You are on page 1of 79

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/279448671

Capacitor Reliability in Photovoltaic Inverters

Research · June 2015


DOI: 10.13140/RG.2.1.1093.2327

CITATIONS READS
0 695

1 author:

Jack Flicker
Sandia National Laboratories
71 PUBLICATIONS   713 CITATIONS   

SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Voltage Regulation and Protection Assurance using DER Advanced Grid Functions View project

Microinverter reliability improvement View project

All content following this page was uploaded by Jack Flicker on 30 June 2015.

The user has requested enhancement of the downloaded file.


SANDIA REPORT
dd
Unlimited Release
Printed September 17, 2012

Capacitor Reliability in Photovoltaic


Inverters
Jack D. Flicker

Prepared by
Sandia National Laboratories
Albuquerque, New Mexico 87185 and Livermore, California 94550

Sandia National Laboratories is a multi-program laboratory managed and operated by Sandia Corporation,
a wholly owned subsidiary of Lockheed Martin Corporation, for the U.S. Department of Energy’s
National Nuclear Security Administration under contract DE-AC04-94AL85000.

Approved for public release; further dissemination unlimited.


Issued by Sandia National Laboratories, operated for the United States Department of Energy
by Sandia Corporation.

NOTICE: This report was prepared as an account of work sponsored by an agency of the United
States Government. Neither the United States Government, nor any agency thereof, nor any
of their employees, nor any of their contractors, subcontractors, or their employees, make any
warranty, express or implied, or assume any legal liability or responsibility for the accuracy,
completeness, or usefulness of any information, apparatus, product, or process disclosed, or rep-
resent that its use would not infringe privately owned rights. Reference herein to any specific
commercial product, process, or service by trade name, trademark, manufacturer, or otherwise,
does not necessarily constitute or imply its endorsement, recommendation, or favoring by the
United States Government, any agency thereof, or any of their contractors or subcontractors.
The views and opinions expressed herein do not necessarily state or reflect those of the United
States Government, any agency thereof, or any of their contractors.

Printed in the United States of America. This report has been reproduced directly from the best
available copy.

Available to DOE and DOE contractors from


U.S. Department of Energy
Office of Scientific and Technical Information
P.O. Box 62
Oak Ridge, TN 37831

Telephone: (865) 576-8401


Facsimile: (865) 576-5728
E-Mail: reports@adonis.osti.gov
Online ordering: http://www.osti.gov/bridge

Available to the public from


U.S. Department of Commerce
National Technical Information Service
5285 Port Royal Rd
Springfield, VA 22161

Telephone: (800) 553-6847


Facsimile: (703) 605-6900
E-Mail: orders@ntis.fedworld.gov
Online ordering: http://www.ntis.gov/help/ordermethods.asp?loc=7-4-0#online

NT OF E
ME N
RT
ER
A
DEP

GY

• •
IC A
U NIT

ER
ED

ST M
A TES OF A

2
dd
Unlimited Release
Printed September 17, 2012

Capacitor Reliability in Photovoltaic Inverters

Jack D. Flicker

Abstract

In order to decrease the cost of ownership of photovoltaic systems, less costly, more reliable
photovoltaic inverters must be developed. Capacitors are a significant cause of inverter fail-
ures and system inefficiencies, so a thorough understanding of their strengths and weaknesses
with regards to inverters is necessary. This paper summarizes the current issues surrounding
the use of capacitors in photovoltaic inverters and discusses the construction, use, lifetime,
and reliability of two types of capacitors, electrolytic and metallized thin film, regularly used
in photovoltaic inverters.

3
Acknowledgment

This work was funded by the DOE Office of Energy Efficiency and Renewable Energy.
Sandia National Laboratories is a multi-program laboratory managed and operated by San-
dia Corporation, a wholly owned subsidiary of Lockheed Martin Corporation, for the U.S.
Department of Energy’s National Nuclear Security Administration under contract DE-AC04-
94AL85000.

4
Contents

Acknowledgment 4

Nomenclature 10

Capacitor Reliability in Photovoltaic Inverters 13

Overview of Inverter Reliability Importance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

Capacitor Requirements in Inverters . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14

Capacitors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

Types of capacitors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

Electrolytic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

Metallized Thin Film . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

Ceramic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

Capacitor Failure Modes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

Electrolytic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

Metallized Thin Film . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

Ceramic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

Lifetime Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

Electrolytic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41

Metallized Thin Film . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

Ceramic . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

Reliability of Capacitors . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54

Summary and Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 63

5
References 65

6
List of Figures

1 Reliability in photovoltaic installations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

2 Circuit diagram of 2-phase inverter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 16

3 Voltage output of unipolar PWM inverter . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

4 Voltage ripple vs. DC-link capacitance . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

5 Effect of voltage ripple of PV MPP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20

6 Maximum allowable ripple voltage vs. PV utilization ratio . . . . . . . . . . . . . . . . 22

7 Maximum allowed voltage ripple vs. VM P P . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22

8 Minimum capacitor size for a given utilization ratio for BP4160 . . . . . . . . . . . 24

9 Physical schematic of generic capacitor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25

10 Equivalent circuit of a generic capacitor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28

11 Schematic of an aluminum elcap. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 29

12 Oxide interaction in an elcap . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

13 Oxide tunnels in elcaps . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31

14 Self-healing and fusing in metallized thin film capacitors . . . . . . . . . . . . . . . . . 36

15 Schematic of a multilayer ceramic capacitor. .......................... 38

16 ESR(t) vs. original ESR for an elcap using the linear inverse model . . . . . . . . 44

17 ESR of a dielectric capacitor versus frequency and temperature . . . . . . . . . . . . 45

18 Heat transfer in a capacitor . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47

19 Metallized thin film capacitance degradation over time . . . . . . . . . . . . . . . . . . . 52

20 Lifetime statistics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56

21 Lifetime distributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

22 Bathtub curve . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

7
23 Availability of a system . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

8
List of Tables

1 Electrical power standards for North American and Europe . . . . . . . . . . . . . . . 14

2 Electrical parameters of a monocrystalline Si solar cell, BP4160 . . . . . . . . . . . 21

3 Relative permittivity and dielectric strength for commonly used dielectrics . . 26

9
Nomenclature

PV Photovoltaic

DOE Department of Energy

MTBF Mean time between failure

MPPT Maximum power point tracking

MPP Maximum power point

CVT Constant voltage tracking

INC Incremental conductance

P&O Perturb and observe

FF Fill factor

V ripple Ripple voltage

BP British Petroleum

Isc Short circuit current

Voc Oper circuit voltage

PM P P Power at maximum power point

VM P P Voltage at maximum power point

IM P P Current at maximum power point

Irs Reverse saturation current

n Diode ideality factor

Q Charge

d Distance
~
E Electric field

Vbreakdown Material breakdown voltage

r Relative permittivity

10
Eds Dielectric strenth

Ileakage Leakage current

ESR Equivalent series resistance

ESL Equivalent series inductance

elcap Electrolytic capacitor

Ca Anode capacitance

Cc Cathode capacitance

Ceq Equivalent capacitance

MLCC Multi-layer ceramic capacitor

Rth Thermal resistance

hconvection Convection thermal transfer coefficient

hradiation Radiative thermal transfer coefficiency

∆t Time step

RH Relative humidity

PDF Probability density function

CDF Cumulative density function

R(t) Reliability at time, t

LF Lifetime

α Confidence level

FIT Failures in time

k, λ Weibull shape and scale functions

MTTF Mean time to failure

MTTR Mean time to repair

Φ System availability

Vrated Device maximum rated applied voltage

11
12
Capacitor Reliability in Photovoltaic
Inverters

Overview of Inverter Reliability Importance

Photovoltaic (PV) power systems have, historically, not been able to compete with con-
ventional fossil fuels without favorable regulatory policies and financial incentives. The US
Department of Energy (DOE), however, has estimated that PV systems can reach economic
competitiveness with other energy source at an installed system cost of $1/Watt ($0.05-
0.06/kWh) [1]. With a $1/Watt price-point, the DOE estimates PV penetration into the US
power market to be upwards of 18% by 2050.

As of 2010, the installed cost of PV systems was $3.40/Watt [1]. While most research,
both historically and currently, focuses on the production costs of PV module technology,
the cost of the necessary grid-connected DC-to-AC inverters has been largely ignored. As
the price of PV modules drops, the price of inverters becomes more important. Inverters
now constitute 8-12% of the total lifetime PV cost [2]. As of 2010, the inverter and asso-
ciated power conditioning components accounted for $0.25/Watt [3], well above the DOE
benchmark of $0.10/Watt by 2017 [1].

One of the key price drivers of the inverter component is inverter reliability [4, 5]. PV
modules have long lifetimes with warranties offered up to 20 years. The mean time between
failure (MTBF) of these systems have been shown, in the field, to be up to 522 years for
residential and 6,666 years for utility systems [6]. In contrast, the inverter component has
shown a field MTBF of between 1 and 16 years [6, 7]. Typical warranties on these devices last
only three to five years [4], with a few (e. g. certain united by PV Powered) with warranties
up to ten years. Even with the most optimistic view of inverter lifetime, it will be necessary
to replace or repair an inverter multiple times over the lifetime of a PV module. Repairing
the inverter is costly not only due to replacement parts and work crews, but also incidental
costs, such as the loss of power not generated during downtime, purchasing replacement
power unsupplied by the offline system, and any performance degradation before failure
identification [4].

Inverter issues are especially problematic since, depending on the PV topology, they may
affect the power production of a large number of modules and/or strings. According to
SunEdison, a North American-based PV plant operator with over 750 PV plants across the
world, inverter performance was responsible for 36% of total energy losses between January
2010 and March 2012 (Figure 1) [8].

13
Table 1: Electrical power standards for North American (IEEE15347) and Europe (IEC61727)

Category IEEE1547 IEC61727


Nominal Power (kW) 30 10
THD <5.0% <5.0%
Power factor >0.94 0.9 (at 50% rating)
DC current injection <0.5% <1%
Voltage range (V) 97-121 (88-110%) 196-253 (85-110%)
Frequency range (Hz) 50±1 59.3-60.5

The inverter is less reliable than other components in the PV system because it is a com-
plicated switching/monitoring system with a number of responsibilities. The main purpose
of an inverter is to output power meeting power quality standards (e.g IEEE 1547 [9] in
North America or IEC61727 [10] in Europe as shown in Table 1). However, depending on
the local/national ordinances or complexity of the device, the inverter may also be required
to manage power output of the PV module, connect/disconnect from the grid, manage var,
read and report status, or monitor islanding [2]. Added inverter functionalities exponentially
increase the difficulty in creating both reliable and affordable components.

In addition to having a number of functionalities, the inverter must also operate in rel-
atively harsh, changing environments. Inverters may experience large temperature swings
from -30 to 70o C, humidity conditions from 0 to 100%, and/or salty, corrosive environments.
Many consider DC bus capacitors to be the weak link in inverter reliability [2, 11], decreas-
ing inverter lifetimes by 50% [12] due to constant temperature and power cycling [13] and
high internal capacitor temperature [14]. According to SunEdison, over a 27 month period,
capacitor issues were responsible for 7% of total energy losses [8]. There is currently much
active, and sometimes all too public, debate in the industry, lead by Enphase Energy and
SolarBridge Technologies regarding the reliability of electrolytic vs. thin film in inverters.

In short, inverter technology must become half as costly and twice as reliable to facilitate
grid parity of solar energy in the future [4, 15]. With capacitors generally considered to be
the most unreliable element of the inverter, they represent the component which can most
easily be altered in the short term to increase inverter lifetime and decrease lifetime PV
system cost.

Capacitor Requirements in Inverters

Assuming that grid voltage and current are in phase and contain only the fundamental
frequency, the power in the grid may be described by (1).

14
Failure Areas: Frequency and Energy Impact
50%

45%
Tickets
40%
Energy Loss
35%

30%

25%

20%

15%

10%

5%
Inverter
0%
Components that Fail
Inverter AC External Other Support DC Planned Modules Weather Meter
Subsys Struct Subsys Outage Stn

0% 5% 10% 15% 20% 25% 30%


INTRODUCTION SYSTEMS INVERTERS MODULES FUTURE
Control Software
Card/Board
AC Contactor
Fan(s)
Matrix/IGBT
Power Supply
Tickets
AC Fuses
DC Contactor Energy Loss

Surge protection
GFI Components
Capacitors
Internal Fuses
Internal Relay/Switch
DC Input Fuses
Additional fields

INTRODUCTION SYSTEMS INVERTERS MODULES FUTURE

Figure 1: Between January 2010 and March 2012, inverter issues were responsible for 43% of
maintenance tickets and 36% of energy lost at 350 PV installations operated by SunEdison (left).
Of the tickets, capacitors issues were responsible for 7% of energy losses [8].

P (t) = V (t) · I(t)


= V · I · sin2 (ωt)
= V · I [1 + sin (2ωt)] (1)

where ω is the grid frequency.

The purpose of an inverter is to inject power into the grid at the same frequency and with
a small phase angle (φ ≈ 0). As can be seen from (1), the power injection from the inverter
is actually accomplished through power pulses at a frequency of 2ω [16]. This pulsewise

15
2!# %2!#
3456## 3456##
&7-+# &7-+#
$# $# 84!# $# $#
%# %#
84!# 84956## %# %#
84956## :)70-;<+#
:)70-;<+#

&'()*+#
,-.(*/0-*+#
$
!"!# #
,B=:# %#
!"!# =(>#?0@0*7/')#
$
!"!# #
%#
?#
1 3
1# A#
2 4
1'0.#

1#

Figure 2: Circuit diagram of a basic PWM inverter utilizing IGBT switches.

injection of power results in an AC component along with the nominally DC power. This
AC component is known as voltage ripple (V ripple ) and exists throughout the inverter/module
circuit.

One of the simplest methods of DC-to-AC signal conversion is to use an H-bridge topology.
In this setup, four switches are used to vary the DC input into a varying output signal which
can then be filtered into an AC signal suitable for injection into the grid or used by a load.
One method of H-bridge control is known as unipolar pulsed width modulation (PWM).
Figure 2 shows a simple PWM inverter utilizing four IGBTs in an H-bridge configuration.
A system of inductors and capacitors are used as output filters across the load. Each IGBT
is gate-controlled by either a comparator or inverting comparator.

A unipolar PWM inverter works by converting an input voltage (VP V ) into a pulsed
voltage at levels of −VP V , 0, and +VP V . IGBT #1 is controlled by a circuit which compares
a 60Hz sine wave to a 10kHz triangle wave. IGBT #2 is controlled by an inverting comparator
with the same inputs, so it is closed when IGBT #1 is open and vice-versa. IGBT #3 is
controlled by a circuit which compares the same 10kHz triangle wave with a sine wave 90o out
of phase with the sine wave used for IGBTs #1 and 2. IGBT #4 is controlled by an inverting
comparator with the same inputs, so it is closed when IGBT #3 is open and vice-versa.

The output of this type of inverter can be shaped into an AC signal the same frequency
as the input sine wave. The specific frequency of the reference triangle wave is unimportant.
Generally, it is a trade-off between AC signal quality and switching losses [17], though the

16
frequency of the triangle wave should be much larger than the sine wave in order to eliminate
undesirable low frequency harmonics which are difficult to filter out. The amplitude of the
output voltage is determined by (2):

max
max Vtri
Vout = V
max P V
(2)
Vsine

where:
out
Vmax is the amplitude of the output AC signal
tri
Vmax is the amplitude of the reference triangle
wave
sine
Vmax is the amplitude of the sine wave

If IGBTs #1 and 3 are closed, then the output of the H-bridge is at 0V. If IGBTs #1
and 4 are closed, the output is +VP V . If IGBTs #2 and 3 are closed, the output voltage
is −VP V . Finally, if IGBTs #2 and 4 are closed, then the output voltage is 0V. Figure 3
shows the voltage input (red), output of the H-bridge (blue) and filtered AC wave (black)
of a unipolar PWM inverter. The inset of Figure 3 shows that due to the inductance in the
circuit and the switching frequency, the current does not smoothly follow the voltage, but
ripples around it.

This ripple in the current is determined by the input voltage, load/filter inductance, and
switching frequency. Assuming a duty cycle of 50%, the maximum current ripple can be
described by (3):

ripple VP V
Imax = (3)
4·L·f

where:
VP V is the PV module input voltage
L is the load/filter inductance
f is the inverter switching frequency

The ensuing voltage ripple in the DC input can be calculated by solving the current flow
across the bus capacitor to be (4):

17
"##$
/0,1"#%2(345(6,&0%#(
%##$
-./012$340540$
&##$ 67$7890:;<$
=8:>$7890:;<$
'##$
!"#$%&'()!*(
#$
'()*+!#&$ '(,*+!#&$ &(#*+!#&$ &(&*+!#&$ &("*+!#&$ &()*+!#&$ &(,*+!#&$ %(#*+!#&$ %(&*+!#&$
!'##$

!&##$

-.$//)0(
!%##$

!)*+,(
!"##$ !"#$%&'( +,-'().*(

Figure 3: Voltage output of a unipolar PWM inverter. The PV input voltage (red) is transformed
into a pulsed voltage of −VP V , 0, or +VP V by an H-bridge (blue). This pulsed voltage supply
is transformed into an AC signal through filtering (black). (Inset) Due to circuit inductance, the
current in the circuit does not smoothly follow the load, but ripples around it [18].

ripple VP V
Vmax = (4)
32 · L · C · f 2

where:
C is the bus capacitance

As can be seen, the ripple voltage is inversely proportional to the bus capacitance. A
large capacitance is necessary to reduce V ripple (Figure 4).

The PV module is hypersensitive to this V ripple . Its presence dramatically reduces avail-
able output power [16]. In order to limit this voltage ripple, each single-phase inverter re-
quires an energy storage element (either a capacitor or inductor) [19]. However, for economic
or engineering reasons, the number or size (either physical or electrical) of these storage el-
ements must be limited and some magnitude of V ripple can be found in all single-phase
inverters at the DC side of the circuit [20].

One of the main function of an inverter, other than power conditioning, is maximum
power point tracking (MPPT). MPPT is the process where the voltage applied to a module
is varied so that the module constantly operates at or near the maximum power point (MPP),
the point on the IV curve of a module which defines the maximum possible output power.
There are multiple algorithms, ranging from relatively simple to complex, which can be used
by the inverter for MPPT. Some of the most popular of these algorithms include constant
voltage tracking (CVT) , perturb and observe (P&O or mountain-climb) [21], incremental

18
120  
Vripplemax  vs.  Bus  Link  Capacitance    
100  
for  Unipolar  PWM  Inverter    

80  

Vripplemax    (V)  
60  

40  

20  

0  
0   500   1,000   1,500   2,000  
Bus  Link  Capacitance  (µF)  

ripple
Figure 4: Vmax vs. Bus Link Capacitance for a unipolar PWM inverter with a load inductance
of 100µH, a voltage input of 325V, and a switching frequency of 9.8kHz

conductance (INC) [22], and periodic scanning [20]. Most of the more complex algorithms
can easily reach efficiencies (known as tracking factors) >95% [23].

For an ideal MPPT algorithm, the applied voltage to the module always results in the
maximum possible power output (PM P P ) from the module. This applied voltage is known
as the voltage at maximum power point (VM P P ). The presence of a voltage ripple in the
inverter/module circuit alters the voltage applied to the module from VM P P to VM P P ±V ripple
(Figure 5). This causes a corresponding ripple in the power output of the module. Depending
on the module fill factor (FF), or the squareness of the module IV curved, a small deviation
in VM P P can have a large deleterious effect on the average module power output (Pavg ).

This decrease in the average power output decreases the module utilization ratio, kP V ,
the ratio between the average power output and PM P P (5).

Pavg
kP V = (5)
PM P P

ripple
For a given utilization ratio, the maximum allowable ripple voltage, Vmax , is defined by
(6).

s
ripple 2 · PM P P · (kP V − 1)
Vmax = (6)
3 · α · VM P P + β

19
./-(

+,,( ,"3$(
,%4&(

!0122#'(

!"#$%&'()!*( !"-(

Figure 5: Ripple Voltage (V ripple ) effect on maximum power point (MPP) of solar module output
power [20].

where:
α and β are the coefficients for a 2nd Taylor Ap-
proximation to the PV module IV curve at the
MPP (IM P P + ∆I ≈ αV 2 +βV+γ) with values

1 d2 IM P P
α= ·
2 dVM2 P P
 2
1 q qVM P P
= −Irs · · e k·A·T ·n (7)
2 kB · A · T · n

where:
Irs is the reverse saturation current (A)
q is the fundamental charge constant, 1.6 ×
10−19 C
kB is Boltzmann’s constant, 13.8−24 J/K
A is the diode quality factor
T is the cell temperature in Kelvin
n is the number of cells in series

20
and
dIM P P 1 d2 IM P P
β= −2· · · VM P P
dVM P P 2 dVM2 P P
 q   q·V 
qVM P P
MP P
= Irs · · − 1 · e k·A·T ·n (8)
k·A·T ·n k·A·T ·n

Equations (6), (7), and (8) depend heavily on the type of PV module used. The values
of A, n, Irs , PM P P , and VM P P can vary greatly (±5%) from manufacturer to manufacturer
and even between two modules of the same model, so α and β are typically only estimates
for any given module. Table 2 shows relevant electrical data from a typical monocrystalline
Si PV module. This module (model BP4160) is produced by British Petroleum (BP) and
contains 72 monocrystalline Si cells in series.

Table 2: Electrical characterization and energy output values for a British Petroleum (BP)
monocrystalline Si solar cell, BP4160 [24].

Figure of Merit Value


Short CIrcuit Current (Isc ) 4.9 A
Open Circuit Voltage (Voc ) 44.2 V
Power at Max Point (PM P P ) 160 W
Voltage at Max Point (VM P P ) 35.4 V
Reverse Saturation Current (Irs ) 13.0 µA
Active Cell Area 1.12 m2
Diode Quality Factor (n) 1.86
Number of Cells in Series(N) 72

Using this data, it is possible to use equations (6), (7), and (8) to determine the maxi-
ripple
mum allowable ripple voltage, Vmax , for a given utilization ratio, kP V (Figure 6). As the
ripple
utilization ratio increases, the allowable Vmax decreases rapidly. This means that V ripple
must be limited to allow for optimal efficiency from the PV module.

V ripple is much more deleterious for larger PV module systems due to the larger value
of VM P P . Figure 7 shows the maximum allowable value V ripple as a function of VM P P for a
constant utilization ratio of 0.9. As value of VM P P increases, the maximum allowable value of
V ripple decrease rapidly until becoming steady at VM P P values above ∼100V. For PV module
setups with large VM P P values, it is of paramount importance to limit V ripple in order to
maximum the utilization ratio of the PV module.

In order to smooth V ripple across the terminals of the PV module and decrease its delete-
rious effects, a capacitive element is usually placed in parallel with the PV module. For a PV
module/inverter system injecting power at the module MPP according to (1), the nominal
applied voltage will be VM P P with a nominal output current of IM P P . The current through
the capacitor, Icap may be described by (9) [24].

21
6#$&"7"*2&''()*!0(8#9)*:;<*-.(&/#.01*2#.0*=0%*6010>%?;8#((&1)*
%#"
@&*60A7()*B5CDEF*

%!"

!"#$%&''()*+!,* $#"

$!"

#"

!"
!" !&$" !&%" !&'" !&(" !&#" !&)" !&*" !&+" !&," $"
-.(&/#.01*2#.03*45!*

ripple
Figure 6: Graph of maximum allowable V ripple vs. Utilization ratio. An increase in Vmax
decreases the Utliization ratio of the solar cell module. Therefore it is important to limit V ripple .

35  
Allowed  Maximum  Ripple  Voltage  vs.  VMPP  for  a  0.9  
U>liza>on  Ra>o  
30  

25  
Vmaxripple  (V)  

20  

15  

10  

5  

0  
0   50   100   150   200   250   300  
VMPP  (V)  

Figure 7: ]
Maximum allowed voltage ripple vs VM P P for a constant utilization ratio of 0.9. As VM P P
increases, the maximum allowed V ripple rapidly decreases. Therefore it is especially
important to decrease V ripple for PV modules with large VM P P values in order to maximize
the module utilization ratio.

22
Icap (t) = IM P P · 1 − 2sin2 (ωt)

(9)

The voltage across this capacitor can be described by (10) [25].

Z
1
Vcap (t) = Icap dt + V (0) (10)
C

By substituting (9) into (10) and choosing V(0)=VM P P , then Vcap (t) becomes (11).

Zt
IM P P
1 − 2 sin2 (ωt) dt + VM P P
 
Vcap (t) =
C
0
 t  t !
IM P P t sin 2ωt
= t − 2 − + VM P P
C 0 2 4ω 0
 t
IM P P sin 2ωt
= + VM P P (11)
C 4ω 0

The voltage across the capacitor reaches a maximum when ωt = π4 . This maximum
ripple
voltage is equal to VM P P +Vmax . The minimum necessary capacitance to smooth a given
π
V ripple can be found by substituting t = 4ω into (11) to be (12).

max ripple
Vcap = VM P P + Vmax
 t= π
IM P P sin 2ωt 4ω
≥ + VM P P (12)
C 4ω 0

The minimum capacitor size can then be found by solving for C to be (13).

PM P P
C≥ ripple
(13)
2 · ω · VM P P · Vmax

ripple
The size of the capacitor is a function of the grid frequency, PM P P , VM P P , and Vmax .
The capacitance can also be found in terms of the utilization ratio by substituting (13) into
(6) to get (14).

23
20  
Minimum  Capacitor  Size  for  Given  U6liza6on  Ra6o  
18  
 for  BP4160    

Minimum  Required  Capacitor  (mF)  


16  

14  

12  

10  

8  

6  

4  

2  

0  
0   0.1   0.2   0.3   0.4   0.5   0.6   0.7   0.8   0.9   1  
U6liza6on  Ra6o,  kPV  

Figure 8: Minimum capacitor size for given utilization ratio for BP4160 monocrystalline solar
cell. The size of the required capacitor increases rapidly with an increase in utilization ratio due
to voltage ripple.

PM P P
C≥ q (14)
2·(kP V −1)·PM P P
2 · ω · VM P P 3·α·VM P P +β

Figure 8 shows how the minimum capacitor size varies by the value of the desired uti-
lization ratio for a module with parameters listed in Table 2. As the value of desired kP V
increases, the necessary capacitance size also increases due to the need for more energy
buffering to decrease V ripple .

For relatively modest utilization ratios, in the 0.8 to 0.9 range, the required capacitance
to limit V ripple can be large, upwards of 6 or 7mF. It is infeasible (both physically and
economically) to put such a large capacitance in an inverter using a single capacitor. To
increase the effective capacitance into the milifarad range, large banks of smaller capacitors
are connected in parallel. This increases the available capacitance by (15).

n
X
Ceq = Ci (15)
i=1

where:
Ceq is equivalent capacitance
Ci is the capacitance of the ith capacitor

24
-.(/(0+1.0"
#" !"
#" !"
#" !"
#" !"
!"
#" 5 !" #"
#" !"
#" !"
$%&'(" #" !" )*+,&'("
#" '" !"
#" !"

2(+*/"3/*+(4"
Figure 9: Physical schematic of a generic capacitor. A capacitor consists of a dielectric sandwiched
between two electrode plates separated by a distance (d). When a voltage is applied to the plates,
~ inside the dielectric.
the dielectric polarizes, forming an electric field (E)

This practice makes inverters far more affordable, however, since many small components
are more likely to fail than one large one, the added complexity and number of parts has
detrimental effects on the reliability of the system.

Capacitors

A capacitor, one of the fundamental building blocks of linear circuits, is an energy storage
~ due to the alignment of dipoles inside a
device which stores energy in an electric field (E)
dielectric [26]. A generallized capacitor is composed of a resistive dielectric sandwiched
between two conductive metal plates (Figure 9). When a voltage difference is applied to
the anode and cathode of the capacitor, the dielectric does not allowed current to flow.
The capacitor uses energy by inducing opposite charges on opposing plates, creating an
electric field inside the dielectric. When the charge on the plates (Q) reaches a maximum
value (Qmax ), current flows through the capacitor and it discharges. This time delay in
discharging means the capacitor is a reactive element and, for an AC signal, the phase of
the current lags the applied voltage by 90o .

The ratio of the charge on the capacitor, Q, to the voltage drop, V, is known as the
capacitance:

Q
C= (16)
V

Capacitance can be solved exactly for an infinite flat plate capacitor using Gauss’ Law
to be (17).

25
0 r A
C= (17)
d

where:
0 is the permittivity of vacuum, 8.854×10−12 m
F

r is the dielectric relative permittivity com-


pared to vacuum
A is the area of the plates
d is the separation between conductive plates

The relative permittivity (Table 3) of a dielectric can range from 1 (vacuum) to around
300,000 for Calcium Copper Titanate (CaCu3 Ti4 O12 ).

Table 3: Table of relative permittivity, r and dielectric strength, Eds , for commonly used di-
electrics [27].

Material r Eds ( MmV )


vacuum 1 20-40 [28]
air 1.00058986 [29] 0.97
PTFE 2.1 60-173
polypropylene 2.2 28 [30]
mylar 3.0 600 [31]
mica 6.0 118
Al2 O3 8.5 13.4
BaTiO3 2000-10,000 [32] 100 [33]
CaCu3 Ti4 O12 (CCTO) ∼300,000 [34] <1 [35]

The energy stored in the electric field is determined by (18)

Z Q Z Q
Q
E= V dq = dq
0 0 C
Q2
=
2C
1
= CV 2 (18)
2

Ideally, a capacitor would prevent current from flowing during charging for all applied
voltages and internal electric fields. However, at large enough voltages, the dielectric ceases

26
to prevent current flow and becomes conductive. This applied voltage is known as the
breakdown voltage (Vbreakdown ) and is the product of the distance between the capacitor
plates (d) and the dielectric strength (Eds ) (19) [36].

Vbreakdown = Eds · d (19)

The value of Eds varies by dielectric and is usually presented for an ideal flat plate
capacitor It is a function of pressure, humidity and temperature [37]. The electric field
inside a real capacitor is critically dependent on the geometry of the plates and any sharp
edges or points which can magnify the electric field due to the small radius of curvature and
cause local failure. Once the dielectric fails locally, the breakdown quickly propagates from
plate to plate, converting the dielectric into conductive carbon and shorting the capacitor
[38].

Actual failure of the dielectric can occur from a number of methods. In an intrinsic
breakdown, the electric field may become large enough to ionize dielectric atoms, thus making
them conductive. Impurities or inhomogeneities in the dielectric may facilitate current flow
so the actual Vbreakdown is less than the ideal. If the dielectric is crystalline, the electric field
can ionize an atom and accelerate the free electron to velocities large enough to ionize other
dielectric atoms in an avalanche ionization process.

The breakdown voltage limits both the operating voltage of the capacitor for a given
thickness of dielectric and the amount of energy which can be stored in the electric field
of the capacitor. Since the breakdown voltage is proportional to d while the capacitance is
inversely proportional to d, the dielectric layer in a capacitor is subject to a classic thick-thin
conundrum. As the dielectric thickness increases, the maximum operation voltage increases,
but the capacitance decreases. Therefore, the area of the plates, A, must be increased
to increase the capacitance. This increases the physical size and corresponding cost of the
capacitor. The dielectric thickness must be carefully balanced for the operational use so that
it is thick enough to withstand operation voltages while thin enough to provide necessary
capacitance and physically small enough to be economically and physical feasible.

An ideal capacitor prevents charge from flowing through the capacitor during charging
and then discharges the energy located in the electric field with no resistance. For a real
capacitor, there are nonideal resistances, capacitances, and inductances. Nonidealities in
real capacitors are represented by an equivalent circuit (Figure 10). In addition to the ideal
capacitance, the dielectric is not perfectly resistive while charging and a small leakage current
(Ileakage ) flows from the anode to cathode during charging. In the equivalent circuit, this
non-infinite resistance is respresented by the resistor Rsp wired in parallel to the dielectric ca-
pacitance (C2 ). When current through the capacitor is flowing, the pins, foils, and dielectric
are not perfectly conducting and there is an equivalent series resistance (ESR) in series with
the capacitance. The terminals and tabs are also not perfectly resistive and posses a para-
sitic capacitance (C1 ). Finally, the capacitor does store some energy in the magnetic field,

27
!"

$2"
,-./0/1."
&'(" $%"

#" )*+" )*&"

Figure 10: Equivalent circuit of a generic capacitor. A capacitor is composed of many non-
ideal elements including capacitance of terminals (C1 ), dielectric capacitance (C2 ), leakage current
(Ileakage ) through the dielectric, series resistance (ESR), and series inductance (ESL).

so there is an equivalent series inductance (ESL) in series with the capacitance and ESR.
The ESL is quite small (∼nH [39]) and is largely ignored in circuit analysis of capacitors.

Types of capacitors

Electrolytic

A majority of capacitors used in PV inverters are electrolytic in nature. Electrolytic ca-


pacitors (elcaps) are composed of a electrolyte impregnated paper layer sandwiched between
two highly roughened metal foils (Figure 11) [40]. Grown directly onto the anode foil is a
thin, very high quality oxide which acts as the dielectric. A well-grown oxide layer possesses
excellent breakdown voltage characteristics (0.667-0.714 V/nm [41]). An electrolyte/foil sys-
tem acts as the cathode. The electrolyte is used as an electrode instead of a metal plate in
order to maintain the integrity of the oxide dielectric via ion current [42] and provide good
contact to the fragile oxide surface.

The manufacturing process for elcaps begins with a pure (∼99.99%, but at least >99%
[43]) metal foil (∼100µm thick) which is used as the anode foil precursor. For elcap, the
most common metal is aluminum. The foil must be relatively free of impurity elements (Bi,
Si, Mg, Zn, etc.) which can inhibit etching and decrease resultant capacity even at the parts
per million (ppm) level [44].

All aluminum forms a native oxide layer approximately 5nm thick when exposed to
oxygen [45]. This oxide is removed by dipping the foil in a hot 4-10% aqueous NaOH
solution [46]. After native oxide removal, the foil is roughened to increase the surface to mass
ratio by pulling the foil along rollers through a hydrochloric acid solution (HCl) at elevated
temperatures (typically <100o C) while applying a voltage (AC, DC, or both) between the
etch solution and the foil [47]. This etch forms many parallel hollow tunnels which are
perpendicular to the foil surface, increasing the area of the foil by as much as 100 times [43].

Small additions of CuCl, FeCl, or AlCl may be added to the HCl solution to increase

28
.%/0$'-'&(,+(1%*+(

,-%&'( !"#$%&'(
)%*+( )%*+(

8+'9#:%+4#'(;<=:'0-"#'&(
2*0$(3/"+*#4(,+567(
>"=':(?="9':(

Figure 11: Schematic of an aluminum elcap. The Al elcap consists of an electrolyte impregnated
paper spacer sandwiched by two highly roughened Al foils. On the anode foil, a high quality Al2 O3
layer is grown to act as the dielectric.

tunnel formation [48]. However, if Cu, Fe, Cl, or Al ions are included in the oxide film, the
leakage current can be drastically increased in the finished device [49], so the etched foil is
typically carefully washed by rinsing in deionized (DI) water and nitric acid (HNO3 ) [46].

Since the thin oxide layer is acting as the only dielectric in this system, the oxide quality
is of paramount importance to the functioning of the capacitor as a whole. This oxide is
typically grown (known as forming) via an anodizing method. Anodizing is an electrolytic
passivation process through which current flows from the anode, made of a valve metal
(aluminum, niobium, tantalum, titanium, or silicon), converting the metal into a metal
oxide (in the case of Al, into Al2 O3 ). For Al elcaps the anodization process is performed by
connecting the foil to the positive terminal of DC voltage source and rolling the foil into an
aqueous electrolyte bath. The negative terminal is connected to the cathode rod placed in
the electrolyte bath. This cathode can be made of any suitable material (C, Pb, Ni, stainless
steel, or other material which is unreactive in the bath [50]).

The electrolyte is chosen so that either the oxide is insoluble or the oxide dissolves into
the electrolyte more slowly than the oxide grows via the anodization process. The porosity
of the oxide film can be controlled by the pH of the bath. Neutral bath electrolytes (typically
ammonium borate, phosphate, or tartrate) cannot dissolve Al2 O3 and create a dense, barrier
oxide layer [51]. Acid electrolytes (typically 10% H2 SO4 , but also oxalic or phosphoric acids)
tend to dissolve some of the as-grown oxide layer and thus can create thicker, more porous
layers [52].

To grow the oxide from the metal foil, a voltage about 135-200% of the capacitor’s final
rated voltage is applied by the DC source [47]. When this voltage is applied, electrons are

29
Al   Al2O3   Electrolyte  

2Al3+   3O2-­‐  
2Al+3O2-­‐   Al2O3+6e-­‐   2Al3++3H 2O
  Al2O3+6H+  

growth   growth  

Original  metal/
electrolyte  interface  

Figure 12: In the anodization of Al to Al2 O3 , the oxide is grown via the diffusion of Al3+ and
O2– ions through the oxide layer. The Al3+ ions react at the oxide/electrolyte interface and the
O2– ions react at the oxide/metal interface. Since the oxide formation reaction is occurring at both
interfaces, the oxide grows both into and from the original surface of the metal.

withdrawn from the foil so metal ions can react with water to form oxide. The electrons are
returned to the bath at the cathode where the electrons react with hydrogen ions to create
H2 . Overall the reaction may be described by (20) [53].

Overall : 2 Al + 3 H2 O −→ Al2 O3 + 3 H2 (20)

The growth rate of oxide is determined by the flow of Al+ 3 ions from the metal and O2

ions from the electrolyte bath [52]. Since these ions are drifting from different directions,
new oxide is grown at both the metal/oxide and oxide/electrolyte interfaces (Figure 12). As
as result, the oxide is grown both out of and into the previous surface of the metal. The drift
of these ions determines the growth rate and is proportional to the current density (A/cm2 )
present in the system [53].

The anodization process add both mass and volume due to the incorporation of O2 into
the Al metal. The volume change of the oxide from the metal is determined by the Pilling-
Bedworth (PB) ratio [54], defined by (21).

Voxide
PB = (21)
Vmetal

The PB ratio not only determines the volume change from oxide formation, but also, for
many systems, whether the oxide is protective. For PB>2, the volume change is so large
that the oxide tends to flake off and offers no protection to the metal. For PB<1, the volume

30
do  

do   df  

Figure 13: (Left) SEM micrograph of a crossview of an Al foil after etching and anodization. The
Al metal has been removed, leaving the Al2 O3 dielectric. The etching of the Al foil has created
dense, parallel tunnels which have been partially filled with Al2 O3 [57]. (Right) Plan view schematic
of a single anode foil tunnel before (top) and after (bottom) anodization. The etching of the foil
forms tunnels with a diameter, do . After anodization, a thickness, t, of Al2 O3 has been grown.
The oxide has grown both into and out of the previous metal surface to occlude part of the tunnel
(diameter, df ).

change is such that the oxide cannot form a complete film on the metal surface. For PB
between 1 and 2, the oxide should form a complete, dense, protective barrier to gas diffusion
into the film. The PB ratio for Al to Al2 O3 is 1.28 [55], which indicates that the formation
of the oxide yields an increase in volume and is also a protective film.

The anodization process forms oxide from both interfaces (Al2 O3 /Al and Al2 O3 /electrolyte).
For the anode foils with etched tunnels, this means that the Al2 O3 layer will both grow
into the pore space and from the pore space into the metal (Figure 13). This yields an
oxide/metal interface with excellent adhesion and fewer defects sites that could affect the
electrical properties of the oxide dielectric [56].

The thickness of the oxide is extremely uniform and is determined by the voltage differ-
ence between the anode and cathode (1.4-1.5nm/V). Since the voltage drop across the oxide
must be constant everywhere, if there are any local non-uniformities in the thickness of the
oxide layer, the localized growth times will change to create a uniform film. The growth of
thinner regions will occur for longer durations while thicker regions will have shorter growth
times. Overall, the thickness of the film will be very uniform. This uniformity in oxide thick-
ness is the reason that elcaps can be operated at voltages much closer (∼80%) to Vbreakdown
than other types of capacitors [58].

The total thickness of the oxide layer is limited by the applied voltage. For a given bath

31
concentration and temperature, there exists a maximum voltage above which breakdown
occurs. During breakdown, reactions other than oxide formation occur, such as O2 gas
evolution, solute oxidation, or electron avalanche breakdown in the oxide. More dilute
solutions have a higher voltage breakdown threshold with the maximum level being ∼1000V
for aqueous solutions. This yields an oxide thickness of ∼1µm.

The density of tunnel formation in the Al foil and the thickness of the oxide must be
carefully controlled for operational voltages and capacitance. Since the oxide layer grows
inside the tunnels, the tunnels are partially occluded and the effective surface area of the
foil is decreased. Foils for low voltage applications should have small, dense tunnels with a
thin oxide layer while high voltage application foils should have larger tunnels with a thicker
oxide layer [47]. The diameter of the tunnels should be optimized for the necessary thickness
of the oxide. If the tunnels are too narrow, the oxide may clog the entirety of the tunnel. In
this case, the oxide will not add to foil capacitance. Tunnels which are too large in diameter
have oxide which does not fill a significant percentage of the space inside, so capacitance is
not maximized.

Once the oxide is grown, a paper spacer is dipped into hot electrolyte and vacuum/pressure
cycles are used to completely wet the spacer and anode [47]. The choice of electrolyte is very
important because the electrolyte must posses low series resistance, high dielectric constant,
high dielectric strength, and the ability to repair, heal, and isolate defects during voltage
operation. The specific electrolyte formulation varies from manufacturer to manufacturer
with most using some proprietary blend of materials closely guarded as trade secrets. In
general, electrolytes generally come in two classes: aqueous or nonaqueous.

Aqueous electrolytes have a larger percentage water content and usually consist of a
boric acid or sodium borate solution to act as an ionic conductor. These electrolytes are also
composed of various sugars, ethylene glycol, or dipropyle ketone to retard evaporation. Non-
aqueous electrolytes have a very small percentage water content and are usually composed
of a weak acid (usually organic or boric), salt of a weak acid (usually ammonium or other
metal), and a solvent. Both types of electrolytes may have any number of thickening agents
and other additives to alter pH (buffers), prevent H2 gas formation (degassers or gatters)
[59], or passivate Al and Al2 O3 surfaces [60].

The majority of series resistance in elcaps is due to the electrolyte solution. In order to
decrease the electrolyte resistance, capacitor companies have invested much effort into the
development of electrolytes with large percentages of H2 O (∼75-85%). The water content in
the electrolyte is key to oxide reformation. Under ideal circumstances, the leakage current
in the elcap drives a H2 O/Al reaction which reforms any compromised Al2 O3 (22).

Al(OH)3 −→ 2 AlO(OH) + 2 H2 O −→ Al2 O3 + 3 H2 O


2 Al + 6 H2 O −→ 2 Al(OH)3 + 3 H2

2 Al + 3 H2 O −→ Al2 O3 + 3 H2 (22)

32
This reaction can reform the oxide layer to lengthen capacitor lifetime, however, it con-
sumes H2 O while producing H2 . The consumption of H2 O leads to increased ESR while the
production of H2 leads to increased internal pressure. Overall, the presence of H2 O can yield
a higher quality dielectric layer with better electrical performance at the expense of high
temperature performance and decreased lifetime [47].

The addition of the cathode foil is the final piece to elcap construction. As mentioned
previously, all aluminum exposed to atmosphere will have a thin native oxide layer. This
native oxide has a capacitance. Since the cathode foil is in series with the dielectric oxide,
the equivalent capacitance is described by (23).

n
1 X 1
= (23)
Ceq i=1
Ci

For an anode capacitance (Ca ) and cathode capacitance (Cc ), the equivalent capacitance
(Ceq ) simplifies to (24). This indicates that the smaller capacitor in a series dominates the
equivalent capacitance.

1
Ceq = 1
+ C1c
Ca
Ca · Cc
= (24)
Ca + Cc

where:
Ca and anode capacitance
Cc is cathode capacitance
Ceq is overall equivalent capacitance

For very small capacitances, the numerator of (24) quickly approaches zero. Therefore,
a very large capacitance is desired for the anode. For a cathode capacitance fifty times the
anode foil, the drop in Ceq is only 2%. In order to increase the capacitance value of the
cathode foil, the foil undergoes no anodization process and may even go through several
processes to remove or thin the native Al2 O3 oxide.

In elcap topology, the dielectric oxide layer is present only on one of the electrode foils.
This results in an operational polarity. A voltage reversal applied to the foils drives an
electrochemical reaction which is the opposite of oxide growth. Namely, the oxide is reduced
back into aluminum metal. Once enough of the oxide has been reduced and the thinner oxide
layer can no longer electrically isolate the anode and cathode, the electrolytic capacitor will

33
short circuit. This short quickly heats the capacitor and the capacitor can will vent due
to the expansion of vaporized electrolyte. A reversed voltage of only 1-1.5V for times as
small as one second can vaporize the electrolyte enough to burst the capacitor can. The
anodization process can be carried out on the cathode foil in order to allow for AC operation
or DC voltage reversals (known as non-polarized or NP elcaps), but there is a subsequent
reduction in equivalent capacitance and energy density (due to increased mass).

Once the cathode foil is connected, the foils and spacer are wound into a a cylinder.
Electrical connections (Al tabs) are attached using cold-welding. The placement of tabs
is key to reduce both foil resistance (etching of the Al foils significantly decreases their
conductance) and inductance to below 2nH [47]. The capacitor is then wound into a cylinder.
After winding, the capacitor is placed in a can and sealed.

Since H2 gas is generated from the electrolyte, pressure release valves are placed in the
can and are usually designed to release interior pressure above ∼7 atm. The pressure relief
valve can either be in the form of a rubber bung or gasket which is pushed out by internal
pressure, or a die-slit on top. Blowout of the pressure valve is considered a failure mode
since it allows the electrolyte to completely evaporate away.

After being sealed in a can, capacitors are aged. In this process, the capacitors are con-
nected to a voltage and slowly brought up to the maximum rated voltage and temperature.
This allows for oxide growth via electrolyte healing in any areas which have insufficient oxide
(such as slit edges and winding cracks). This also allows for capacitors with manufacturing
defects to be identified and removed from shipments sent to consumers.

As was mentioned previously, anodization can be carried out on any valve metal. It is
possible to purchase elcaps made of other metals besides Al. There has been some market
penetration by Tantalum and Niobium capacitors. Ta elcaps have better electrical charac-
teristics and higher stability over time and temperature (400% increase in capacitance and
90% lower resistance). Niobium elcaps have similar capacitances to Ta capacitors. Ta and
Nb elcaps are not used at a larger scale mostly due to price. While Al costs around $0.04/g
for the anode material, Ta and Nb cost around $2/g and $1/g, respectively. Tantalum elcaps
are usually available only in small (<5cm3 ) capacitors while Nb elcaps are generally too new
to have widespread applications.

Al elcaps are mostly used in inverters because of their low cost and high capacitance
per volume, making them ideal for larger applications. Unfortunately, they tend to suffer
from high instability with temperature and time, high leakage current, and low operational
temperature [61].

Metallized Thin Film

The second most commonly used capacitor in inverter technologies is a class called met-
allized thin film capacitors. These capacitors consist of a dielectric plastic film 10-100µm
in thickness which is metallized on both sides to form electrodes [62]. The electrodes are

34
usually Al and Zn in pure or alloyed states [63], deposited via evaporation or physical vapor
deposition [64] in thickness of ∼20-100nm [65]. Two layers of film are wound together as
tightly as possible to avoid any air voids. Electrical connections are made by end spraying
(or schooping) a metal (typically Zn) the exposed edges of the roll [64]. The final electrical
connection is made by soldering a tab onto the end-sprayed metal [66].

A variety of plastic film dielectrics can be used in the capacitors, including polyester
(PET), polystyrene, mylar, polyimide, polycarbonate, and teflon [67]. The majority of high
performance capacitors utilize biaxially oriented polypropylene films as the dielectric due
to its low cost, low resistance [68], and highly consistent manufacturing [69]. Most of the
polymer films require surface treatment (e.g. coronal treatment) to polarize the nonpolar
surface and increase adhesion with the metal electrode layer [64].

The use of plastics in capacitors have been utilized for over 50 years [70] with polypropy-
lene film-foil capacitors available since 1970 [61]. Many early capacitors utilized thicker metal
foils as the electrodes. These types of capacitors had poor lifetime characteristics as local
imhomogeneities in the plastic dielectric would quickly lead to local, then global, failure.
Poor lifetime performance of these types of capacitors was improved by thinning the metal
electrodes to the nanometer range. The thinner metallization layer allowed for local failure
without global failure of the capacitor in a process known as self-healing or clearing (Figure
14).

All manufactured thin film capacitors contain some amount of defects. These defects
can range from air pockets to foreign matter to inconsistencies in the polymer film. The
Vbreakdown of the capacitor is lowered in the regions of these defects so that, during operation,
the capacitor will locally short at some voltage V< Vrated . The short quickly heats the
capacitor via Joule heating to such a degree that the thin metallized electrode layer is
evaporated. This locally isolates the short and creates a small pinhole ( ∼5-8mm2 [64]) in
the dielectric film. The capacitor suffers a small decrease in capacitance, but lifetime is
increased by avoiding catastrophic failure.

Clearing behavior should allow for the total avoidance of catastrophic failure of metallized
thin film capacitors. However, for very large localized defects, the short may be large enough
that the entire capacitor can catastrophically fail during operation [63]. To avoid this event,
modern metallized thin film capacitors use a patterned electrode layer [72, 73]. Larger
patches of metal electrode are separated by much thinner current gates [74]. For normal
clearing behavior, the current gates are unaffected. However, for very large failures, the
current gates should readily evaporate, isolating the failed metallized electrode segment
before it can affect neighboring segments [75].

In a completed capacitor, the films are wound as tightly as possible to prevent the for-
mation of air voids [76]. These voids are deleterious to capacitor lifetime because, as the
number of clearing event increases, conductive metallized vapor builds inside the capacitor
casing. This vapor may congregate in any air voids present in the package. After too many
clearing events, these voids can become conductive, which can cause glow discharge in the
voids and catastrophic failure of the capacitor [74]. If the films are wound with no im-

35
Dielectric  Film  
Defect  
Metalliza,on  
Local  Breakdown  

Current  Gates  
Self-­‐healing  

Metallized  
 Electrode  
Segments  
Current  Collector  
Dielectric  Film  

Figure 14: (Left) Self-healing or clearing of a metallized thin film capacitor [64]. Defects in
the manufacturing process decrease the local Vbreakdown of the polymer film and facilitates local
breakdown of the capacitor. In this region the capacitor shorts at V< Vrated . The short quickly
heats the capacitor to such a degree that the metallization layer is evaporated in the area of the
short leaving a small pinhole (∼5-8mm2 ) through the dielectric film. The evaporated electrode
layer prevents a short from forming. The capacitor has a small decrease in capacitance, but no
catastrophic failure. (Right) Fusing of a metallized thin film capacitor [71]. Although self-clearing
should not result in catastrophic failure of the capacitor, for very large localized defects, the shorting
may be large enough to cause failure of the entire capacitor. Modern thin film capacitors have
metallization layers which are segmented into larger electrode areas separated by thin metallized
current gates. These gates act as fuses which will fail before the failure of the entire capacitor.

pregnating substances, the capacitors are known as dry capacitors. Some capacitors may be
impregnated with oil or epoxy (or some other non-swelling, low viscosity, low surface tension,
highly oxygenated liquid [64, 77]) during winding which increases breakdown strength [78],
high voltage behavior, and overall lifetime [76].

Since the metallization layer is symmetrical on both sides of the dielectric, this type of
capacitor is nonpolar [66]. Overvoltage and voltage reversals are typically accepted without
affecting lifetime [58]. However, the metallization film structure limits peak current and
energy density handling. Thin film capacitors also tend to suffer faster degradation in ca-
pacitance at higher temperatures and have a relatively low operating temperature due to
phase transition of the polymer (e.g. 160-166o C for polypropylene, depending on percent
tacticity [79]). Typical maximum operating temperatures are around 125-150o C.

One of the primary conflicts regarding capacitors is whether metallized thin film or elcaps
are more appropriate for use in inverters. Even though metallized thin film capacitors are
safer with longer lifetimes (up to 150,000 hours compared to ∼40,000 for elcaps), electrolytic
capacitors are still used in most inverters due to cost and past history. While electrolytic

36
capacitors have been used successfully for many years and have well-defined lifetime calcu-
lations, much less is known about the long term use of metallized thin film capacitors.

In addition to longer history, electrolytic capacitors typically have a cost which is varies
linearly to the capacitance [24] (typically around $0.0045/µF [18]). The cost of thin film
capacitors increases exponentially with increasing capacitance [24], so in order to limit volt-
age ripple across the PV module terminals to a manageable voltage, the cost of thin film
capacitors may be prohibitive when looking purely at capacitance. However, the cost is not
so unbalanced between the two types of capacitors since electrolytic capacitors are typically
limited by their ripple current handling [18]. This means that inverters which utilize elcaps
usually have much more capacitance than this strictly necessary in order to handle the nec-
essary ripple current. The improved ripple current handling of thin film capacitors means
that less capacitance can be used, which can decrease the cost disparity somewhat. In order
to truly determine which type of capacitor is more economical, it is necessary to perform a
detailed lifecycle calculation considering not just the capacitor unit installation costs, but
lifetime, and replacement cost.

Ceramic

A third type of capacitor, multi-layer ceramic capacitors (MLCC), are currently not used
in high power electronics applications due to cost considerations. However, their price has
been rapidly decreasing and, in the near future, they may begin to be utilized in inverter
systems. For now, they are mostly used in low power consumer electronics.

MLCCs consist of repeating layers of ceramic sandwiched by metal electrodes (Figure


15). A wide range of ceramics may be used as the dielectric medium with the most popular
being variations of barium titanate (BaTiO3 ). The ceramic is usually created via sintering
of a “green tape”, a ceramic powder or slurry in a wax carrier [80]. Since the electrodes are
co-fired with the ceramic powder, the electrode is limited to materials with a higher melting
temperature than the sintering temperature of the ceramic (∼1100-1700K, depending on the
ceramic).

Metal electrodes have been traditionally composed of precious metals (usually a 85%
silver/15% palladium alloy) [81]. In order to decrease device cost in the face of rising pal-
ladium costs (palladium in automobile catalytic converters accounts for 50% of overall use
[82]), many manufacturers have been moving towards base metal electrodes (usually Ni with
Cu terminations) [80].

Although progress towards base metal electrode MLCC have been progressing rapidly,
there are still manufacturing issues to overcome. Since the Ni electrodes require a reducing
environment during sintering to avoid oxidation [83], oxygen vacancies are created in the
ceramic. It is generally believed that vacancy ordering along grain bounderies eventually
leads to a large, irreversible capacitance degradation in humid environments [84]. It has
also been postulated that due to humid environments, moisture may penetrate cracks in the

37
!"#$%&'(

)"*$+(,+"'*#-."/(

Figure 15: Schematic of a multilayer ceramic capacitor. The capacitor consists of a stack of metal
electrodes interspersed with a ceramic dielectric sintered from a powder [39].

MLCC and dissolve ionic impurities to form conductive bridges that decrease capacitance
[85].

Before MLCCs can compete with other capacitor technologies, a number of disadvantages
must be overcome. While ceramic capacitors have high operating temperatures with good
reliability, they suffer from poor temperature drift and stability. They also possess high
dielectric constants, but poor Vbreakdown due to any remaining porosity after the sintering
process. Finally, although they have high volumetric capacitance, they are currently too
costly to be used for high power electronic systems.

Capacitor Failure Modes

Capacitor end-of-life can be described by two basic categories of failure, catastrophic and
degradation [86]. Catastrophic failure is complete loss of functionality of the capacitor in
the circuit. These failures can result in an open or short circuit, explosion of the capacitor
can, destruction of the dielectric, harm to other electrical components, or leakage of liquids
or gasses from the capacitor interior. Degradation failures are failures of definition where
a capacitor parameter has fallen outside of “acceptable” performance limits. These limits
and parameters vary widely by manufacturer and may include a leakage current increase,
ESR increase (typically 300% for electrolytic and 150% for metallized thin film capacitors),
dielectric value decrease (typically 40% for electrolytics), or capacitance decrease (typically
20% for electrolytics and 5-10% for metallized thin film [39, 87]. Electrolytic capacitors
tend to fail catastrophically while metallized thin film capacitors tend towards degradation
failures.

38
Electrolytic

Electrolytic capacitors have two primary modes of failure, physical and chemical. The
primary physical failure mechanism for elcaps is vaporization of the electrolyte through
polymer seals. A decrease in the volume of the electrolyte leads to an increase in the capacitor
ESR [88]. Vaporization of the electrolyte may be driven by increased internal temperatures
due to ripple current or simple evaporation through poor quality sealing. Many researchers
cite electrolyte vaporization as the primary wear out mechanism for elcaps [88, 42].

Chemical failure of elcaps has been mentioned previously. Leakage current through the
electrolyte drives the reformation of Al2 O3 through the consumption of H2 O according to
(22). This reaction leads to three methods of failure. The first is that reformed oxide is of
a poorer quality than the oxide formed via anodization. The replacement of high quality
oxide with poor quality oxide decreases the capacitance of the elcap over time. Secondly,
the consumption of H2 O leads to higher ESR values from the decreased electrolyte volume.
Finally, in addition to the oxide, one of the products of the reforming reaction is H2 gas. Since
the capacitor can is airtight to prevent evaporation of the electrolyte, H2 gas accumulates
inside the canister over time, increasing the interior pressure. Once the interior pressure
builds to ∼7 atm, the pressure relief valves release, bursting the capacitor can. After the
can is exposed to air, the electrolyte readily evaporates and the capacitor fails.

Metallized Thin Film

With the advent of segmented electrode films, metallized thin film capacitors seldom fail
catastrophically. Over time, metallized thin film capacitors suffer a gradual degradation
until a parameter falls below specification. This is known as soft failure [65].

There exist two categories of degradation in metallized thin film capacitors [64]. The
first, and most common [39], is due to clearing behavior. For each clearing event, there is a
small decrease in the overall capacitance of the capacitor. With enough clearing events, the
capacitor eventually fails in an open circuit. This failure is much more advantageous from
a safety aspect compared to elcaps, which tend to short circuit and can extensively damage
the capacitor itself or other electrical components [66, 89].

During a clearing event, joule heating from the localized discharge vaporizes a certain
volume of metallized electrode (the vaporized volume is proportional to applied voltage [90]).
This discharge creates a localized plasma that interacts with the polymer surface, creating
low molecular weight species from the degradation of polymer. Polymers which are rich in
oxygen tend to form CO upon vaporization. Polymers which have a relatively low oxygen
composition tend to form conductive graphitic filaments which are more likely to bridge
opposing electrodes and form shorts [64]. Low oxygen component polymers are said to be
poor clearing polymers and are more likely to clear incorrectly, shorting the electrodes and
catastrophically failing.

39
Even for polymers with good clearing behavior, the formation of gas inside the capacitor
can lead to problems. As more clearing events occur, the production of gas increases pres-
sure inside the casing. This increase in pressure tends to increase the rate of degradation
[39] via surface erosion, tracking, and polymer decomposition. This means that localized
failure in one area leaves neighboring areas ripe for further electrical breakdown [64]. Excess
gas can also lead to can bulging and pressure release [91] (though this usually occurs only
after degradation failure has already occurred). Before pressure release, the gas may be
dense enough to conduct charge between electrodes in a process known as flashover (also a
catastrophic failure). To prevent this, manufacturers may choose to fill metallized thin film
capacitor windings with a liquid (liquid impregnated), instead of air (dry-wound) [78].

The second degradation mechanism is due to direct electron impacts on polymer segments
at the electrode/dielectric interface. The impact creates free radicals in the polymer which,
in turn, reacts with absorbed H2 O or O2 . This interaction causes a chain reaction which
can cause scission or crosslinking of polymer chains, deteriorating the initially good physical
properties of the film [64].

Ceramic

Ceramic capacitors can suffer from either catastrophic or degradation breakdown. Degra-
dation breakdown tends to be easily controlled and, in most cases, is reversible by heating
above the Curie Temperature for one to four hours [81]. This aging is due to the ferroelectric
and volumetric response of the ceramic upon cooling from the Curie Temperature.

Catastrophic failure in ceramic capacitors tend to result in crack propagation through


the dielectric ceramic during breakdown [61]. There are three modes of dielectric breakdown
in ceramic capacitors, intrinsic, thermal, and ionization.

The most common cause of failure is intrinsic breakdown. This type of breakdown occurs
when the applied voltage is larger than the ceramic material can support. The intrinsic
strength of the ceramic depends on the ceramic porosity, grain size, conductive impurities
[85], manufacturing methods, past applied voltage [92], applied voltage frequency, voltage
ramp rate [93], and any mechanical defects present. Ceramics tend to crack under high fields
[32]. This cracking is facilitated by any strain due to clamping or electrode connections to
the ceramic [94]. Because ceramics are brittle materials, any small cracking quickly grows
and propagates throughout the entire dielectric.

Catastrophic breakdown of an MLCC dielectric may also occur due to thermal runaway.
In this failure mode, a localized portion of the ceramic dielectric is chemically reduced.
This new reduced ceramic material has a greater conductivity allowing more current to flow.
Through Joule heating, the temperature is rapidly increased, converting more ceramic to the
chemically reduced form and increasing the conductivity (and amount of current) even more.
This loop continues until the temperature is high enough that the ceramic melts and, either
a conduction path shorts the capacitor [95], or the CTE difference between the ceramic and
electrodes causes delamination and cracking [96].

40
The final failure mode for ceramic capacitors is ionization breakdown. In ionization
breakdown, current flows due to Pool-Frenkel emission [97]. In Pool-Frenkel emission, an
electric field lowers the necessary thermal energy for a bound electron to hop into the con-
duction band from localized states. Porosity and other structural features in the ceramic can
enhance the local internal electric field [98] so that it is far greater than the average applied
electric field [85]. This enhanced electric field causes electrons to have posses enough energy
that they can knock multiple neighboring electrons into the conduction band in an avalanche
ionization [85, 99]. This local ehancement of the electric fields leads to breakdown at lower
Vapplied than would otherwise occur from intrinsic breakdown [100]. The electric field is en-
hanced by structural features so, the dielectric strength of the dielectric in a MLCC varies
with ceramic porosity [101, 102] and defect density (e.g. grain boundaries or voids).

Lifetime Models

Electrolytic

Most elcaps have rated lifetimes anywhere from 1,000-15,000 hours (0.11-1.71 years)
[103], though these base lifetimes may be accelerated by increases in temperature, applied
voltage, and ripple current. The acceleration rates differ from manufacturer to manufacturer
due to different electrolyte chemistries and capacitor topologies. Most acceleration rates and
formulas are empirically derived to fit experimental data, rather than derived from underlying
theory. So care must be taken when applying acceleration rates from one manufacturer to
another, since capacitor construction and testing methods may differ widely.

Typical lifetime testing for many manufacturers occurs at the maximum rated voltage,
however, manufacturers may publish lifetimes for capacitors at temperatures other than
those tested. Most researchers and manufacturers assume that the lifetime of a capacitor is
affected by temperature according to the Arrhenius equation [104]. The Arrhenius equation
is a very generic equation used to describe the temperature relationship with a reaction rate,
k (25).

−Ea
k = Ae kB T (25)

where:
A is a constant prefactor
Ea is the reaction activation energy,
kB is Boltzmann’s constant (1.38 × 10−23 KJ )
T is the absolute temperature in Kelvin

For elcap manufacturers, typically the degradation rate (or, more specifically, the lifetime
L1 ) of a capacitor is known at a certain temperature, T1 . To find the lifetime, L2 at some
elevated temperature T2 , (26) is used.

41
−Ea
L1 Ae kB T1
= −Ea
L2 Ae kB T1
Ea
− k EaT
= e kB T2 B 1
h i
Ea 1
− T1
=e kB T2 1
h i
Ea ∆T
=e kB T1 T2
(26)

The reduction/oxidation of anodic alumina has a value in the literature of approximately


1.506 × 10−19 J (0.94eV) [42]. If it is further assumed that the temperature of the capacitor
is around the maximum operating temperature (T1 · T2 ≈ 3982 ), then an approximation of
the new lifetime can be made by (27) [105].

L1 1.091×105 ∆T
= e 3982 ( 10 )
L2
∆T
≈ eln 2 10
∆T
≈ 2 10 (27)

The equation presented in (27) is a simple “rule of thumb” for elcap lifetime in the
industry and is known as the so-called “rule of two”. This rule states that the lifetime of
a capacitor will halve every 10o C above rated operation temperature (or will double every
10o C below rated temperature).

Since the value of two is only approximate, care must be taken when applying the “rule
of two”. Equation (27) assumes that the product of the rated and operation temperature is
≈ 3982 . This may not be true and, for low temperature, underestimates the lifetime by 10-
30% [106]. The calculated lifetime may also underestimate the lifetime because the equation
assumes a simplified constant voltage. Any ripple current present gradually increases the
internal temperature of the capacitor so that the ambient operational temperature, capacitor
surface temperature, and internal core temperature may all be significantly different [88].

For degradation failures of elcaps, it is generally assumed that the evaporation of solvent
in the electrolyte is responsible for decreased performance parameters. The vapor pressure of
ethylene glycol, a component in many electrolyte recipes, can change multiple orders of mag-
nitude over standard capacitor operation temperatures. Depending on the temperature and
the quality of capacitor construction, the solvent may readily evaporate at higher tempera-
tures. This evaporation decreases electrolyte volume and increases the elcap ESR. Failure
is generally considered to have occurred for ESR values 300% of the initial ESR value. In
1984 Rhoads and Smith [107] determined that, over time, the ESR could be described by a
simple linear inverse model (28).

42
ESR 1
= −Ea (28)
ESRo 1−k·t·e T

where:
t is the operation time in hours
Ea is an activation energy (≈4700) based on
experimental fit
k is an empirical constant depending on the size,
design, and construction of the capacitor [108].
It is usually (though not always) reported by
manufacturers. An average value found by ex-
perimentation would be around 1.77 [109, 110]

Figure 16 shows the percent change in ESR over time according the linear inverse model
(28). The expected lifetime of capacitors using this model is quite good (∼50,000 hours),
especially since failure is by degradation, so the capacitor continues to function, albeit in a
reduced capacity. However, this offers an ideal model for capacitor lifetime and does not take
into account any effects of ripple current which may significantly alter the temperature of
the capacitor. The linear inverse model also does not take into account how the evaporation
affects the capacitor integrity. The buildup of pressure inside the capacitor can may lead to
venting of gas and ensuing leakage of liquid electrolyte long before the capacitor fails in a
degradation mode.

Generic lifetime equations differ from manufacturer to manufacturer and take into ac-
count both theoretical and empirical relationships depending on capacitor construction and
design. In general, capacitor lifetime is affected by operating temperature, applied voltage,
and ripple current. Most lifetime equations taking these parameters into account and, al-
though the specifics are different, they are generally structured in the form of (29) [40]. Most
of the lifetime acceleration factors are based on empirical models derived from experimental
data of capacitor operation. Therefore, the formulations, values, and definitions of the accel-
eration constants may differ widely depending on the type of capacitor tested, manufacturer
or researcher, or experimental setup.

L(T, I, V ) = Lo KT KR KV (29)

As mentioned previously, ripple current present in a circuit results in gradual internal


heating of the capacitor which is not accounted from by KT . A high internal temperature
decreases ESR (due to an increase in ion mobility) and promotes vaporization of electrolyte

43
ESR  Linear  Inverse  Model  
350  

300  

250  

ESR/ESRo  (%)    
200  

150  

100  

50  

0  
0   10,000   20,000   30,000   40,000   50,000   60,000  
Time  (hr)  

Figure 16: Prediction of ESR at time t versus original ESR using the Linear Inverse Model. This
model assumes ESR decreases due to evaporation or vaporization of electrolyte. However it does
not take into account any effects of ripple current or possible catastrophic capacitor failure due to
pressure venting.

where:
L is the capacitor lifetime as a function of tem-
perature, voltage, and Iripple
Lo is the base lifetime measured by the manu-
facturer
KT is an operational temperature acceleration
factor determined by (27)
KR is the ripple current acceleration factor
KV is the applied voltage acceleration factor

material leading to smaller capacitor lifetimes. The ripple current acceleration factor, KR
is meant to account for this increased internal temperature due to ripple current [47]. One
possible acceleration factor due to ripple current is published by capacitor manufacturer
Cornell-Dublier [47] (30).

 1/2
Tc − Ta
KR = (30)
Tc − Tr

Since the capacitor is not at a steady state temperature, there are three separate applica-
ble temperatures, the ambient temperature, Ta , the capacitor surface temperature, Ts , and
the capacitor core temperature, Tc . The ambient and surface temperatures are relatively
easily measured. Unfortunately, while Tc is the most important temperature, it may difficult

44
generation.
Manufacturers often provide generic curves showing
1 the effect of temperature on ESR for their products.
Gap = 1 + R I +R, -- j 1)
2n fC1 These curves can be used with the ripple current
-+ j2z f C2
R, L
multipliers to obtain an estimate of the values for the
Where: where: T c is the components
core temperature in the model. However, a more
Zcap Complex Impedance of T Capacitor accurate alternative is to experimentally measure the
a is the ambient temperature
ESR and capacitance over a range of temperatures
Ro Resistance of Foil, Tabs, Terminals
Tr& is (Q)
the rated temperature
and frequencies and use least squares to fit the
RI Resistance of Electrolyte (Q)
values to the data. Fig. 3 and 4 illustrate the good fit
R2 Dielectric Loss Resistance (Q) achieved by this process and Table 1 shows the
C1 Terminal Capacitance (F) values for an actual capacitor
Cp Dielectric Loss Capacitance (F)
200%
f Frequency (Hz)
180%
I
w 120Hz
I
I
I
-
Equation
ESR = Real(Z,,,) (2)

Real(ZCa,) = R2
+ R I +R, (3)
1 + (2n f)2C: RZ

Increasing temperature causes a decrease in ESR


due
ytic Capacitor Model to the increased conductivity of electrolyte R1.
High rated voltage devices have more sensitivity to
temperature than low rated
allel resistor capacitor 100voltage devices lo00 because 10000 25 50 I5 100
to both ESR they and
use lower conductivity electrolytes[l 1. Equation
Frequency (Hz) Temperature ("C)
141. However, (4) captures
one the temperature
Fig. 3. Effect sensitivity
of Frequencynormalized
on ESR Fig. 4. Effect of Temperature on ESR
to
for estimating heat Figure
the base 17: ESR
resistance at of a
room dielectric
temperature
Actual at 50°C and Fit using
capacitor
Rlbase. versus
It
Equation 3
frequency (left) and temperature (right)[112]. Ex-
Actual at 120Hz and fit using Equation 4
is common for the
perimental elevated
points aretemperature
compared to ESR to be
equations (32)-(37). ESR decreases with increasing frequency
half the room temperature
because value.
of dipole relaxation time in the dielectric oxide1. and
Table Values for 400V Rated
increasing Voltage 470uF
temperature because elec-
Manufacturers often provide generic curves showing
R I +R, -- j trolyte
1) the effect
viscosity of temperature
decreases, on ESR
allowing for
for their
greaterproducts.
ion mobility.
2n fC1 These curves can be used with the (4) ripple current
Where: multipliers to obtain an estimate of the values for the
Rlbase RI at components in the Tbase
Base Temperature model.(Q) However, a Sensitivitv more Factor E
Capacitor e
T
,, accurate
Core alternative
Temperature ('
K ) is to experimentally
to obtain direct data while the capacitor is in measure the
I Dielectric
operation. Loss For large
Resistor I 0.038 manufacturers
R:, capacitors, SZ I
s, & TerminalsTbase ESR
Base and capacitance
MeaSU~Wnent Temp over a range of temperatures
=23'C=300eK I
~

(Q)may offer units with embedded thermocouples. SeriesThese


Capacitor C1
special samples 470vFbe run at full
may
E and frequencies
Temperature SensitivityandFactor (OK-') squares to fit the
use least
e (Q) voltage andtoripple forFig.
prototyping. However, Loss Capacitor
Dielectricmay C2 I 11,000 vF
values the data. 3 and 4 illustrate the good this
fit not be feasible for smaller capaci-
nce (Q) achieved by this process and Table 1 shows the
F)
tors and many researchers and manufacturers either substitute the surface temperature or
values for an actual capacitor 1043
ance (F) approximate the core temperature from surface temperature measurements.
200%
The core180%temperature
I
w inside
I
- I
120Hz theEquation
I
capacitor due to ripple current can be calculated from the
(2)
temperature rise, ∆T , due to a power (P) flowing across an element with thermal resistance of
Rth . The electrical power is dissipated across the capacitor due to resistive heating (P = I 2 R)
+ R I +R, and
(3) the temperature rise can be calculated by (31).
RZ

a decrease in ESR
vity of electrolyte R1.
ve more sensitivity to
age devices because 25 50 Tc = IT 5a
+ ∆T100
ctrolytes[l 1. Equation Temperature ("C)
sensitivity normalized = Ta + P · Rth
Fig. 4. Effect of Temperature on ESR
emperature Rlbase. It Actual at 120Hz and fit = Ta +
using
2
IripRM
Equation 4S · ESR(f, Tc ) · Rth (31)
mperature ESR to be
. Table 1. Values for 400V Rated Voltage 470uF

(4)
The ESR of an elcap is not constant, but is a function of both internal capacitor tem-
e Tbase (Q) peratureSensitivitv
and frequency
Factor E (Figure 17). The ESR can be described by (32) and is composed of
I Dielectric LossRResistor
three components, c , Rf , R:
and
, RIT [111].
0.038 SZ I
mp =23'C=300eK ~

Series Capacitor C1 I 470vF


y Factor (OK-')
Dielectric Loss Capacitor C2 I 11,000 vF
45
1043
ESR(f, Tc ) = Re(Zcapacitor )
!
1 i
= Re 1 + Rc + RT (Tc ) −
Rf
+ i2πf C2 2πf C1
Rf
= + Rc + RT (Tc ) (32)
1 + (2πf C2 Rf )2

where:
Rf is the oxide dielectric loss is ≈0.038 for an-
odic Al2 O3 [112]
f is frequency
C1 is parasitic terminal capacitance
C2 is primary capacitance
RT is temperature dependent resistance of the
electrolyte
Rc is constant resistance of tabs, foils, and con-
nectors and is ∼10mΩ

Rc is a constant resistance due to the foils, tabs, electrolyte, and other capacitor compo-
nents. This resistance is a relatively small value (∼10mΩ) which is constant with temperature
and frequency. Rf and RT are the frequency and temperature dependent resistances of the
leakage resistance, Rsp in Figure 10. The frequency dependent resistance Rf is due to the
the alignment of dipoles in the dielectric oxide during voltage reversals [113].

The temperature dependence of ESR contained in RT is due to viscosity change of the


electrolyte. This change in viscosity changes ion mobility and electrolyte resistance and can
be modeled by (33) [42, 113].

Tc −25 B
RT (Tc ) = RT o · 2−[ A ] (33)

where:
RT o is the measured electrolyte resistance at
25o C
Tc is core temperature (o C)
A and B are best fits and, for ethylene glycol-
based electrolytes
are typically 40 and 0.6, respectively [113]

The final component needed to calculate the ripple voltage acceleration factor is the ther-
mal resistance (Rth ) of the capacitor. The value of Rth determines how quickly a capacitor

46
+,&-.(

!"#$%&'"#(
/"*#0(

!"#)*&'"#(
Figure 18: The internal temperature of a capacitor depends on three modes of heat dissipation,
convection, radiation, and conduction. Convection is heat dissipated through the movement of
surround air. Radiation is energy dissipated through emission of thermal radiation. Conduction is
energy dissipated through contact with solids [40].

may dissipate heat to lower Tc towards Ta . Rth is simply the reciprocal of the total heat
transfer coefficient, htot , and the surface area (A) of the capacitor (34).

1
Rth = (34)
htot · A

where:
W
htot is the total heat transfer coefficient ( cm 2)

A is the surface area of the capacitor (cm2 )

The heat transfer coefficient is composed of three different heat dissipation modes, con-
duction, convection, and radiation (35) shown in Figure 18.

htot (W/m2 K) = hconv + hrad + hcond (35)

Convection is the transfer of heat from the capacitor to a liquid (usually air, though
possibly some other cooling liquid). Convection cooling can be described in two regimes.
For free air convection (v<0.5m/s), the heat transfer coefficient can be described by the
temperature difference between the surface and ambient, the number of capacitors (m), the
diameter of a single capacitor, (D), and a constant (G) (36)[114]. For forced air cooling (i.e.

47
v>0.5m/s), convection is the dominant cooling mode and can be modeled using only the
capacitor surface area (A) and the air velocity, v [14].

  s −Ta 1/4
 G · Tm·D
 v . 0.5 m/s
hconv = (36)
 500A−7/8 (v + 1)−2/3

v > 0.5 m/s

where:
G is a constant equal to 1.32 for single capaci-
tors and 1.2 for banks
Ts and Ta are the surface and ambient temper-
ature, respectively
m is the number of capacitors
D is the diameter of a single capacitor
v is the forced air velocity

Radiation is heat transfer from the capacitor through the emission of electromagnetic
(i.e. infrared, λ >740 nm) radiation. Electromagnetic radiation emission can be nicely
modeled for a capacitor. Assuming that the capacitor acts as a grey body emitter, the heat
transfer coefficient can be calculated using the Stefan-Boltzmann law [115]. This equation,
shown in (37), depends on the capacitor emissivity (percentage of energy absorbed, ) and
the Stefan-Boltzmann constant (σ). This heat transfer mode is typically the least important
and is often ignored.

hrad = σA Ts4 − Ta4




(37)

where:
A is the surface area of the capacitor or capac-
itor bank
Ts and Ta are the surface and ambient temper-
atures, respectively
 is the capacitor emissivity (≈0.4 for bare and
≈0.85 for sleeved elcaps)
σ is the Stefan-Boltzmann constant; σ =
π 2 kB
4

60~3 c2
= 5.67 × 10−8 s·mJ2 ·K 4

Conduction is the transfer of heat through contact with a solid material in physical
contact with the capacitor. Conduction usually occurs primarily through the capacitor

48
connection to the exterior mount or a heat sink. This heat transfer mode is usually only
a significant portion of the total heat dissipation for small, axial elcaps, elcaps with a heat
sink attachment, or liquid cooled capacitors. For high power electronics, heat conduction is
largely ignored.

Calculating the effect of ripple current on capacitor interior temperature requires cal-
culation of the frequency and temperature dependent ESR, thermal resistance (Rth ), and
convection and radiation heat transfer coefficients (hconvection and hradiation ). This calculation
can be quite complicated because the calculation of Tc requires calculating the internal ESR,
which is a factor of Tc . So a reductive loop is established when running calculations.

These strings of equations cannot be simultaneously solved analytically, so many re-


searchers and manufacturers either use estimations and simplifications or choose to numer-
ically integrate equations (31) to (37) [88]. This can be accomplished by choosing a time
step (∆t) which is sufficiently small compared to changes in the calculated values. For each
time step, equations (31) to (37) are solved for new values of Tc and ESR. This continues
until a steady state solution is reached.

The final lifetime acceleration constant, KV is meant to model voltage dependent ef-
fects on elcap lifetime. High applied voltages tend to produce H2 gas from the electrolyte
constituents that is then trapped inside the capacitor can. If the H2 is dissolved in the
electrolyte, ESR and Tc will increase [106]. Capacitors which are operated at voltages higher
than the rated voltage suffer from a decreased lifespan. Conversely, capacitors which are
operated at voltage lower than Vrated can have significantly longer lifespans due to reduced
stress on the dielectric layer [40] and beneficial healing of the dielectric [116].

A widespread empirical model for the voltage acceleration factor is (38), which is used
mainly for medum to large sized capacitors. For smaller capacitors, electrolyte loss is gov-
erned mainly by the temperature acceleration factor, KT , determined by (27).

Vapplied −N
KV = (38)
Vrated

where:
Vapplied is the applied voltage
Vrated is the manufacturer’s rated voltage
N is an exponential fit that varies by manufac-
turer, but usually 0≤N≤6

The lifetime of an elcap is affected by operational temperature, ripple current, and applied
voltage. The approximate acceleration effect on lifetime can be modeled. However, great care
should be used when using multiple acceleration factors. Since most of the factors are based
on empirical models with a number of simplifications and approximations, the uncertainty

49
in the calculated lifetime is compounded and may only hold for a certain capacitor from a
particular manufacturer utilized in a particular way. Skepticism is necessary when comparing
lifetimes of different types of capacitors from different manufacturers using different modeling
equations and definitions.

Metallized Thin Film

Metallized thin film capacitors have rated lifetimes under pulsed voltage operation rang-
ing from 60,000-150,000 hours (6.85-17.1 years) [117]. Typical lifetime testing for these types
of capacitors is at 125% rated voltage and 10o C above ambient temperature. Since the ma-
jority of metallized thin film capacitors are used in a pulsed power mode, manufacturers
usually report test charge/discharging cycles and it may be difficult to find manufacturer
data on constant voltage operation.

The manufacture process for biaxially oriented polypropylene thin film (the most common
thin film dielectric) involves extrusion of the polymer feedstock followed by stretching into
thin sheets. This manufacturing method introduces oxidation damage and residual stress
to the final polymer, causing the dielectric to begin aging relatively early [118]. To prevent
capacitors which age too quickly from being sent to consumers, most manufacturers will
apply voltages up to 125% of rated in a process known as pre-clearing [64].

When operated in a constant voltage mode, it has been shown in the past that metallized
thin film capacitors have significantly shorter lifetimes when held at peak voltage. During
constant voltage operation, clearing is constantly occurring so capacitance steadily drops.
Capacitors which have long lifetimes for pulsed applications would be expected to have much
shorter (<3,000 hours) lifetimes under constant voltage conditions [119, 91, 120, 76, 72].

Lifetime calculations must include some voltage multiplication factor, as lifetime de-
creases rapidly as DC voltage approaches Vrated . Most lifetime calculations for metallized
thin film capacitors define a voltage acceleration factor, N. This acceleration factor is usually
in the range of 10-20 [121, 122, 91, 123], depending on the quality, manufacture, and type
of capacitor. This very large voltage acceleration factor means that voltage rating of metal-
lized thin film capacitors is of primary importance, since any over-voltage can have drastic
deleterious effects on capacitor lifetime.

Much like electrolytic capacitors, metallized thin film capacitors have decreased lifetimes
as temperatures increase. The decrease due to temperature is greater than elcaps, with the
general rule being a halving of lifetime for every 8o C over rated temperature [39]. The voltage
and temperature acceleration factors can be combined into an overall lifetime equation as in
(39).

 −N
t Tr −Tc Vw
=2 8 · (39)
to Vrated

50
where:
to and t are base and accelerated lifetime
Tr and Tc are reference and core temperatures
Vw and Vrated are working and rated voltages
N is voltage acceleration factor (typically 10-20)

The working voltage is the sum of the DC and ripple voltages (40).

Vripple
Vw = VDC + (40)
2

As in electrolytic capacitors, ripple current increases the capacitor core temperature due
to resistive heating. This heat is partially transferred to the outer casing through conduction,
convection, and radiation with an efficiency that is determine by the thermal resistance of
the capacitor. The core temperature can be approximated by (41).

Tc = Ta + (Pd + Pt ) · Rth (41)

where: Ta and Tc are ambient and core temper-


atures
Pd and Pt are the dielectric and thermal power
losses, respectively
Rth is the capacitor thermal resistance

The dielectric loss is due to energy absorption by the polymer dielectric medium from
the rearrangement of dipoles at a frequency, f.

 
1 2
Pd = · Cn · Vripple · f · DF (42)
2

Thermal power loss is due to joule heating of the capacitor.

51
where:
Cn is the nominal capacitance
f is the ripple voltage frequency
DF is the dielectric loss factor (2×10−4 for
polypropylene) [66]

Co

!C = "2%+,-,."/,0.$%&µ1*%

!"#$%&'()*%

Figure 19: Typical results of a metallized thin film capacitance degradation over time. If the
degradation rate is not quadratic in time, then the degradation is roughly linear until soft failure.
After soft failure, the degradation rate rapidly increases. Examples of nonlinear degradation rates
can be seen in [72, 89, 122, 124, 125].

2
Pt = ESR · IRM S (43)

It has been mentioned previously that metallized thin film capacitors typically fail due
to a decrease in capacitance brought on by a number of small clearing events. These clearing
events are stochastic in nature, however, the number of events increases with time [65]. This
means that the degradation rate of thin film capacitors tends to increase in time (Figure 19).

The typical capacitance degradation for thin film capacitors is either quadratic with
respect to time or is roughly linear until soft failure. After soft failure, the degradation
rate increases exponentially. The increased degradation rate leads to an increased number
of clearing events. This, in turn, leads to higher pressure inside the capacitor can and may
eventually lead to flashover and catastrophic failure of the capacitor.

While one of the main advantages of thin film capacitors is soft failure and the ability to
use them completely safely even after this point, it is important to note that their electrical
attributes degrade markedly and may negatively affect the corresponding circuitry. Also,
if capacitors are not replaced relatively soon after soft failure, the buildup of gases due to
clearing events could lead to catastrophic failure. Although metallized thin film capacitors

52
can be used long after soft failure, it is important not to use them for too long without
replacement or catastrophic failure may result.

Ceramic

Ceramic capacitors can either suffer from degradation or catastrophic failure. Although
ceramic capacitors have a tendency to catastrophically fail during operation, there are no
models for catastrophic failure. The parameter of note in degradation failures of MLCCs is
capacitance.

MLCCs have a tendency for capacitance degradation during use according to (44) [126]:

C(t)
= 1 − k · log t (44)
Co

where:
C(t) is the capacitance at time, t
Co is initial capacitance
k is the dielectric aging rate, which is dielectric
dependent, but is usually between 120 and 130
[127]

This aging is due to the ferroelectric and volumetric response of the ceramic upon cooling
from the Curie Temperature [81]. The Curie Temperature is the point at which a ferromag-
netic material will become paramagnetic. At this point, the thermal energy overcomes the
attraction between magnetic moment of domains and magnetic spins become randomized
in a material. The capacitance of an MLCC can be restored by heating above the Curie
Temperature for 1-4 hours [81]. Commercial MLCC manufacturers frequently add doping
materials to the ceramic in order to lower the Curie Temperature and facilitate “de-aging”
of the capacitor by heating.

For most ceramic capacitors with Ag/Pd alloy electrodes, the lifetime can be upwards
of 8000 hours until degradation failure due to capacitance decrease (C(t)/Co =30%) under
85o C/85%relative humidity (RH) accelerated testing. For the less costly MLCCs with Ni
electrodes, failure can occur in less 500 hours with only limited reversibility [81].

Much like elcaps, MLCCs have a voltage dependent lifetime acceleration. This degrada-
tion is due to Poole-Frenkel emission [128] which leads to avalanche breakdown [99]. The
lifetime of the capacitor is inversely related to the applied voltage raised to the power, N
(45).

53
This value of N is highly dependent on ceramic type and morphology. For example,
delemanitions, cracks, voids, pores, etc. (known as extrinsic factors [96]) may result in local
electric fields much larger than the average [128]. For example, N =2.6 for V=50-150V across
a 2.5×10− 3cm BaTiO3 dielectric [129]. The value of N drops to 1.2 for the same dielectric
with thicker dielectric layers due to the reduction of the average internal electric field [129].
In general N=1-3 [128].

 N
t Vo
= (45)
to V

where:
t is voltage accelerated lifetime
to is baseline lifetime
Vo is baseline applied voltage
V is accelerated lifetime applied voltage
N is voltage acceleration factor

This voltage accelerated lifetime does not account for any temperature dependent effects.
An increase in leakage current under increased temperature is generally due to thermal
runaway [99]. This thermal runaway is general caused by intrinsic factors in the ceramic.
These intrinsic factors can be electronic disorders, dislocations, grain boundaries, etc. a
inside the ceramic [96].

Both avalanche and thermal breakdown are occurring during capacitor operation. The
combination of the two breakdown modes can be combined to to find an overall lifetime
acceleration according to (46) [130].

" #
N Ea
t1 V2 kB ( 1 − 1
)
= ·e T1 T2
(46)
t2 V1

This equation has been shown to fit well to data for high quality MLCCs under JIS
and MIL-STD testing conditions (C·R>20 MΩF with 200% Vrated at 85o C for 1000 hours)
[131, 132].

Reliability of Capacitors

The lifetime of a single average device can be approximated by calculations similar to


those in the previous section. For a large population of similarly constructed devices, the

54
where:
t1 and t2 are the baseline and accelerated mean
time to failure
V1 and V2 are the applied voltages
kB is Boltzmann’s constant
N is the voltage acceleration factor
Ea is activation energy for degradation,
∼1.87eV for purely thermal breakdown and
∼1.49eV for purely avalanche breakdown [99]

calculated lifetime would be equal. However, during operation, the actual failure of these
devices occurs at variable times due to slight differences in manufacturing, material quality,
and/or operational environment. This illustrates the importance of reliability. The reliability
of a device is defined as the probability that a device will perform its intended function during
a specified period of time under stated conditions [133]. This can be stated mathematically
as (47).
Z ∞
R(t) = Pr[T > t] = f (t) dt (47)
t

where:
R(t) is the reliability (number of devices not
failed) of a device
f(t) the distribution of failures

The distribution of failures, f(t), for a device in time is known as the probability density
function (PDF) and may be described by a number of different mathematical expressions
depending on the type of device. The number of failures, F(t), which have occurred over
some time period, T, is called the cumulative density function (CDF) and is related to the
reliability, R(t) by (48).

Z t
F (t) = 1 − R(t) = f (t) dt (48)
−∞

where:
F(t) is the cumulative distribution function
(CDF)

55
&*#*('5/$%
=">+0"?*5@A%@4%6'"(*0$>%
;<<% 8&=6:C%68+:%
100 + !

&*#*('5/$%6'"(*0$%7'+$%89:%
2

&'()*('+$,%-./$0'1$2%
3"4$5#$%

=">+0"?*5@A%@4%
3"4$5#$>%8B=6:C%48+:%
100 ! !
2
<%
!"#$% 36%
!3;% !3D%

Figure 20: Lifetime can be stated as a number for single, average device, LF . For a large popula-
tion of device, the individual lifetime of each device will vary slightly due to different manufacturing,
usage, and testing conditions. The distribution of lifetimes is known as a probability density func-
tion (PDF) and is the solid black line. The integral, or cumulative number of failures over time is
the cumulative distribution function (CDF) in red. The lifetime of a population of devices should
be described in terms of a confidence level (α).

Figure 20 shows this relationship graphically. The lifetime of a single device could be
measured or calculated to occur at some time (LF ). However, the lifetime of other similar
devices may be slightly different due to environmental, manufacturing, or test conditions.
For a large population of devices, the lifetimes will take on some distribution over time. This
distribution of lifetimes can be described by some function, f(t), known as the PDF.

If a number of devices are tested, the number of devices which fail continue to grow.
The number of expected failures in time can be described by the CDF, which is simply
the integral over time of the PDF. This number of expected failures is usually reported as
failures in time (FIT) and is defined as the number of failures per 1 billion (1×109 ) hours
(49), though it may be reported per 1 million hours or other time period as data dictates.

# failures
FIT =
1 × 109 hours
f (t)
= 1 × 109 ·
R(t)
f (t)
= 1 × 109 · (49)
1 − F (t)

Manufacturers are well aware of the variability of device lifetime. A much more indicative
way of communicating device reliability is to give a lifetime with a certain confidence interval

56
(α) rather than a single number, such as the average lifetime. When taking confidence level
into account, the lifetime of a device type is usually stated as (50).

LF − ∆L1 ≤ LF ≤ LF + ∆L2 (50)

The low confidence threshold, ∆L1 , is defined as (51). ∆L2 is defined in a similar way.
The values of ∆L1 and ∆L2 are determined both by the distribution of failures and the
confidence level. For symmetric distributions, ∆L1 = ∆L2 = ∆L and the lifetime can be
described as LF ± ∆L.

1−α
F(LF − ∆L1 ) =
2 
1−α
−1
LF − ∆L1 = F
2
 
−1 1−α
∆L1 = LF − F (51)
2

where:
F(LF − ∆L1 ) the value of the CDF at the lower
confidence limit
α is the confidence level
F −1 1−α

2
is the inverse CDF at the lower con-
fidence limit

If a manufacturer quotes a lifetime with a confidence level of 90%, this means that around
10% of capacitors will fail after LF −∆L1 and before LF +∆L2 . This statement says nothing
about which devices will fail prematurely or have an exceptionally long life, but that overall
90% will be between the low and high confidence levels. Confidence intervals are similar to
standard deviations. For example, a confidence interval of 95.45% is equal to two standard
deviations for a normal distribution.

Another way that many manufacturers may choose to present reliability information is
to say that 1+α
2
% of devices will have a lifetime greater than the lower confidence limit
(LF > LF − ∆L1 ). This is a completely equivalent to (50) for symmetric distributions.
However, for asymmetric distributions there may still be a very large distribution of lifetimes
since ∆L2 >> ∆L1 .

Some manufacturers may give a single number for device lifetime with no information
regarding device failure distribution or confidence level. This is bad practice and should be
avoided because, at best, it discards necessary information and, at worst, it may be used

57
Gaussian  CDF   Lognormal  PDF  
Mean   Mean  
Median   Median  
Mode   Mode  
Guassian  PDF   Lognormal  CDF  

0   2   4   6   8   10   12   0   0.2   0.4   0.6   0.8   1   1.2   1.4   1.6  

0   2   4   6   8   10   12   14   16   18   20   0   0.5   1   1.5   2   2.5   3   3.5   4   4.5   5  

Figure 21: (Left) Two examples of devices which have Normal or Gaussian PDFs. Both distribu-
tions have the same mean, median, and mode values, but the confidence level of the top distribution
would be much tighter than the bottom distribution. (Right) Two examples of a lognormal PDFs.
Both distributions have the same median value, but, for the bottom figure, over 75% of devices
would have failed before the given mean failure time.

to obfuscate poor device performance. The right side of Figure 21 shows how the mean of
lifetime distributions may be good, but almost 75% of devices will have failed before the
mean. The opposition may possibly be true, where the majority of lifetimes are longer than
the mean lifetime. This may be undesirable because it makes it very difficult to accurately
control maintenance schedules. Many operators may not care how short or long the mean
lifetime may be, just that the confidence levels are tight enough to accurately optimize
maintenance schedules. In other words, they do not care when the fail, just that all the
devices fail at the same time.

There exist many different formulations for PDF that can describe different manufac-
turing processes or device applications. The most popular formulations for PDF include a
Gaussian, log-normal, exponential, and Weibull. A Gaussian distribution is a simple bell
curve which is symmetric around some average value. This distribution is seldom used for
capacitor reliability because most of the time, failure distribution is non-symmetrical [117].
A log-normal distribution has a long, non-symmetric tail. This distribution was used ex-
tensively in the past, but in the past 20 years, has been largely replaced with the Weibull
distribution. An exponential distribution is sometimes used because of its ease of calculation,
however, it is simply a speciallized form of the Weibull distribution.

For most capacitor applications, the PDF can be described by a Weibull function (52).
The Weibull distribution for a time (t) is dependent on two variables, the shape function

58
(k), and the scale function (λ).

 k−1
k t tk
PDF = f (t, k, λ) = e− λ (52)
λ λ

where:
k is the shape function
λ is the scale function

By integrating the Weibull PDF, the CDF can be found to be (53).

Z t
CDF = F (t, k, λ) = f (t, k, λ) dt
−∞
t k
= 1 − e−( λ ) (53)

The FIT is found for a Weibull distribution to be (54).

f (t, k, λ)
FIT(t, k, λ) = 1 × 109 ·
1 − F (t, k, λ)
 k−1
9 k t
= 1 × 10 · (54)
λ λ

The failure rate over time for a given device can generally be described by a figure known
as a bathtub curve (Figure 22). This type of curve is fairly well-controlled by three regions.
Region I occurs at small times and shows a large failure rate which decreases as time increases.
Failure in this region is known as “infant mortality” failure and is controlled by quality
assurance, consistency in the manufacturing process, and adequate design practices [134].
After infant mortality, the overall failure rate drops to a (hopefully) small, constant value
in region II. In this region, which generally lasts for the majority of device lifetime, device
failure occurrence and cause are random. These failures do not adhere to any particular
pattern and are not predictable. After the constant failure region, the failure rate begins to
dramatically and rapidly increase in region III. This region is controlled by wear-out failure
which indicates that temperature, electrical stress, or other usage conditions have degraded
the device past its expected lifetime.

The bathtub curve in total is usually the combination of three (or more) Weibull dis-
tributions whose cumulative effect describe the bathtub curve for all regions. Region I is

59
I
II
III

Failure  Rate  (%)  


Observed  Failure  
Rate  

“Infant  Mortality”  
Failure   Wear  Out  Failure  
Constant  Random  
Failure  

Time  

Figure 22: Typical distributions of failures over time take the form of a bathtub curve (black).
This curve is split into three regions. Region I is dominated by a decreasing failure rate known
as “infant mortality”. Region II is dominated by a relatively constant failure rate which occur
randomly. Region III is dominated by a steeply increasing failure rate as devices reach their end-
of-life [62].

generally described by a Weibull distribution with k<1. The exact value of k and λ depend
on the length and slope of this infant mortality failure region. Region II is described by
an exponential distribution, which is a simplified Weibull distribution with k=1. Region III
is usually described by a Weibull distribution with k>1. The exact values of k and λ are
determined by the length and severity of the infrant mortality, random failure, and wear-out
regions [135].

Typically, for a number of devices, such as those used in an inverter, it is more useful
to discuss the average time until a component fails, regardless of the cause. This average is
known as mean time to failure (MTTF) and is a much more common parameter for discussing
how many components will affect an overall system. MTTF is basically a measurement of
how long a component or system of components should be expected to work without failing
[4] and is defined as the expectation value of the PDF:
Z ∞
MTTF = tf (t) dt (55)
0

MTTF is closely related to a value known as the mean time between failure (MTBF).
These two values are sometimes substituted for each other, although there is a subtle differ-
ence. MTTF is meant to be used in systems or components which cannot be repaired (such
as isolated or residential systems), so the repair time approaches infinity. MTBF is meant
to be used for systems or components which are continually replaced (such as utility scale
systems which undergo maintenance), so that the repair time approaches zero.

60
In reality, no repair time is zero. Once a device fails, there is a certain period between the
time that a device fails and the time that a device can be repaired or replaced. Between these
two times, the device cannot perform its function. This “failed” time period is described by
the mean time to repair (MTTR). The combination of MTTF and MTTR determine that
availability (Φ), or percentage of operational time, of a device according to (56):

MTTF
Φ= (56)
MTTF + MTTR

where:
MTTF is mean time to failure
MTTR is mean time to repair

Both MTTF and MTTR affect the overall availability of a system according to (56).
It is worth noting that availability can either be maximized (Φ →1) as MTTF→ ∞ or
MTTR→ 0. This has significant implications in the design of a system of components.

Both MTTF and MTTR must be extracted from experimental data if it is available in
significant volume [4]. It should be noted that MTTF should not be confused with the useful
life of a particular component, only the bulk failure behavior of a number of devices [136].
For example, published data for the MTTF of a solar cell module can be hundreds of years
for a residential unit. If a manufacturer claims a module MTTF of 500 years in the field, it
does not mean that a single module should be expected to fail only after 500 years, but that
one in every 500 modules will fail in a given year. Even this seemingly large MTTF may
not be ideal for an end-user. If, for example, a utility scale PV plant has 5,000 modules in
a field, a MTTF of 500 years means that every year they would expect 10 modules to fail.
This may represent unacceptable maintenance and replacement costs for the user.

A complicated system, such as an inverter, is composed of a large number of devices


which each have a different MTTF, MTTR, bathtub curve, PDF, and CDF. Each device
may be in different regions of their bathtub curve. The overall system availability is the
product of the availabilities of N components or subsystems:

N
Y
Φsys = Φi (57)
i=1

It is important to note that the overall availability of the system cannot be larger than
the shortest component availability. Simply, the availability of the system is determined by
its weakest link [4]. However, the availability may be much lower than the lowest component
availability since the system may fail due to the failure of one or multiple components.

61
where Φsys is overall system availability
Φi is the availability of the ith component or
subsystem

)% )% )%

&% &% &% &%


&'%

&(%

!"#$%

Figure 23: Over time, a system can either be operational (O) or non-operational (F). MTTF is
the average of t O times and MTTR is the average of F times. The overall availability (ratio of O
to F) may be minimized by increasing the average of O or decreasing the average of F.

Figure 23 shows the relationship between Φ, MTTF, and MTTR graphically. The MTTF
is the average of the on state times (O). The average of the off times of the system (F) is
MTTR.

As was mentioned previously, the availability of a system can be maximized either through
the decrease of MTTR or the increase of MTTF. The choice of optimization of either or both
of these terms should be considered when a system is designed. MTTF is usually determined
by the quality of components, so it is dependent on the research and development of original
equipment manufacturers. MTTF of a system can be increased by using high quality devices,
such as those with military specifications. The devices demonstrate no infant mortality
because they have been through a “burn-in” process where devices have been stressed to
remove any infant mortality failures before shipment to the consumer. MTTF can also
be increased (for a system) through adding redundant components (however, the increased
complexity increases cost and decreases component MTTF) [136]. Although MTTF can be
increased through the utilization of high quality devices, it is usually more cost effective to
decrease MTTR.

MTTR is usually in the hands of system designers and can be decreased by good en-
gineering (easy swap-in of components) and error detection (so the interval of decreased
output efficiency due to component failure is shortened). MTTR can also be decreased by
using devices which do not necessarily have the greatest lifetime, but have very tight confi-
dence intervals. By choosing devices where the time period of failure is fairly well known,
maintenance can be optimized and repair times and costs can be reduced.

Other than obtaining reliability data from manufacturers, one of the main problems fac-
ing system engineers is the lack of consistent terminology and experimental practices from
manufacturer to manufacturer. This makes comparing components or systems across man-

62
ufacturers extremely difficult. This is especially true in the case of capacitor manufacturers.

Some capacitor manufacturers may describe lifetimes or reliability according to an en-


durance test. These tests are usually carried out in accordance with IEC 60684 standards
(Vrated and maximum operation temperature until C, ESR, or Ileakage reached). However,
some companies may instead described a useful life (typically endurance test with an added
ripple current). Other manufacturers may describe terms like load life, life expectancy, oper-
ational life, and service life. All of these terms may mean the same, or different things with
completely different testing procedures.

In addition to the varied terminology which is used, there is no consensus for standards in
reliability testing. different manufacturers define reliability according to different thresholds
and experimental conditions for different parameters. The Electronic Industries Alliance
(IEC) has recently passed a standard for capacitor testing (IEC IS-749), though the stan-
dard is not currently used by all capacitor manufacturers. Some manufacturers may test
their capacitors in compliance with standard JESD-22 THB and HAST (high accelerated
stress test) while others may be more concerned with complying with standard CEI 1071
(voltage spikes). Some manufacturers may test their capacitors according to MIL217F, EIA
198, HALT (highly accelerated lifetime test, 40-60o C/minute cycling with vibration) or IEC
60749-33 depending on the application concerns for their capacitor. Many manufacturers
may develop their own internal standard testing method which is not recognized by any
governing body [81].

It can be quite challenging to compare lifetime and reliability data from one manufacturer
to another due to the difference in standards, testing conditions, and definitions for failure.

Summary and Conclusions

This work has investigated the construction, utility, lifetime, and reliability of capacitors
for use in PV inverter systems. The cause of ripple voltage in inverters due to DC-to-AC
conversion was discussed along with associated deleterious effects on solar module MPPT and
how bus capacitor use diminishes voltage ripple. Two types of capacitors were introduced
(elcap and metallized thin film) along with their operating principles, construction, and
degradation/failure modes. Several lifetime models were then introduced for each of these
capacitor types. Finally, this work introduced the importance of reliability in inverter usage.

Lifetimes of solar cell modules are typically in the range of 30 years. It is completely
possible to manufacture an inverter system which can function for the entire lifetime of a
solar cell module through the use of MIL quality parts and component redundancy. However,
such an inverter would be so costly that the price per watt of energy produced could never
be competitive with other energy sources. Instead of having a single-minded goal towards
increasing the lifetime of inverters, an inverter designer should instead consider maximiz-
ing the overall availability of the system. This can be done via a detailed lifecycle analysis
considering the cost-effective use of high quality parts to increase component MTTF, re-

63
dundant parts to increase system MTTF, better engineering practices to decrease MTTR,
and improved error identification to prevent inverter operation at decreased efficiency due
to component failure.

It is an unfortunate fact that, for the time being, capacitors failure will negatively impact
inverter operation. This impact can be minimized through proper lifetime and reliability
testing that can then be used to optimize maintenance costs. Unfortunately, if published,
every capacitor manufacturer uses different standards (if they use them at all). In the future,
the standardization of capacitor lifetime and reliability testing is sorely needed and would
be a valuable resource for inverter designers.

Through a combination of improved component manufacture, standardization of lifetime


and reliability testing, and better engineering practices regarding overall inverter lifecycle
costs, it will be possible to decrease the price per watt of inverter energy output to meet the
DOE goals by 2017.

64
References

[1] “$1/watt photovoltaic systems,” US Department of Energy, White Paper, Aug. 2010.

[2] Y. Xue, K. C. Divya, G. Griepentrog, M. Liviu, S. Suresh, and M. Manjrekar, “Towards


next generation photovoltaic inverters,” in 2011 IEEE Energy Conversion Congress
and Exposition (ECCE), 2011, pp. 2467–2474.

[3] L. Bony, S. Doig, C. Hart, and E. Maurer, “Achieving low-cost solar pv: Industry
workshop recommendations for near-term balance of system cost reductions,” Rocky
Mountain Institute (RMI), Tech. Rep., 2010.

[4] A. Ristow, M. Begovic, A. Pregelj, and A. Rohatgi, “Development of a methodol-


ogy for improving photovoltaic inverter reliability,” IEEE Transactions on Industrial
Electronics, vol. 55, no. 7, pp. 2581–2592, Jul. 2008.

[5] B. Jablonska, H. Kaan, M. Leeuwen, and G. Boer, “Pv-prive project at ecn: Five years
of experience with small-scale ac module pv systems,” in 20th European Photovoltaic
Solar Energy Coference and Exhibition, 2005.

[6] A. Maish, “Defining requirements for improved photovoltaic system reliability,”


Progress in Photovoltaics, pp. 165–173, 1999.

[7] A. B. Maish, C. Atcitty, S. Hester, D. Greenberg, D. Osborn, D. Collier, and M. Brine,


“Photovoltaic system reliability,” in Conference Record of the Twenty Sixth IEEE Pho-
tovoltaic Specialists Conference, 1997, 1997, pp. 1049–1054.

[8] A. Golnas, “Pv system reliability: An operator’s perspective,” in 38th Photovoltaic


Specialists Conference, Austin, TX, Jun. 2012, pp. 1–32.

[9] IEEE, “1547-ieee standard for interconnecting distributed resources with electric power
systems,” IEEE, Tech. Rep., 2003.

[10] “Iec61727-photovoltaic (pv) systems: Characteristics of the utility interface,” Interna-


tional Electrotechnical Commission, Tech. Rep., 1996.

[11] W. Bower and D. Ton, “Summary report on the doe high-tech inverter workshop,”
DOE Office of Energy Efficiency and Renewable Energy, Tech. Rep., 2005.

[12] C. Rodriguez and G. A. J. Amaratunga, “Long-lifetime power inverter for photovoltaic


ac modules,” IEEE Transactions on Industrial Electronics, vol. 55, no. 7, pp. 2593–
2601, Jul. 2008.

65
[13] T. Von Kampen and E. Sawyer, “Reliability considerations of inverter/dc link ca-
pacitor using pp film and 105c engine coolant,” in Proceedings of the International
Microelectronics and Pakaging Society (IMAPS), 2008.

[14] S. J. Parler, “Thermal modeling of aluminum electrolytic capacitors,” in Conference


Record of the 1999 IEEE Industry Applications Conference, 1999, pp. 2418–2429.

[15] R. Morgolis, “A review of pv inverter technology cost and performance projections,”


National Renewable Energy Laboratory, Tech. Rep., 2006.

[16] S. B. Kjaer, J. K. Pedersen, and F. Blaabjerg, “A review of single-phase grid-connected


inverters for photovoltaic modules,” IEEE Transactions on Industry Applications,
vol. 41, no. 5, pp. 1292–1306, Sep. 2005.

[17] A. M. Trzynadlowski, Introduction to Modern Power Electronics. Wiley, Mar. 2010.

[18] M. Salcone and J. Bond, “Selecting film bus link capacitors for high performance in-
verter applications,” in IEEE International Electric Machines and Drives Conference,
2009 (IEMDC ’09), 2009, pp. 1692–1699.

[19] C. Siedle and V. D. Ingenieure, Comparative Investigations of Charge Equalizers to


Improve the Long-Term Performance of Multi-cell Battery Banks. VDI Verlag, 1998,
vol. 245.

[20] A. Luque and S. Hegedus, “Handbook of photovoltaic science and engineering,” Wiley,
pp. 1–1138, Mar. 2011.

[21] O. Wasynezuk, “Dynamic behavior of a class of photovoltaic power systems,” IEEE


Transactions on Power Apparatus and Systems, vol. 102, no. 9, pp. 3031–3037, 1983.

[22] K. Hussein, I. Muta, T. Hoshino, and M. Osakada, “Maximum photovoltaic power


tracking: An algorithm for rapidly changing atmospheric conditions,” IEEE Proceed-
ings on Generation, Transmission and Distribution, pp. 59–64, 1995.

[23] M. de Brito, L. Sampaio, L. Junior, and C. Canesin, “Evaluation of mppt techniques


for photovoltaic applications,” 2011 IEEE International Symposium on Industrial Elec-
tronics (ISIE 2011), pp. 1039–1044, 2011.

[24] S. B. Kjaer, “Design optimization of a single phase inverter for photovoltaic applica-
tions,” Ph.D. dissertation, University of Aalborg, Aalborg, Denmark, May 2005.

[25] R. C. Dorf and J. A. Svoboda, Introduction to Electric Circuits. John Wiley & Sons,
2001.

[26] M. A. Laughton and D. F. Warne, Electrical Engineer’s Reference Book. Newnes,


2002.

[27] D. R. Lide, CRC handbook of Chemistry and Physics. CRC Press, Jun. 2004.

66
[28] S. Giere, M. Kurrat, and U. Schumann, “Hv dielectric strength of shielding electrodes in
vacuum circuit-breakers,” 20th International Symposium on Discharges and Electrical
Insulation in Vacuum, 2002., pp. 119–122, 2002.

[29] L. G. Hector and H. L. Schultz, “The dielectric constant of air at radio frequencies,”
Physics-A Journal of General and Applied Physics, vol. 7, no. 1, pp. 133–136, 1936.

[30] D. Tripathi, Practical Guide to Polypropylene. Smithers Rapra Press, Apr. 2002.

[31] Y. Inuishi and D. A. Powers, “Electric breakdown and conduction through mylar films,”
Journal of Applied Physics, vol. 28, no. 9, pp. 1017–1022, 1957.

[32] A. J. Moulson and J. M. Herbert, Electroceramics: Materials, Properties, Applications.


John Wiley & Sons Inc, Jul. 2003.

[33] M. H. Yeh, Y. C. Liu, K. S. Liu, I. N. Lin, J. Y. M. Lee, and H. F. Cheng, “Electrical


characteristics of barium titanate films prepared by laser ablation,” Journal of Applied
Physics, vol. 74, no. 3, pp. 2143–2145, 1993.

[34] T. Lebey, V. Bley, S. Guillemet, M. Boulos, and B. Durand, “Origin of the colossal
permittivity and possible applications of cct ceramics,” in 55th Electronic Components
and Technology Conference, 2005 (ECTC), 2005, pp. 1248–1253.

[35] M. Arbatti, X. Shan, and Z. Y. Cheng, “Ceramic-polymer composites with high di-
electric constant,” Advanced Materials, vol. 19, no. 10, pp. 1369–1372, May 2007.

[36] F. T. Ulaby, Fundamentals of Applied Electromagnetics. Prentice Hall, 2007.

[37] J. O. Bird, Electrical Circuit Theory and Technology. Routledge, Mar. 2003.

[38] P. Scherz, Practical Electronics for Inventors. McGraw-Hill/Tab Electronics, Nov.


2006.

[39] W. Sarjeant, J. Zirnheld, and F. MacDougall, “Capacitors,” IEEE Transactions on


Plasma Science, vol. 26, no. 5, pp. 1368–1392, 1998.

[40] A. Albertsen, “Electrolytic capacitor lifetime estimation,” Jianghai Europe GmbH,


Tech. Rep., 2010.

[41] A. Morley and D. Campbell, “Electrolytic capacitors -their fabrication and the inter-
pretation of their operational behavior,” Radio Electronic Engineering, vol. 43, pp.
421–429, Jul. 1973.

[42] H. Ma and L. Wang, “Fault diagnosis and failure prediction of aluminum electrolytic
capacitors in power electronic converters,” in 31st Annual Conference of IEEE Indus-
trial Electronics Society, 2005 (IECON 2005), 2005, pp. 842–847.

[43] O. G. Palana, Engineering Chemistry. McGraw-Hill Education, 2009.

67
[44] K. Arai, T. Suzuki, and T. Atsumi, “Effect of trace-elements on etching of aluminum
electrolytic capacitor foil,” Journal of the Electrochemical Society, vol. 132, no. 7, pp.
1667–1671, 1985.

[45] M. Saif, S. Zhang, A. Haque, and K. Hsia, “Effect of native al2o3 on the elastic response
of nanoscale al films,” Acta Materialia, vol. 50, no. 11, pp. 2779–2786, 2002.

[46] P. M. Deeley, “Electrolytic capacitors,” The Cornell-Dubilier Electric Corporation,


Tech. Rep., 1938.

[47] S. Parler, “Application guide, aluminum electrolytic capacitors,” Cornell-Dublier,


Tech. Rep.

[48] I. Nagata, “Aluminum dry electrolytic capacitors,” Japan Capacitor Industry, Tech.
Rep., 1983.

[49] W. Lin, G. Tu, and C. Lin, “The effect of indium impurity on the dc-etching behaviour
of aluminum foil for electrolytic capacitor usage,” Corrosion science, vol. 39, no. 9, pp.
1531–1543, 1997.

[50] P. G. Sheasby, R. Pinner, and S. Wernick, The Surface Treatment and Finishing of
Aluminium and Its Alloys. ASM International, Jan. 2001.

[51] A. Zagiel, P. Natishan, and E. Gileadi, “Plating on anodized aluminum-the effect of


the metal, the anion and the aluminum alloy,” Electrochimica Acta, vol. 35, no. 6, pp.
1019–1030, Jun. 1990.

[52] J. O’Sullivan, “The morphology and mechanism of formation of porous anodic films
on aluminium,” in Proceedings of the Royal society of London. Series A, Mathematical
and Physical Sciences, vol. 317, no. 1531, 1970, pp. 511–543.

[53] G. E. Thompson, “Porous anodic alumina: Fabrication, characterization and applica-


tions,” Thin Solid Films, vol. 297, no. 1-2, pp. 192–201, Apr. 1997.

[54] N. Pilling and R. Bedworth, “The oxidation of metals at high temperatures,” Journal
of the Institute of Metals, vol. 29, pp. 529–582, 1923.

[55] L. J. Korb and Committee, ASM Handbook. Volume 13. Corrosion. ASM Interna-
tional, 1998, vol. 13.

[56] R. S. Timsit, W. G. Waddington, C. J. Humphreys, and J. L. Hutchison, “Structure


of the al/al2o3 interface,” Applied Physics Letters, vol. 46, no. 9, p. 830, 1985.

[57] S. G. Parler, “Improved spice models of aluminum electrolytic capacitors for inverter
applications,” IEEE Transactions on Industry Applications, vol. 39, no. 4, pp. 929–935,
Jul. 2003.

[58] G. Terzulli and B. W. Peace, “Film technology to replace electrolytic technology,”


AVX Corporation, Tech. Rep., Jul. 2005.

68
[59] N. Helmold and C. Hillman, “Identification of missing or insufficient electrolyte con-
stituents in failed aluminum electrolytic capacitors,” in 24th Annual Capacitor and
Resistor Technology Symposium (CARTS), 2004, pp. 122–127.
[60] J. Stevens and T. Marshall, “The effects of electrolyte composition on the deforma-
tion characteristics of wet aluminum icd capacitors,” in The Capacitor and Resistor
Technology Symposium (CARTS), 2006.
[61] J. Ho, T. R. Jow, and S. Boggs, “Historical introduction to capacitor technology,”
IEEE Electrical Insulation Magazine, vol. 26, no. 1, pp. 20–25, 2010.
[62] M. El-Husseini, P. Venet, G. Rojat, and M. Fathallah, “Effect of the geometry on
the aging of metalized polypropylene film capacitors,” in IEEE 32nd Annual Power
Electronics Specialists Conference, 2001 (PESC 2001), 2001, pp. 2061–2066.
[63] L. Hua, L. Fuchang, Z. Heqing, D. Ling, H. Yongxia, and K. Zhonghua, “Study on
metallized film capacitor and its voltage maintaining performance,” IEEE Transactions
on Magnetics, vol. 45, no. 1, pp. 327–330, 2009.
[64] C. W. Reed and S. W. Cichanowskil, “The fundamentals of aging in hv polymer-film
capacitors,” IEEE Transactions on Dielectrics and Electrical Insulation, vol. 1, no. 5,
pp. 904–922, 1994.
[65] J. Zhao and F. Liu, “Reliability assessment of the metallized film capacitors from
degradation data,” Microelectronics and Reliability, vol. 47, no. 2-3, pp. 434–436, Feb.
2007.
[66] G. Buiatti, S. Cruz, and A. Cardoso, “Lifetime of film capacitors in single-phase re-
generative induction motor drives,” IEEE International Symposium on Diagnostics
for Electric Machines, Power Electronics and Drives, 2007 (SDEMPED 2007), pp.
356–362, 2007.
[67] P. Harrop and D. Campbell, “Selection of thin film capacitor dielectrics,” Thin Solid
Films, vol. 2, no. 4, p. 273, 1968.
[68] N. Henze and J. Liu, “Reliability considerations of low-power grid-tied inverter for
photovoltaic application,” in 24th European Photovoltaic Solar Energy Conference and
Exhibition, Hamburg, Germany, Sep. 2009, pp. 21–25.
[69] M. Rabuffi and G. Picci, “Status quo and future prospects for metallized polypropylene
energy storage capacitors,” IEEE Transactions on Plasma Science, vol. 30, no. 5, pp.
1939–1942, Oct. 2002.
[70] J. Cozens, “Development of plastic dielectric capacitors,” IRE Transactions on Com-
ponent Parts, vol. 6, no. 2, pp. 114–118, 1959.
[71] M. Carlen, P. Bruesch, and R. Gallay, “Electrical endurance characterization of
polypropylene winding capacitors for traction applications: a new experimental
method,” in Seventh International Conference on Dielectric Materials, Measurements
and Applications, 1996, pp. 350–353.

69
[72] J. B. Ennis, F. W. MacDougall, X. H. Yang, R. A. Cooper, K. Seal, C. Naruo,
B. Spinks, P. Kroessler, and J. Bates, “Recent advances in high voltage, high energy ca-
pacitor technology,” in IEEE International Pulsed Power Plasma Science Conference,
2007 (PPPS 2007), 2007, pp. 282–285.

[73] D. Xin, L. Fuchang, L. Jin, Y. Zonggan, and W. Nanyan, “Influence factors for the
self-healing of metallized polypropylene capacitors,” in Annual Report Conference on
Electrical Insulation and Dielectric Phenomena, 2000, pp. 461–465.

[74] A. Schneuwly, P. Groning, and L. Schlapbach, “Uncoupling behaviour of current gates


in self-healing capacitors,” Materials Science and Engineering B-Solid State Materials
for Advanced Technology, vol. 55, no. 3, pp. 210–220, 1998.

[75] A. Gully, “Failure mechanisms in film-based power capacitors,” in Seventh Interna-


tional Conference Dielectric Materials, Measurements and Applications, no. 430, 1996,
pp. 358–363.

[76] D. Shaw, S. Cichanowski, and A. Yializis, “A Changing Capacitor Technology - Fail-


ure Mechanisms and Design Innovations,” Ieee Transactions on Electrical Insulation,
vol. 16, no. 5, pp. 399–413, 1981.

[77] F. Lin, D. Xin, L. Jin, Y. Zonggan, and W. Nanyan, “On the failure mechanism of
metallized polypropylene pulse capacitors,” in Conference on Electrical Insulation and
Dielectric Phenomena, 2000, 2000, pp. 592–595.

[78] A. Schneuwly, P. Groning, L. Schlapbach, C. Irrgang, and J. Vogt, “Breakdown Be-


havior of Oil-impregnated Polypropylene as Dielectric in Film Capacitors,” Dielectrics
and Electrical Insulation, IEEE Transactions on, vol. 5, no. 6, pp. 862–868, 1998.

[79] C. Maier and T. Calafut, Polypropylene: Definitive User’s Guide. William Andrew,
Dec. 1998.

[80] D. Donahoe and C. Hillman, “Failures in base metal electrode (bme) capacitors,” in
23rd Capacitor and Resistor Technology Symposium, 2003, pp. 129–138.

[81] D. Donahoe, M. Pecht, I. Lloyd, and S. Ganesan, “Moisture induced degradation of


multilayer ceramic capacitors,” Microelectronics and Reliability, vol. 46, no. 2-4, pp.
400–408, Feb. 2006.

[82] E. Zysk, Precious Metals and Their Uses: Nonferrous Alloys and Pure Metals, 9th ed.,
Zysk, Ed. Metals Park, Ohio: American Society for Metals, 1979, vol. 2.

[83] T. Nomura, Y. Nakano, and A. Satoh, “Multilayer ceramic chip capacitor,” US Patent
5,319,517, 1994.

[84] X. Zhang, T. Hashimoto, and D. Joy, “Electron holographic study of ferroelectric


domain-walls,” Applied Physics Letters, vol. 60, no. 6, pp. 784–786, 1992.

70
[85] J. M. Herbert, “Production of ceramic material,” US Patent 3,041,189, Tech. Rep.,
1962.

[86] C. Kulkarni, G. Biswas, X. Koutsoukos, J. Celaya, and K. Goebel, “Experimental


studies of ageing in electrolytic capacitors,” in Annual Conference of the Prognostics
and Health Management Society, 2010, pp. 1–7.

[87] D. Guo, “Wear analysis and degradation mechanism of film metallized capacitor,”
Power Capacitor (China), vol. 2, pp. 12–15, 1995.

[88] M. L. Gasperi, “Life prediction model for aluminum electrolytic capacitors,” in Con-
ference Record of the 1996 IEEE Industry Applications Conference (IAS 1996), 1996,
pp. 1347–1351.

[89] H. Matsui, T. Fujiwara, and K. Fujiwara, “Metalized film capacitors with high energy
density for rail vehicles,” in Conference Record of the 1997 IEEE Industry Applications
Conference (IAS ’97), 1997, pp. 1079–1091.

[90] B. Walgenwitz, J. Tortai, N. Bonifaci, and A. Denat, “Self-healing of metallized poly-


mer films of different nature,” in Proceedings of the 2004 IEEE International Confer-
ence on Solid Dielectrics (ICSD 2004), vol. 1, 2004, pp. 29–32.

[91] W. Sarjeant, F. MacDougall, D. Larson, and I. Kohlberg, “Energy storage capaci-


tors: Aging, and diagnostic approaches for life validation,” in IEEE Transactions on
Magnetics. Aerovox Inc,New Bedford,Ma, 1997, pp. 501–506.

[92] H. Lee, K. Lee, J. Schunke, and L. Burton, “Leakage currents in multilayer ceramic ca-
pacitors,” IEEE Transactions on Components Hybrids and Manufacturing Technology,
vol. 7, no. 4, pp. 443–453, 1984.

[93] A. L. Young, G. E. Hilmas, S. C. Zhang, and R. W. Schwartz, “Mechanical vs. electrical


failure mechanisms in high voltage, high energy density multilayer ceramic capacitors,”
Journal of Materials Science, vol. 42, no. 14, pp. 5613–5619, Apr. 2007.

[94] S. Lucato, D. Lupascu, M. Kamlah, J. Rodel, and C. Lynch, “Constraint-induced crack


initiation at electrode edges in piezoelectric ceramics,” Acta Materialia, vol. 49, no. 14,
pp. 2751–2759, 2001.

[95] W. Bahn and R. Newnham, “Macrodefects in batio3 capacitors,” Materials Research


Bulletin, vol. 21, no. 9, pp. 1073–1082, 1986.

[96] J. Yamamatsu, N. Kawano, T. Arashi, and A. Sato, “Reliability of multilayer ceramic


capacitors with nickel electrodes,” Journal of Power Sources, vol. 60, pp. 199–203,
1996.

[97] E. Loh, “A model of dc leakage in ceramic capacitors,” Journal of Applied Physics,


vol. 53, no. 9, pp. 6229–6235, 1982.

71
[98] L. L. Hench and J. K. West, Principles of Electronic Ceramics. Wiley-Interscience,
1990.

[99] B. Rawal and N. Chan, “Conduction and failure mechanisms in barium titanate based
ceramics under highly accelerated conditions,” in Proceedings of the 34th Electronic
Components Conference, 1984, pp. 184–188.

[100] S. Freiman and R. Pohanka, “Review of mechanically related failures of ceramic capac-
itors and capacitor materials,” Journal of the American Ceramic Society, pp. 2258–
2263, 1989.

[101] E. K. Beauchamp, “Effect of microstructure on pulse electrical strength of mgo,” Jour-


nal of the American Ceramic Society, vol. 54, no. 10, pp. 484–487, 1971.

[102] R. Gerson and T. Marshall, “Dielectric breakdown of porous ceramics,” Journal of


Applied Physics, vol. 30, no. 11, pp. 1650–1653, 1959.

[103] S. G. Parler Jr, “Reliability of cde aluminum electrolytic capacitors,” Cornell-Dublier,


Tech. Rep., Nov. 2004.

[104] W. Q. Meeker and G. J. Hahn, How to Plan an Accelerated Life Test: Some Practical
Guidelines. MIlwaukee: American Society for Quality Control, Statistics Division,
1985, vol. 10.

[105] P. Venet, A. Lahyani, G. Grellet, and A. Ah-Jaco, “Influence of aging on electrolytic


capacitors function in static converters: Fault prediction method,” European Physical
Journal-Applied Physics, vol. 5, no. 1, pp. 71–83, 1999.

[106] S. Parler Jr, “Selecting and applying aluminum electrolytic capacitors for inverter
applications,” Cornell-Dublier, Tech. Rep., 2010.

[107] G. Rhoads and A. Smith, “Expected life of capacitors with non-solid electrolyte,” in
IEEE 34th Electronic Component Conference, 1984, p. 156.

[108] C. Kulkarni and G. Biswas, “A prognosis case study for electrolytic capacitor degra-
dation in dc-dc converters,” Annual Conference of the Prognostics and Health Man-
agement Society, 2010, 2010.

[109] G. Biswas, X. Koutsoukos, and K. Goebel, “Physics of failure models for capacitor
degradation in dc-dc converters,” in The Maintenance and Reliability Conference,
MARCON, 2010.

[110] R. C. Rosenberg and D. Karnopp, Introduction to Physical Systems Dynamics.


McGraw-Hill College, 1983.

[111] F. Hayatee, “The equivalent series resistance in electrolytic capacitors,” Electrocompo-


nent Science and Technology, vol. 2, pp. 67–72, 1975.

72
[112] M. L. Gasperi, “A method for predicting the expected life of bus capacitors,” in Con-
ference Record of the 1997 IEEE Industry Applications Conference (IAS 1997), 1997,
pp. 1042–1047.

[113] Determining End-of-life, ESR, and Lifetime Calculations for Electrolytic Capacitors
at Higher Temperatures, Aug. 2008.

[114] W. McAdams, Heat Transmission, 3rd ed. McGraw-Hill Book Co. Inc., 1954.

[115] M. L. Gasperi and N. Gollhardt, “Heat transfer model for capacitor banks,” in The
1998 IEEE Industry Applications Conference (IAS 1998), 1998, pp. 1199–1204.

[116] S. Parler Jr, “Deriving life multipliers for electrolytic capacitors,” IEEE Power Elec-
tronics Society Newsletter, vol. 16, no. 1, pp. 11–12, 2004.

[117] “Capacitors Age and Capacitors Have an End of Life,” Emerson Network Power, Tech.
Rep., Aug. 2008.

[118] J. Nash, “Biaxially Oriented Polypropylene Film in Power Capacitors,” Polymer En-
gineering and Science, vol. 28, no. 13, pp. 862–870, 1988.

[119] H. Fuhrmann, M. Carlen, D. Chartouni, T. Christen, C. Ohler, and T. Votteler, “Novel


Measurement Methods for In-depth Analysis of AC Metallized Film Capacitors,” in
IEEE International Symposium on Electrical Insulation, 2004, pp. 568–571.

[120] M. Carlen, P. Bruesch, and R. Gallay, “Electrical Endurance Characterization of


Polypropylene Winding Capacitors for Traction Applications: a New Experimental
Method,” in Seventh International Conference on Dielectric Materials, Measurements
and Applications,, 1996, pp. 350–353.

[121] T. Umemura and K. Akiyama, “Accelerated-life Testing of Power Capacitor Dielectric


Systems,” IEEE Transactions on Electrical Insulation,, no. 3, pp. 309–316, 1987.

[122] M. Kemp, C. Burkhart, and T. Tang, “Lifetime Tests on a High Ohms/square Met-
alized High Crystalline Polypropylene Film Capacitor with Application to a Marx
Modulator,” in IEEE International Power Modulator and High Voltage Conference
(IPMHVC), 2010, pp. 647–650.

[123] “AVX medium power film capacitors for power applications,” AVX Corporation, Tech.
Rep., Jan. 2012.

[124] P. Michalczyk and M. Bramoulle, “Ultimate Properties of the Polypropylene Film


for Energy Storage Capacitors,” IEEE Transactions on Magnetics, vol. 39, no. 1, pp.
362–365, Jan. 2003.

[125] Q. Sun, J. Zhou, Z. Zhong, J. Zhao, and X. Duan, “Gauss-Poisson Joint Distribution
Model for Degradation Failure,” IEEE Transactions on Plasma Science, vol. 32, no. 5,
pp. 1864–1868, 2004.

73
[126] W. Mason, “Aging of the properties of barium titanate and related ferroelectric ce-
ramics,” Journal of the Acoustical Society of America, vol. 27, no. 1, pp. 73–85, 1955.

[127] I. Burn, Ceramics and Glasses, S. Schneider, Ed. ASTM, 1991, vol. 4.

[128] E. Loh, “Development of a model for voltage degradation of various dielectric mate-
rials,” IEEE Transactions on Components, Hybrids, and Manufacturing Technology,
vol. 4, pp. 536–544, 1981.

[129] J. Minford, M. Yatsko, and W. Baker, “Accelerated aging study of multilayer ceramic
capacitors,” in American Ceramic Society Annual Meeting, Cincinnati, Ohio, 1979.

[130] T. Prokopowicz and A. Vaskas, “Research and development, instrinsic reliability, sub-
miniature ceramic capacitors,” Sprague Electric Corporation, Tech. Rep., Oct. 1969.

[131] W. Minford, “Accelerated life testing and reliability of high-k multilayer ceramic capac-
itors,” IEEE Transactions on Components, Hybrids, and Manufacturing Technology,
vol. 5, no. 3, pp. 297–300, 1982.

[132] R. Munikoti and P. Dhar, “Highly accelerated life testing (halt) for multilayer ceramic
capacitor qualification,” IEEE Transactions on Components, Hybrids, and Manufac-
turing Technology, vol. 11, no. 4, pp. 342–345, 1988.

[133] A. E. Clifton II, Concise Encyclopedia of System Safety: Definition of Terms and
Concepts. Hoboken, New Jersey: John Wiley and Sons, 2011.

[134] J. Lapp, “Power capacitor reliability concepts,” in IEEE/PES Transmission and Dis-
tribution Conference and Exposition, 1979, pp. 544–548.

[135] S. Speaks, “Reliability and MTBF Overview,” Vicor Reliability Engineering, Tech.
Rep., 2005.

[136] E. Vargas, J. Bianco, and D. Deeths, Sun Cluster Environment: Sun Cluster 2.2.
Prentice Hall, Apr. 2001.

74
DISTRIBUTION:

MS ,
,
1 MS 0899 Technical Library, 9536 (electronic copy)

75
76
v1.38
View publication stats

You might also like