r ( ) i r ( ) = ( ) i[ ( ) + 0 / 0 ] = 1 + ( ) i[ ( ) + 0 / 0 ] and the dielectric dissipation factor, tan ( ), tan ( ) = ( ) + 0 / 0 r ( ) = r ( ) r ( ) (18)

(19)

(20)

The real part of eq. (19) represents the capacitance of a test object, whereas the imaginary part represents the losses. Both quantities depend on frequency. Often, this fact is not appreciated, if a C - tan measurement is
September/October 2003 Vol. 19, No. 5

Performing Frequency Domain (FD) Measurements Typical results of FD measurements are shown in Figures 6 and 7. The example is taken from recently performed investigations [9] and carried out on the same oil-impregnated pressboard samples with different moisture content (m.c.) as used for the PDC measurements in Figure 4. A special dielectric spectrometer manufactured by Dielectric Instrumentation [12] was applied enabling measurements over an extremely wide frequency range for these small samples (3.16 10-4 Hz to 10 kHz). By applying good EM shielding of the instrumentation and the test cells, a test voltage of only 3 V was sufficient to perform the measurements. In Figures 6 and 7 the actual test frequencies can be identified as points on the curves. FD measurements always need a pair of magnitudes for each individual frequency, see eq. (19). Instead of displaying the results by the real and imaginary parts of the complex magnitudes for permittivity or susceptibility, the frequency dependence of capacitance and dissipation factor are shown in both figures. Figure 6 displays the real part of the capacitance C = C0 and Figure 7 the dissipation factor tan (eq. 20). The results confirm that the m.c. of the pressboard affects the low and very low frequency results much more pronounced than that at power frequency only. Figure 7 also shows that power and higher frequencies may not identify moisture contents with reliability. The increase in C (Figure 6) is obviously mainly caused by the conductivity of water and thus by increased interfacial polarization inside the pressboard. Please note that the vacuum capacitance C0 of the test cell, which can be computed from the dimensions of this cell, is only 44.3 pF. Increased interfacial polarization also produces the increase in dissipation factor, mainly in the low and very low frequency range. Measurements in the frequency domain need voltage sources of variable frequencies and, for applications re11

lated to HV power equipment, output voltages up to at least some hundreds of volts. Such measurements become quite lengthy if very low frequencies are considered. At least two cycles of an ac voltage are necessary to quantify the amplitudes and phase shift between voltage and currents. Therefore, up to more than 2000 seconds is necessary to get a single value for a C - tan measurement at a frequency of 1 mHz.

Alternative Measurement Techniques Related to Dielectric Response/Spectroscopy Measuring changes of its original dielectric response in the time and frequency domains are not the only means of detecting aging or the destructive decomposition of insulation materials and/or complete insulation systems. There are also many other methods to detect deterioration, e.g. changes in chemical, mechanical or optical behavior. As far as typical power transformer insulation is concerned, for which a combination of oil gaps and oil-impregnated pressboard is used, examples of such methods are the determination of the DP (degree of polymerization) for quantifying decomposition and thus mechanical strength of cellulose, oil parameter analysis including DGA (dissolved gas in oil) or HPLC (high performance liquid chromatography) for identifying decomposition products due to high temperatures (hot spots) or partial discharges, or Karl Fischer tests for detecting moisture in oil and paper or pressboard. A survey of these and similar techniques related to insulation monitoring can be found in [2] together with some hints to dielectric monitoring, to which this article is exclusively addressed. Before we go into an overview of alternative measurement techniques as applied to HV power equipment, the following must be mentioned: G Dielectric response techniques are global methods, i.e. each test object is treated as a black box accessible only by its electric terminals. Therefore, only global changes of the insulation can be identified but not localized de-

fects. Partial discharge (PD) measurements for instance can be applied to find such localized defects [2]. Inherent to all dielectric response measurements in either of the two domains is their off-line character, i.e. equipment in operation must be removed from service to perform the measurement. As dielectric response measurements should be performed on equipment installed on site in the field, and as the measurement instruments used are usually quite sensitive to electromagnetic disturbances, the electromagnetic compatibility of such instruments must be guaranteed. Therefore, the test voltage levels of the instruments cannot be too low.

Alternative Time Domain Techniques


One of the oldest methods to qualify dielectric properties (permittivity and losses) of anomalous dielectrics was, and still is, to perform a single Return or Recovery Voltage measurement. The expression anomalous was earlier used to define a dielectric whose after-effects also an earlier expression to identify polarization and depolarization currents and effects - could not have been expressed by a single number for capacitance and resistance, see e.g. [14]-[15]. The measurement principle is shown in Figure 8. Similar to Figure 3, a dc voltage of known amplitude UC is applied to the test object (which was previously completely discharged) during an interval t1 = TC, which should be so long that at its end the after-effects produced by the application of the potential UC have dispersed completely [15]. After a short, but often not well defined short-circuiting period, a return or recovery voltage, UR(t), can then be measured across the test object, if the input impedance of the voltmeter used is very high. (Note: The condition of a very high input impedance, even one century ago, was easy to fulfill by using electrostatic voltmeters; which is why the effects of relaxation or residual charge could be detected). The origin and source of the recovery voltages are the depolarization currents, idepol(t), as defined in eq. (10), i.e. the still active relaxation processes inside the dielectric material which did not relax during the short-circuit period. Early measurements confirmed [15] that the magnitudes and shapes of the recovery voltages, UR(t), are strongly dependent on the amplitude UC of the voltage and its duration t1, as well as on the duration of the short-circuit or grounding period, (t2 - t1), see Figure 8. If, therefore, any quantitative results of this technique have to be judged, all three factors have to be taken into account, i.e. they should be well quantified. All these effects and the results gained from the Return Voltage Technique can be easiest explained if we represent the behavior of a dielectric by an equivalent circuit, shown in Figure 9. This circuit can be traced back to Maxwell [8] and is thus sometimes called the Maxwell
IEEE Electrical Insulation Magazine

A) The Common Return Voltage Technique

Figure 6. Real part of the complex capacitance of the pressboard samples, Figure 4.

12

model [16], [27]. A simple background exists for this general equivalent circuit: As already shown, for slow polarization processes the depolarization currents and thus the dielectric response functions f(t) are monotonically decreasing, see Figure 4. Such dependencies can always be simulated by a superposition or a sum of exponential functions as shown in [9], [17] and elsewhere. This sum can be modeled most easily by a parallel connection of series RiCi elements, together with a high frequency capacitance C = C0 according to eq. (9), which can be determined by conventional methods, and the insulation resistance R0 (representing 0 in eq. (9)) of the test object. This resistance, R0, can either be taken from a more or less final value of the polarization/charging current, or it can be calculated from an already Figure 7. Dissipation factor tan of the pressboard samples, known dc conductivity, 0, of the dielectric and Figure 4. the geometry of the test sample. If this circuit, according to Figure 9, is suddenly charged with a dc voltage source UC during 0 t t1, the individual polarization currents including the constant current through R0 will flow into the circuit, charging instantaneously C and, with some delay, the capacitors Ci of the RiCi -elements according to their time constants (the magnitude and number of which depends on the simulation procedure applied). Depending on how long (t1) the object is charged, the different polarization processes as represented by the RiCi -elements become either fully or only partly activated. A very short grounding period from t1 < t t2 will only discharge C, but if this period is larger, the slower polarization processes with large time constants will also start to relax. As soon as the short circuit is opened for t > t2, the recovery voltage UR(t) can be Figure 8. Principle of a Return Voltage measurement. Charging measured under open circuit conditions, as the polarizawith U C during 0 t t 1 , grounding period from t 1 < t t 2 , for t > t 2 , tion processes (or RC elements), some of which were parthe recovery voltage is measured at open circuit conditions. tially relaxed during the short circuit period, will partly discharge into C and R0. The magnitude of the return voltage which is thus always proportional to UC, is deexist, the magnitude of which, however, become very pendent on the charging time t1 = TC and grounding pesmall and could no longer be measured in practice. The riod, (t2 - t1). The initial slope of the return voltage (SR in calculation assumes ideal conditions for the measureFigure 8) is proportional to the active depolarization curment of the return voltage slopes. The pronounced decay rent at time instant t2 and thus proportional to the of the voltages after its peak values is due to the magniintensity of the polarization process at this very specific tude of the insulation resistance, R0, in Figure 9. instant. One advantage of the return voltage measurement is The proportionality of the magnitudes of the return the fact that only one terminal of the test object is necesvoltages to UC is self-evident, but the other dependencies sary to connect the charging voltage to the test object and are more sophisticated. An example of the dependence of to perform the return voltage measurement, assuming the shape of the return voltage on the grounding period, the second terminal is at ground potential. If, however, Td = (t2 - t1), is shown in Figure 10. The results were calthe test object is a combination of individual sub-systems, culated from the measured depolarization current of the it is a considerable disadvantage not to be able to sense test sample, Figure 4 with an m.c. 2.5 %, and its specific such individual sub-systems. Another advantage is the equivalent circuit, Figure 9, for which the insulation reself-calibrating effect of this method: The magnitude of sistance is not too large. A charging duration of 100 s was the return voltages is always more or less related to the assumed to demonstrate the effect that, even for groundgeometric or high frequency capacitance C of the test ing periods of as high as 10000 s, a return voltage will still
September/October 2003 Vol. 19, No. 5

13

Figure 9. Equivalent circuit to model any linear dielectric.

objectsee Figure 8which can also be used to qualify the magnitude of the polarization processes. Even in the 1950s, some specialized return voltage measurement instruments for the detection of moisture in the insulation of transformer and for controlling the transformer drying process after batch production became obviously attractive. As a result, special instruments were developed in France, UK, and Russia. It is not known if such instruments are still in use, but further information can be found in [10], [19]-[20]. (NOTE: R0 may also be approximately evaluated by applying a voltage response measuring method [18] not discussed in this paper. This method is based on the Return Voltage Technique, but at the end of the charging period of between 10 to 1000 s duration, the test sample is disconnected from the dc source and a slowly decaying discharge voltage curve is measured. It is claimed that the negative slope of this curve is directly proportional to the conductivity of the test object. However, this statement is not completely accurate as the decay of this voltage is also somewhat influenced by depolarization processes. In [18], no quantitative instructions are provided about the duration of this open circuit period, which will be followed by a shorter short-circuit period of 1 to 100 s before the normal return voltage UR(t) and its initial slope are measured.)

B) Measurements of Polarization Spectra by Means of a Return Voltage Meter


About a decade ago, a particular return voltage measurement technique for the on-site assessment of the bulk dielectric properties for power transformers appeared. This method, now called RVM-technique, was originally proposed in [21] and in succeeding papers (e.g. in [22]-[23]). It became very attractive, as it was, and still is, claimed that the moisture content in the pressboard of a transformers composite insulation system can be quantified by analyzing a polarization spectrum evaluated from the measurement procedure. This spectrum is produced by applying a series of individual charging voltages UC to the test object, followed by a short-circuiting period as explained in Figure 8, at each cycle or step increasing the charging period t1 = Tc as well as the short-circuiting or grounding period (t2 - t1) = Td, and using a fixed ratio of (Tc/Td) = 2 for the measurement series. During the individual charging process as well as during the grounding period, no measurements are taken. After 14

Td has elapsed, the recovery voltage, UR(t), for a particular cycle, is recorded and from its peak value, the amplitude URmax is quantified together with the charging period Tc for that cycle. As the charging periods are usually subdivided by cycles of 1, 2, and 5 for each decade in time, either a discontinuous graph, URmax as a function of Tc, can be drawn or a continuous function, if the individual measuring points are combined by some kind of interpolation. The resulting curve is called the polarization spectrum as each maximum appears at a Tc-value, which approximately corresponds to a single relaxation time constant RiCi of the equivalent circuit, see Figure 9, if all other time constants are neglected in this circuit. A prerequisite for applying this method is the existence of a finite, neither too high nor too low an insulation resistance R0 of the test object, which is responsible for the pronounced peak value in an individual return voltage, UR(t). Otherwise the return voltage will decay too quickly or not decay at all. Such a simplified circuit with only four elements can analytically be solved as shown in [23]. There some computed return voltage curves are shown, but each only for a single relaxation time constant in our equivalent circuit, see Figure 9. Such an individual return voltage starts development at about 0.1% of this time constant and becomes close to zero at about 20 times this time constant, if an adequate insulation resistance (R0 in Figure 9) is assumed. Therefore, the resolution of more than one relaxation time constant, which differ by less than a factor of about 5 in magnitude, is very limited. A fundamental drawback of this method is the time necessary to perform a complete series of measurements necessary to obtain a polarization spectrum for any test object: A series usually starts with the shortest charging period Tc of 10 ms, that is then changed in steps of 2/5/10 and up to the longest, at 10000 s. The insulation system under investigation must be adequately discharged between each step before starting the next one. Thus, the time necessary to execute one step is not only the time of Tc + Tc/2+ time of UR(t) up to URmax , but there is also an additional time to discharge the test object to zero conditions. As the RVM-technique is based on quantifying the effects of a set of incomplete depolarization currents and incomplete specific return voltage curves, it provides no direct information about the insulation resistance R0 of the test object. It is, however, possible to calculate this resistance using the Newton iteration method if the initial return voltage increase rate (dUR/dt) is also measured and quantified. The results of such calculations are subject to uncertainty as was shown in [24]. Part II of this article will show some applications of this RVM-Technique. As this instrument is not able to perform measurements on test objects of small size or low capacitance, calculated polarization spectra are shown in Figure 11, based on the results of the PDC measurements shown in Figure 4. As will be seen in Part II, such
IEEE Electrical Insulation Magazine

calculations are possible and accurate since the dielectric response functions f(t) of these samples are already known. The results demonstrate the influence of moisture content (m.c.) on the shape of the polarization spectra, i.e. the shifting of the peak values to shorter values of charging duration (TC) with increasing m.c. Nevertheless, the RVM-technique is popular in diagnosing oil-paper insulation in power transformers due to the assumption that the moisture content in the pressboard can reliably be identified. However, there has been much controversy for some years surrounding this technique [25], which has been criticized for various reasons, e.g.: the moisture content of the pressboard evaluated from the polarization spectra very often yields much higher values than that obtained by other methods; the recommended interpretation scheme is too simplistic, and finally, the technique does not take into account the effects of the geometry and properties of the transformer oil. Therefore, in 1999 a CIGR Task Force (15.01.09) was set up to clarify these discrepancies. In Part II of this article, an example of these investigations, a summary of which was recently published [26], is provided. If the insulation resistance R0 of the test object becomes too large, the RVM-technique with its polarization spectrum will fail. This was experienced about ten years ago, but only one publication as an example will be mentioned [27] in which some individual return voltages with 15 minutes charging period displayed no peak value, URmax. Therefore, this technique and its evaluation procedure based on the polarization spectrum can only be applied for dielectrics whose insulation resistance R0 is not too large.

the specific IRC-Plot will show a bell-shaped curve whose maximum appears at the time constant of this single exponential function. As measured IRC-plots may also contain some local maxima, this measured plot is thenby means of somewhat complex and sophisticated evaluation proceduressimulated by the superposition of single exponential functions with amplitudes ai and relaxation times i as parameters. The latter parameter is related to charge traps or trap energy levels in the insulation, which may change with material, temperature and, for example, aging products. The shape of IRC-plots is quite different from the shape of the plots for the depolarization currents, as can be seen in Figure 12. Again, some of the measured depolarization currents in Figure 4 have been used to display typical shapes, but in this figure the time scale is much larger than those for other published results (see Part II), as the depolarization currents were available up to 200000 s, whereas the charging periods of common measurements last about 1800 s. The plot for an m.c. of 2.5 % indeed shows a bell-shaped curve whose maximum is obviously quite close to a dominant time constant in the range of some 1000 seconds which can also be identified

C) Isothermal Relaxation Current (IRC) Analysis


In recent years, another alternative diagnostic technique appeared in the literature called IRC-Analysis. Originally mainly applied to the investigation of PE and XLPE cables [29]-[30], the authors currently use this method for the diagnostics of other kinds of insulation materials used in HV technology [31]. The method is based on a depolarization current measurement which is usually performed with a charging voltage of 1 kV and a charging period of 1800 s followed by a discharging (short-circuiting) period of five seconds, during which no measurements are made [32]. The depolarization current, named relaxation current iR(t), is then measured and multiplied by the time t elapsed during the specific measuring period which is in general equal to the discharging period. But charging as well as discharging times may be variable and set by the experience of on-site investigations [29]; as a result, numerical values of these quantities are usually not indicated in the publications. From the results, an IRC-Plot, t times iR(t) as a function of log t, is made and used as an evaluation procedure. This procedure is based on the following: if iR(t) decays according to a single exponential function in the form of (a e-t/),
September/October 2003 Vol. 19, No. 5

Figure 10. Calculated return voltages for one sample (2.5 % m.c.) of Figure 4.

Figure 11. Calculated polarization spectra for 3 samples of Figure 4.

15

from the relevant depolarization current in Figure 4 or the relevant polarization spectrum in Figure 11. Apart from the fact that this method cannot quantify the dc insulation resistance R0 or the conductivity 0 of the test object, published results very likely indicate a systematic error of the evaluation procedure, as the maximum will always be dependent on the charging time. This can be well understood by the inherent behavior of a depolarization or relaxation current, which always decays more or less as an exponential function after about half of the charging period which is not sufficient to get a complete depolarization current up to the full discharging time, see Figure 5.
Figure 12. Calculated IRC-plots for 3 samples of Figure 4.

Alternative Frequency Domain Techniques The common measurement techniques for capacitance, C, and the dielectric dissipation factor, tan , according to eq. (20) at power frequency (i.e. at one single value in the frequency domain) by means of bridge circuits, which are based on standard capacitors, are well known and explained in many textbooks related to HV technology [33]. In general, such measurement can reveal adverse effects in insulation systems. Depending on the type of HV equipment under test, however, aging effects are not always sensitive to a single set of C- tan - values measured at power frequency. Many investigations, especially related to medium voltage power cables, have shown that many kinds of aging phenomena are more apparent at a much lower frequency range [34]. Thus, for some years C and tan have been measured at 0.1 Hz using 0.1 Hz HV test equipment, which has low power consumption. One of the reasons for the development of low frequency techniques is that for diagnostic tests of medium-voltage PE/XLPE cables, the loss factors are mainly sensitive to water treeing [35]. Measurements performed at one or even two single frequencies, however, cannot be treated as an alternative method to an extended dielectric spectroscopy, apart from the additional expenses as necessary for the 0.1 Hz HV test and measurement equipment. The measurement of C- tan - values (or complex permittivity values, eq. (19)) within a large frequency range and high voltages up to some kilovolts is a difficult task, as standard capacitors with precisely known and stable dielectric properties essentially in a frequency range of less than 1 Hz are not available. Higher voltages are necessary to avoid EMC problems during on-site measurements. This difficulty, which will be present in the already mentioned instrument for the measurements of Figures 6 and 7, was recently solved by the development of a new instrument based on frequency domain spectroscopy (FDS, [35]). Results of this, as well as of other new instruments, will be discussed and shown in Part II of this article.
16

Coherency Between Different Techniques Figures 10 to 12 have shown that results of alternative measuring techniques can be computed if the dielectric response function f(t) in an adequate time period is measured and known. Equation (14) also shows that an unique relationship exists between the time and frequency domains as long as the dielectric properties under investigation are linear. As most of the methods currently applied are time-domain methods and, as for all conversions the equivalent circuit of Figure 9 is very important, the modeling procedures based on this circuit, and described in more detail in [9], are briefly explained. For insulation systems in HV power equipment, all measured polarization and depolarization currentssee eqs. (9) and (10)are governed, within the time periods of interest, by monotonously decreasing functions. The behavior of the currents is essentially determined by the dielectric response function f(t) and the conductivity 0 of the test object with its geometric capacitance C0 or C respectively. If both currents have been measured for time periods large enough to identify the dc-component within the polarization current, their sum can then be used to identify the dc resistance R0 (= C0 0 / 0) of the test object. But the equivalent circuit of Figure 9 contains many more circuit elements: The still unknown elements Ri, Ci with their corresponding time constants i = Ri Ci represent the polarization and depolarization processes and must next be determined by the modeling procedure. This can be done by fitting the depolarization current with the equation
i depol ( t ) = where Ai = U c [1 exp( T c / i )] / R i , for i = 1iK n. (22)

Ai exp( t / i ),
i =1

(21)

In eq. (22), Tc is the period of time during which the step voltage Uc was applied.
IEEE Electrical Insulation Magazine

With a special, sequential algorithm not explained here, the a priori unknown coefficients Ai and time constants i, which are chosen equidistantly in the logarithmic time scale, are determined. In general, only two or three time constants per time decade are necessary to fit the measured depolarization current with sufficiently high accuracy. It should be emphasized that by this procedure the currents are now simulated by algebraic equations covering the time scale from zero to infinity although their validity is restricted to the measured time periods only. Now our equivalent circuit with its known circuit elements can be used for all kinds of transformations. The transformation into the frequency domain may easiest be understood by re-writing eq. (18) in the form of its relation between a measured sinusoidal current I( ) and a measured sinusoidal voltage U( ) as I ( ) = iC( )U ( ). (23)

tan ( ) =

n R iC i2 1 + R 0 i=1 1 + (R iC i ) 2

C +
i =1

1 + (R iC i ) 2

Ci

. (29)

Here, C() is the complex capacitance which is related to the relative dielectric permittivity, eq. (19), as C( ) = C ( ) iC ( ) = C 0 { r ( ) i[ r ( ) + 0 / 0 ]}.

(24)

Here, C0 and 0 again represent the vacuum or geometric capacitance and the dc conductivity of the test object, respectively. The dissipation factor tan () can be re-written using eq. (20) as: tan ( ) = C ( ) , C ( )

(25)

as the losses due to dc conductivity are already included in C ( ). For the equivalent circuit, Figure 9, the complex capacitance C() can now be calculated according to eq. (23) from its complex admittance, Y() as: C( ) =
n Ci Y ( ) 1 = C + + . i iR 0 i=1 1 + iR iC i

Equations (26) to (29) are simple algebraic equations dependent only on frequency as all other magnitudes are already known. A numerical computation of the results for any frequency is thus straightforward. Figures 6 and 7 display measured results of the test sample of Figure 4, and are thus not computed. As the measurements of Figure 4 have only be made within a time period of five decades, it would not be possible to simulate and calculate more than seven decades in the frequency domain. But the very good agreement of calculated and measured results was already shown in [9] or [17]. Figures 10 and 11 showed calculated results for return or recovery voltages and polarization spectra respectively. As already explained, polarization spectra are based on the results of a series of return voltage measurements. The following calculation procedure, again based on the equivalent circuit of Figure 9, can thus be applied for any transformation of a PDC measurement into return voltages. According to Figure 8, the return voltage uR(t) is measured under open circuit after charging the sample with a charging voltage, UC, during a charging period, TC = t1, and a discharging (grounding) period of Td = t2 - t1. The time scale for uR(t) will thus start for t t2. During TC not all of the circuit elements i = 1...n may have been charged completely and during the grounding period Td all of the capacitors Ci have already been discharged to some extent, so that the boundary conditions at time t = t2 for solving the differential equations are: u R ( t 2 ) = 0, (30)

and for the remaining voltages at each RiCi element u i ( t 2 ) = U c [1 exp( T c / R iC i )]exp( T d / R iC i ).

(26) (31) The return voltage can now be computed by the following differential equations: du i ( t ) 1 = [u R ( t ) u i ( t )], for i = 1K n dt R iC i and (28) du R ( t ) 1 = dt C0 u r ( t) n u r ( t) u i ( t) R . Ri 0 i =1 (32)

The real and imaginary parts of C() are then given as C ( ) + C +


i =1 n

1 + (R iC i ) 2

Ci

(27)

and C ( ) =
n R iC i2 1 + . R 0 i=1 1 + (R iC i ) 2

Finally, tan () can be calculated from:


September/October 2003 Vol. 19, No. 5

(33) 17

As TC as well as Td in these equations can have any values, a PDC measurement from which the equivalent circuit has been derived thus provides any kind of a return voltage measurement. For the calculation of an IRC-plot, Figure 12, no transformation but only a multiplication by the time t is necessary. If Frequency Domain Spectroscopy is used, transformations into the time domain are also possible. The procedures, however, are more difficult to apply. The reader interested in such successfully investigated and applied algorithms can find an in-depth treatment in [36]-[37].

Conclusions This review introduces the theoretical background of dielectric spectroscopy in the time and frequency domains and provides an overview about the specific measuring methods based on this background. The specific methods treated are used for diagnosing electric insulation materials used in power engineering. It indicates that some of these methods may not be sufficient to gain full information about the actual conditions of a test object and that either measurements of polarization and depolarization currents (PDC) in the time domain or measurements of C- tan - values (or complex permittivity) in the frequency domain (FDS) should be preferred to obtain a dielectric response function which offers much more information to judge the actual state of an insulation material or system. In Part II, the application of some of these methods are treated in more detail and the results of some investigations related to the PDC-, RVM- and FDS-methods will be shown. Acknowledgement Special thanks are due to my former student, V . Der Houhanessian, for providing permission to publish some results of his thesis [9], and for preparing Figures 10 to 12.
Walter S. Zaengl (SM1970-F1998), was born in 1931 and received his Dipl-Ing and doctorate (Dr.-Ing.) in electrical engineering from the Technical University of Munich/Germany in 1955 and 1964 respectively. In 1955 and 1956 he was with AEG, Kassel, Germany, working on the development of HV power circuit breakers and HV testing. From 1956 to 1969 he worked with the Technical University, Munich to support the construction of new HV laboratories for research and testing, to perform research on impulse voltage measurements for very high voltages (thesis on capacitor voltage dividers) and to teach courses on HV measurement and testing techniques. In 1969, he joined General Electric, in Pittsfield, MA, USA, for a special project.

In October 1969, he joined the Swiss Federal Institute of Technology (ETH) of Zurich, Switzerland as a full professor and head of the High Voltage Laboratory. Research under his direction was mainly focused on the development of advanced techniques for HV measuring and testing systems, on basic research in gaseous discharges (SF6 and gas mixtures), on applied research and diagnostic methods for HV power equipment, and on hazardous gas treatment by means of non-thermal electrical discharges. He retired from full service at ETH in October 1996. Professor Zaengl has authored about 120 publications and is the co-author of three books on HV engineering. He is a Fellow of the IEEE and a member of VDE (Germany), SEV (Switzerland) and a distinguished member of CIGR. He was convenor of the IEC working group establishing the latest IEC-Standard (No. 60270) for Partial Discharge (PD) Measurements, issued in 2000, and is member of a CIGR Task Force, Group (15-01-09), dealing with the application of dielectric phenomena for transformer diagnostics.

References
[1] O.H. Hoelsaeter, Transmission system operators in Europe, Electra, Special Issue 2000, pp. 17-19. [2] B.H. Ward, A survey of new techniques in insulation monitoring of power transformers, IEEE Electrical Insulation Magazine, vol. 17, no. 3, 2001, pp. 16-23. [3] A.R. von Hippel, Dielectric Materials And Applications, New York: The Technology Press of MIT and J. Wiley & Sons, 1958. [4] A.K. Jonscher, Dielectric Relaxation in Solids. London: Chelsea Dielectrics Press, 1983. [5] A.K. Jonscher, The universal dielectric response: part I, IEEE Electrical Insulation Mag., vol. 6, no. 2, 1990. [6] A.K. Jonscher, The universal dielectric response: part II, IEEE Electrical Insulation Mag., vol. 6, no. 3, 1990. [7] A.K. Jonscher, The universal dielectric response: part III, IEEE Electrical Insulation Mag., vol. 6, no. 4, 1990. [8] J.C. Maxwell, A Treatise on Electricity and Magnetism, vol. 1, 3rd ed. Oxford: Clarendon Press, reprint by Dover, pp. 450-464, 1981. [9] V . Der Houhanessian, Measurement and analysis of dielectric response in oil-paper insulation systems, Ph.D. thesis, Swiss Federal Institute of Technology, No. 12832, Zurich, 1998. [10] M.J. Fabre, Un nouvel appareil dessai des isolations au papier imprgn: Labsorptiomtre rcurrence, Revue Gnrale de Lctricit, vol. 66, pp. 447-457, Sept. 1957. [11] U. Gfvert, G. Frimpong, and J. Fuhr, Modeling of dielectric measurements on power transformers, paper 15-103, presented at the International Council on Large Electric Systems (CIGR), Paris, France, 1998 Session. [12] J. Pugh, Dielectric measurements using frequency response analyzers, in Proc. 4th Int. Conf. on Dielec. Mats., Mes. and Apps., London, England, 1984, IEE Conference Publication 239, pp. 247-250.

18

IEEE Electrical Insulation Magazine

[13] U. Gfvert: Condition assessment of insulation systems - analysis of dielectric response methods, in Proc. Nord IS 96, Bergen, Norway, 1996, pp. 1-20. [14] J.B. Whitehead, Lectures on Dielectric Theory and Insulation. New York: McGraw-Hill, 1927. [15] B. Gross, On after-effects in solid dielectrics, Physical Review, vol. 57, pp. 57-59, 1940. [16] B. Gross, Electret research - stages in its development, IEEE Trans. Elect. Insulation, vol. EI-21, pp. 249 269, June 1986. [17] V . Der Houhanessian and W .S. Zaengl, Application of relaxation current measurements to on-site diagnosis of power transformers, in 1997 CEIDP , Minneapolis, MN, 1997, IEEE Publication 97CH36046, pp. 45-51. [18] E. Nemeth, Measuring the voltage response, a diagnostic test method of insulation, paper 72.06, presented at the 7th International Symposium on High Voltage Engineering (ISH), Dresden, Germany, 1991. [19] G. Mole, Improved test methods for the insulation of electrical equipment, in Proc. Inst. Electr. Engr., IIA, no.3, 1953. [20] B.A. Alekseev, A new method for detecting moisture in the insulation of transformer windings, Translation from Elektrichestvo, 1958. [21] A. Bognar, L. Kalocsai, G. Csepes, E. Nemeth, and J. Schmidt, Diagnostic tests of high voltage oil-paper insulating systems (in particular transformer insulation) using dc dielectrometrics, paper 15/33-08, presented at the International Council on Large Electric Systems (CIGR), Paris, France, 1990. [22] A. Bognar, G. Csepes, L. Kalocsai, and I. Kispal: Spectrum of polarization phenomena of long time-constant as a diagnostic method of oil-paper insulation systems, in Proc. 3rd Int. Conf. on Prop. and Apps. of Dielec. Mats., Tokyo, Japan, 1991, pp.723-726. [23] A. Bognar, G. Cspes, I. Hamos, I. Kispal, and P . Osvath, Comparing various methods for the dielectric diagnostics of oil-paper insulation systems in the range of low frequencies or long time constants, paper 21.01, presented at the 8th International Symposium on High Voltage Engineering (ISH), Yokohama, Japan, 1993. [24] S.M. Gubanski, M. Kvarngren, and C. Stec, Return voltage characteristics of oil-paper insulated HVDC cables, paper 67.05, presented at the 8th International Symposium on High Voltage Engineering (ISH), Yokohama, Japan, 1993. [25] A.J. Kachler, Aging and moisture determination in power transformer insulation systems: contradiction of RVM methodology, effects of geometry and ion conductivity, presented

at the 2nd International Workshop on Transformers, Lodz, Poland, 1999. [26] CIGR Working Group 15.01, Task Force 09, Dielectric response methods for diagnostics of power transformers, Electra, 24-37, June 2002. [27] E. Ildstad, and H. Faremo, Water treeing and diagnostic testing of different XLPE cable insulations, paper 20.05, presented at the 8th International Symposium on High Voltage Engineering (ISH), Yokohama, Japan, 1993. [28] B. Gross, Dielectric relaxation functions and models, J, Appl. Phys., vol. 67 (10), pp. 6399-6404, 1990. [29] H.-G. Kranz, D. Steinbrink, and F. Merschel, IRC-analysis: a new test procedure for laid XLPE-cables, in Proc. Vol. 4, 10th Int. Symp. High Voltage Engineering (ISH), Montral, Canada, 1997, pp. 207-210. [30] G. Hoff, and H.-G. Kranz, Correlation between return voltage and relaxation current measurements on XLPE medium voltage cables, presented at the 11th International Symposium on High Voltage Engineering (ISH), London, England, 1999. [31] R. Hofmann, and H.-G. Kranz, IRC-analysis: destruction free dielectric diagnosis of epoxy resin impregnated insulation, contribution 4-93, presented at the 12th International Symposium on High Voltage Engineering (ISH), Bangalore, India, 2001. [32] R. Plath, W . Kalkner, and I. Krage, Vergleich von Diagnosesystemen zur Beurteilung des Alterungszustandes PE/VPE-isolierter Mittelspannungskabel, Elektrizitaetswirtschaft, Jg. 96, pp. 1130-1140, 1997. [33] E. Kuffel, W .S. Zaengl, and J. Kuffel, High Voltage Engineering, Fundamentals (Second Edition), ISBN 0-7506-3634-3, Burlington, MA: Newnes (Butterworth-Heinemann), 2000. [34] R. Neimanis, On estimation of moisture content in mass impregnated distribution cables, Thesis, Dept. of El. Eng., Royal Ins. of Tech. (KTH), Stockholm, Sweden, 2001, ISSN 1100-1592 [35] P . Werelius, Development and application of high voltage dielectric spectroscopy for diagnosis of medium voltage XLPE cables, Thesis, Dept. of El. Eng., Royal Ins. of Tech. (KTH), Stockholm, Sweden, 2001, ISSN 1650-674x. [36] A. Helgeson, Analysis of dielectric response measurements methods and dielectric properties of resin-rich insulation during processing, Thesis, Dept. of El. Eng., Royal Ins. of Tech. (KTH), Stockholm, Sweden, 2000, ISSN 1100-1593. [37] R. Neimanis, On estimation of moisture content in mass impregnated distribution cables, Thesis, Dept. of El. Eng., Royal Ins. of Tech. (KTH), Stockholm, Sweden, 2001, ISSN 1100-1592. pp.

September/October 2003 Vol. 19, No. 5

19

Potrebbero piacerti anche