You are on page 1of 315

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

net/publication/285594381

Characterization and analysis of photovoltaic modules and the solar resource


based on in-situ measurements in southern Norway

Thesis · November 2012


DOI: 10.13140/RG.2.1.1515.6568

CITATIONS READS

20 2,126

1 author:

Georgi Hristov Yordanov


KU Leuven
54 PUBLICATIONS   582 CITATIONS   

SEE PROFILE

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

PhD work on temperature coefficients of multicrystalline compensated silicon solar cells View project

PV OpMaat View project

All content following this page was uploaded by Georgi Hristov Yordanov on 04 December 2015.

The user has requested enhancement of the downloaded file.


Georgi Hristov Yordanov

Characterization and Analysis


of Photovoltaic Modules and the
Solar Resource Based on
In-Situ Measurements in
Southern Norway

Thesis for the degree of Philosophiae Doctor

Trondheim, November 2012

Norwegian University of Science and Technology


Faculty of Information Technology, Mathematics
and Electrical Engineering
Department of Electric Power Engineering
NTNU
Norwegian University of Science and Technology

Thesis for the degree of Philosophiae Doctor

Faculty of Information Technology, Mathematics and Electrical Engineering


Department of Electric Power Engineering

© Georgi Hristov Yordanov

ISBN 978-82-471-4028-4 (printed ver.)


ISBN 978-82-471-4029-1 (electronic ver.)
ISSN 1503-8181

Doctoral theses at NTNU, 2012:350

Printed by NTNU-trykk
3

ABSTRACT

In broad terms, this thesis is concerned with outdoor testing of individual


photovoltaic (PV) modules and characterization of the solar resource in Southern
Norway. The work aims to evaluate the irradiance and temperature dependency of the
efficiency for various types of PV technology (mostly crystalline-silicon, c-Si), and to
get insight into and explain using device physics the performance differences within the
generic c-Si class which have been reported in the literature. In addition, the thesis
discusses and compares the local irradiation results obtained in this and earlier studies,
and the predictions of the Photovoltaic Geographical Information System (PVGIS) of
the European Commission’s Joint Research Centre (JRC).
Two research installations have been built for regular measurement of current-
voltage characteristics (I-V curves) of ten c-Si and one thin-film CuInGaSe2 (CIGS)
modules, and the key environmental variables (solar irradiance in the plane of the
modules, ambient and module temperatures). Data recorded during one full year (2011)
are used to characterize each of the tested devices in terms of identification of their I-V
curve model parameters and their relative efficiency (with respect to that at standard
testing conditions, STC). The solar resource in the city of Grimstad is analyzed from
several perspectives including the annual irradiation, its seasonal distribution,
magnitudes and durations of overirradiances due to cloud enhancement, the optimal
azimuthal orientation of a PV array, and solar energy lost due to clouds. This is
complemented by a study of local meteorological data from fifteen consecutive years
which are publically available on the Internet.
The work reveals and quantifies the opposite effects of the series resistance RS
and the ideality factor n on a module’s performance at different levels of illumination. It
shows the inevitable trade-off faced by PV manufacturers when optimizing the cell
structure, e.g., using two instead of three front busbars in screen-printed c-Si cells
increases the annual energy yield per kilowatt-peak installed at the expense of reduced
STC power. These results contribute to a better scientific understanding of PV
performance and to the industry by providing quantitative guidelines for product design.
A key finding reported in the thesis is that cloud enhancement of sunlight can
boost the plane-of-modules irradiance well beyond the extraterrestrial levels even at
such a high latitude as Southern Norway, almost at sea level. A burst of 1528 W/m2 has
been recorded in the year 2012. Statistics of overirradiance magnitudes and durations
are presented. The work also shows that the clouds block nearly a half of the generally
available solar energy in Grimstad, and that the useful fraction exceeds the predictions
of the PVGIS by 10 %. These findings increase the scientific knowledge of the
terrestrial sunlight in general and the PV potential in Southern Norway in particular.
4

Furthermore, the thesis documents new or improved methods for I-V curve
analysis and for measurement of the environmental variables: identification of model
parameters; calculation of the equivalent cell temperature (ECT) from open-circuit
voltage; and the use of the short-circuit currents of multiple co-planar PV devices to
improve the accuracy of irradiance measurements. The thesis also demonstrates some
limits of the applicability of the widely used 1- and 2-exponential I-V curve models.
5

To my beloved family, friends and teachers who made me who I

am – my parents, my elder sister, my wife Diana, my wonderful

friend Svetoslav Velinov, and my teachers in physics, mathematics

and biology Aneta Etimova, Vanya Boyadzhieva, Radoslav

Etimov and Jenna Staykova. I thank you and love you!

““My frogs are blue!”” 

“Writing for Scholars”

Lynn P. Nygaard

By accident, the thesis contains about 100π pages.

Request a digital copy from the author at g.yordanov dir.bg.


6

ACKNOWLEDGMENTS

I would like to express my gratitude to many people who kindly supported me


while I was doing my studies and the research presented in the thesis.
I first thank my supervisor Ole-Morten Midtgård for his enormous support,
hospitality, positive attitude, and the great deal of encouragement during all the stages.
He has always told me to believe in myself since I was doing a great job, and helped me
overcome my doubts and keep going. He has been also a wonderful teacher in scientific
writing; a key co-author of many joint papers; a very successful applicant for research
funding; and a good team- and institution-builder. I also appreciate the help, support and
advice of my co-supervisor at NTNU Lars Norum. Giving me access to a couple of
preinstalled PV modules at the early stage of my LabVIEW™ software development
has helped a lot. Tor Oskar Saetre from the University of Agder (UiA) is recognized for
sharing valuable Matlab™ know-how; for giving useful advice and training before my
very first oral conference presentation; for inviting several research groups from abroad
to UiA, thus letting me practice my presentation skills and receive feedback on my
work. He has also been a very fruitful co-author of conference and journal papers.
The continuous help and support of Jan Ove Odden and Torfinn Buseth from
Elkem Solar is especially appreciated. It was the data from their research installation in
Kristiansand where I saw the first independent confirmation of the really strange
overirradiances. Later on, Torfinn shared his observation of a burst exceeding 1800
W/m2 at Elkem’s station in equatorial Kenya. He has also provided several important
papers cited in the thesis. Jan Ove and Torfinn kindly shared all the reports from the lab
at PI-Berlin, where all the PV modules provided by Elkem Solar were independently
tested.
Erling Andresen was a very kind host at the Kvadraturen Skolesenter in
Kristiansand. The technical staff and the students from Kvadraturen Skolesenter
contributed in an important way to my experiment there by building the mechanical
frames and support structures for the PV modules, and by doing the electrical cabling,
all done very professionally. Henrik Kofoed Nielsen from UiA is acknowledged for his
efforts and mediation in establishing the research platform on the roof of UiA’s new
campus in Grimstad. I also thank Anne Gerd Imenes (with UiA and Teknova) very
much for her fruitful collaboration, including joint publications and many useful
discussions; for providing relevant literature in Norwegian language; for her hospitality,
as well as for sharing the Norwegian culture.
I am also grateful to my colleagues Abozar Alabakhshizadeh, Chee Lim Nge,
Deepak Verma and Muhammad Tayyib who, in the cold winds, installed and connected
the last two modules in Grimstad, while I was celebrating a warm New Year with my
family in Bulgaria. Deepak and Muhammad also dismantled the very same modules one
7

year later. My special gratitude goes to Abozar, who turned on the electronic loads on
the New Year’s Day and tied all my cables tightly and very professionally. Chee Lim
also provided the DC power supply and the probes used to measure the resistance of the
cables connecting the modules with the electronic loads. I must admit that the idea of
this simple experiment (described in Chapter 3) crystallized after discussions with Chee
Lim and Hans Georg Beyer. The latter is acknowledged for his technical expertise,
numerous discussions, co-authorship, and defending his points in the German way.
Chee Lim helped me a lot in the transportation of heavy and bulky equipment between
Grimstad and Kristiansand. I highly appreciate all the discussions we have had over the
years, especially those that we had during our course studies in Trondheim. I also
appreciate his distinguished social skills and the happy parties that we had, usually upon
his suggestion. Nils Randulf Kristiansen directed me to the Grimstad-based company
Alternativ Energi which, to my big surprise, marketed PV modules with back-contact
cells from SunPower. By that time I had given hope to have this technology as part of
my experiment. Thanks a lot, Nils!
Financial support for the present work was provided by Elkem Solar (now a
Bluestar company), the Research Council of Norway, the University of Agder, and the
Municipality of Kristiansand. NTNU is acknowledged, among the other things, for
gifting me with the wonderful booklet “How to get a PhD” [Phi05] which contains
comprehensive definitions of a doctorate, the research process and the constituents of a
thesis. Eva Schmidt from NTNU is acknowledged for her continuous administrative
support.
As a Christian, I thank our Savior and our Father very much for creating a
universe so interesting and wonderful; for giving us great love; and for the everyday
care, support, guidance and forgiveness. Finally, I would like to express deepest thanks
to my wife Diana for her love, faithfulness and patience during the long periods of my
absence in the past four years.
8

CONTENTS
1 INTRODUCTION ………………………………………………………... 11
1.1 Aims and objectives ………………………………………………….... 11
1.2 Outline of the thesis ……………………………………………………. 12

2 PERSPECTIVES ………………………………………………………… 15
2.1 PV market development ……………………………………………...... 15
2.2 PV potential in Southern Norway …….……………………………….. 16

3 TEST SETUP AND DATA SETS ………………………………………… 19


3.1 Chapter aims and objectives …………………………………………… 19
3.2 Outdoor vs. indoor characterization of PV modules …………………... 19
3.3 Best practices for outdoor PV testing …………………………………. 20
3.4 Choosing PV technologies for testing ………………………………… 27
3.5 New vs. old experimental setup in Grimstad: Improvements ………….. 30
3.6 Test site and module configuration ……………………………………. 35
3.7 Inspection, observation and maintenance …………………………….. 39
3.8 Data acquisition hardware …………………………………………… 40
3.9 LabVIEW code; data types, formats and pre-processing ……………... 43
3.10 Daily visual data inspection ………………………………………….. 55
3.11 Post-processing of raw I-V curves ………………………………….. 59

4 IRRADIANCE ……………………………………………………………. 61
4.1 Chapter aims and objectives …………………………………………... 61
4.2 Sunlight at Earth’s surface ……………………………………………. 61
4.3 Measurement methods and sensors; data sources …….……………….. 63
4.4 Analysis of local meteorological data …………………………………. 67
4.5 Irradiance measurements in the present work …………………………. 73
4.6 Effects of low temporal resolution of irradiance data ………………… 78
4.7 Data recovery and self-referenced irradiance …………………………. 84
4.8 Solar and cloud resources measured in the year 2011 ………………… 98

5 TEMPERATURES ………………………………………………………. 122


9

5.1 Chapter aims and objectives …………………………………………… 122


5.2 Best practices in module temperature measurements .………………... 122
5.3 Back-of-module and ambient temperatures …………………………... 125
5.4 Equivalent cell temperature (ECT) vs. back temperature .…………... 129
5.5 Temperature coefficients ………………………………………………. 146
5.6 Thermal resistances of the tested modules …………………………… 153
5.7 Discussion …………………………………………………………….. 154

6 I-V CURVE ANALYSIS …………………………………………………. 159


6.1 Chapter aims and objectives …………………………………………… 159
6.2 I-V curve analysis and modeling …………………………………….... 159
6.3 Review of the existing I-V curve models ……………………………. 161
6.4 Methods for determination of RS, n and I0 …………………………….. 175
6.5 Series resistances of the tested modules ……………………………… 190
6.6 Ideality factors determination ………………………………………... 200
6.7 Local ideality factor analysis ……………………………………….... 209
6.8 Reverse saturation currents …………………………………………….. 224
6.9 Chapter summary ………………………………………………………. 226

7 POWER RATING AND MODELING ………………………………… 229


7.1 Chapter aims and objectives …………………………………………… 229
7.2 Review of prominent PV performance models ……………………….. 229
7.3 Power rating of the tested modules ……………………………………. 234
7.4 Modeling performance in terms of relative efficiency ………………... 242
7.5 Chapter summary ………………………………………………………. 257

8 CONCLUSIONS ……………..…………………………………………… 258


8.1 Main scientific contributions …………………………………………… 259
8.2 Main contributions to PV industry …………………………………….. 260
10

APPENDICES

A List of papers 261

B Matlab™ code used to calculate the solar position, the clear-sky air mass (AM) and
the angle of incidence (AOI) of direct sunlight on the tested modules
264

C ECT error analysis with respect to the ideality factor n 267

D Matlab™ functions used to extract RS, n and I0 from one I-V curve 268

E Matlab™ scripts used to determine RS(I) with the double-light method 270

F Derivation of Eq. (6.5) from Eq. (18) in [Gre03] 273

G Effect of the temperature coefficient of ISC on the relative efficiency ηREL


275

H Derivation of the coefficients k1 and k2 of the relative efficiency model 276

I List of symbols 279

BIBLIOGRAPHY 284
11

CHAPTER 1 INTRODUCTION

1.1 AIMS AND OBJECTIVES


The present work aims to investigate the potential for photovoltaic (PV)
technology as a source of electricity in Southern Norway, and to identify causes for
performance differences between various crystalline-silicon (c-Si) technologies. A main
objective is to characterize and compare the solar resource available in the Grimstad
area to results found in prior studies, as well as to predictions of the Photovoltaic
Geographical Information System (PVGIS) of the European Commission’s Joint
Research Centre (JRC) [PVGIS]. If available, also local meteorological data should be
analyzed. Another key task is to measure and model the performance of appropriate PV
module types, thus testing the behaviors found in the literature and those assumed in
software tools such as the PVGIS.
To achieve these goals two outdoor test stations have been built, one in the city
of Grimstad and the other in the city of Kristiansand. The test stations have been used
for detailed studies of both the solar resource (the energy source itself), and various
photovoltaic modules.
The methodology has been to operate the modules continuously at the maximum
power point (MPP) and to record very detailed current-voltage characteristics (I-V
curves) at regular intervals. At the same time the solar irradiance in the plane of the
modules has been measured by different methods, temperatures have been recorded, and
care has been taken to clean the modules regularly. In addition, a TV-camera and time
stamping has been used in one of the test stations in order to be able to identify causes
of anomalies.
Will the rapid introduction of PV in Germany and many other states suffice for
preventing the climate catastrophe and its dire consequences? This thesis answers many
other questions but this one. However, acting after seeing the bad thing happened would
be too late. The increasing environmental and geopolitical awareness was the author’s
main motivation to specialize in a renewable energy technology. This PhD research is a
part of the project “End Use of Photovoltaic Technology in Norway” co-funded by the
Research Council of Norway, Elkem Solar AS (producer of solar-grade silicon for the
PV industry, SoG-Si), the Municipality of Kristiansand, and the University of Agder
(UiA). The project’s overall objective is to investigate the hypothesis that large-scale
PV generation is viable in Norway as well. This requires studies of both the local solar
resource and the suitable PV technologies. The investigation of power electronics
interfaces between PV generators and the utility grid is the working topic of another
PhD researcher [Nge09, Nge10], and is outside the scope of the present work. However,
the obtained environmental and PV module data are generally available to other
12

researchers and groups from the PV, power electronics and weather/climate
communities, as well as to policy makers.
Some of the results presented in the thesis have been presented at peer-reviewed
international conferences and journals related to PV and renewable energy. A list of
publications by the present author, produced in cooperation with his supervisors and
other research colleagues, is given in Appendix A. The list includes 20 conference
papers and 5 journal papers. The present author is the main author of 3 of the journal
papers and 12 of the conference papers.

1.2 OUTLINE OF THE THESIS


Chapter 2 starts with a description of the geopolitical and economic contexts of
the present work. The PV potential in Southern Norway is then briefly introduced.
Chapters 3 through 5 contain most of the data theory in the present work.
Chapter 3 begins with a review of the best practices for outdoor testing of PV modules,
and specifies the improvements of the new experimental setup in Grimstad versus an
older one described in [Mid10]. Then, the current status of the terrestrial PV module
market is discussed in the context of selecting devices to be tested in Grimstad. Later
on, a description of that experimental setup is given, including details on the outdoor
installation, the data acquisition system, and the maintenance activities performed in the
course of the year 2011. Some challenges faced during the measurement campaign are
also discussed.
Chapter 4 focuses on the measurement of solar irradiance in the plane of the PV
modules and characterization of the local solar resource. It starts with a review of the
available methods and established practices for measuring of terrestrial irradiance.
Local meteorological records of the global horizontal irradiance (GHI) are analyzed,
focusing on the year-to-year variability and the seasonal distribution of annual
irradiation. The sensors and the methods used in the present work are then described,
and their limitations are outlined. Then, the effects of sampling irradiance with 20-
minute and with 1-minute resolution on the monthly irradiation uncertainty are
investigated. The results are discussed in the context of the monthly irradiation values
presented in [Mid10] and the ones obtained in the present work from the 1-minute data
set. Later on, the method used for recovery of contaminated irradiance data is presented.
Chapter 4 continues with a presentation and analysis of the measured seasonal
distribution of solar irradiation. The results obtained from the present experimental
setup and from local meteorological data are compared to those of prior studies, and to
long-term predictions of the PVGIS. Statistics of the measured irradiance values are
given, and special attention is paid to the registered overirradiance events. The ‘cloud
resource’ available in Grimstad and the optimal azimuthal orientation of a PV array are
also investigated.
Chapter 5 focuses on measuring the ambient and module temperatures, and on
the evaluation of the temperature effects on the open circuit voltage (VOC), which is
13

used in the calculation of the equivalent cell temperature (ECT). The chapter begins
with a review of the established practices for measuring a PV module’s operating
temperature and discusses the associated challenges. The calculation of equivalent cell
temperature (ECT) with the standard procedure IEC 60904-5:1993 is then presented.
The measurements of daytime and nocturnal back-of-module temperatures are described
next, and their accuracy is analyzed. Later on, an error analysis is given of the standard
ECT calculation procedure with respect to each of the involved parameters. The
challenge of a variable ideality factor is then addressed and an improved equation is
derived, which enables ECT calculation for devices with an ideality factor that varies
linearly with irradiance and module temperature. Two methods used for determination
of the temperature coefficient β of VOC are presented and discussed next, and the
thermal resistances of the modules are evaluated. The chapter ends with a discussion of
the results and the methods, as well as the questions which remain open regarding them.
Chapter 6 focuses on I-V curve analysis and modeling, as well as on the
derivation of the series resistance RS, the ideality factor n, and the reverse saturation
current I0, which are defined in the 1-exponential model of PV cells and modules. The
chapter begins with a review of the classic 1- and 2-exponential models. Different
versions of the latter are discussed with respect to the values assumed for the two
ideality factors. Possible complications, such as a variable series resistance and non-
linear shunts, are considered. The Ohmic shunt-current term is tested with experimental
data and is generally refuted in the case of modules. The analytical Saetre-Midtgård-Das
model is briefly discussed, but not reviewed in detail. After that, different methods for
evaluation of model parameters are reviewed and discussed. Then, a new method for
determination of RS proposed by the present author is explained, and results are
presented for the eight c-Si modules and the CIGS module tested in Grimstad. The
alternative double-light method is used to probe for a current dependency of RS. Later
on, results for n and I0 are presented, and their respective dependencies on irradiance
and device temperature are modeled. The classic local ideality factor m and its RS-
corrected modification are then used (in the form of m-V and m-I plots) to analyze the
behavior of the ideality factor along the I-V curve, in particular near open circuit and
around the MPP. At last, a new method for estimation of both RS and n from a single I-
V curve, which was discovered using this methodology, is reported.
The main topic of Chapter 7 is rating and modeling of the power performance of
PV modules, as well as identifying quantitative and qualitative relationships of the
performance to the I-V curve parameters. First, various performance models are
reviewed. Then, the relative efficiency model implemented in the European PVGIS tool
is discussed, and the model coefficients are determined for eight c-Si modules and for
the CIGS module. The modules are first rated at standard testing conditions (STC), and
the possible methods for such rating from outdoor data are compared. An improvement
14

of the 'Northern' method is proposed. The modeled relative efficiencies of the modules
with respect to STC are then compared and discussed in the context of final energy
yield per kilowatt-peak of rated PV power (kWh/kWP). Data from the two mc-Si
modules installed in Kristiansand are also presented. Theoretical expressions (based on
the 1-exponential I-V curve model) are derived for two key coefficients (k1 and k2) of
the relative efficiency model, and their limits of applicability are discussed. The
theoretical and fitted values of k1 and k2 are then compared, as well as the predicted and
measured energy yields.
The thesis ends with a brief chapter on the main conclusions.
15

CHAPTER 2 PERSPECTIVES

2.1 PV MARKET DEVELOPMENT

Figure 2.1: Atmospheric CO2 record of the US National Oceanographic and


Atmospheric Administration [USD11]. The applied extrapolation is exemplary yet
possible.

This research on solar resource and photovoltaic (PV) modules started in the
context of globally increasing environmental and economic concerns. The greenhouse
effect of the anthropogenic carbon dioxide (CO2) has been recognized with a very high
level of scientific understanding [IPC07]. The trend of increasing its atmospheric
concentration does not show signs of slowing down (see Fig. 2.1). It has been estimated
that mankind must start reducing its greenhouse gas emissions (not just CO2) by the
year 2015 in order to prevent a dangerous climate change [McG08].
The need of a more sustainable basis for economic growth has been widely
recognized worldwide. Over the time, subsidiary regimes for renewable electricity
generation have been introduced in Europe, the US and elsewhere. The most famous
16

governmental incentives for PV are the German 100’000 solar roofs program and the
American 1’000’000 solar roofs initiative [Hof06]. Investments in rooftop and utility-
scale PV systems have risen exponentially over the past decade, with a combined annual
growth rate of approximately 50 % [Min10].
In Norway, the main source of electricity is hydropower. Feed-in tariffs (FIT)
for domestic and corporate PV generation similar to those in Germany have not been
introduced so far. Norwegians use PV modules for various stand-alone applications: on
boats, at remote cottages in the mountains or on the many islets, and for powering
equipment with modest requirements. However, there are few larger scale, grid-
connected installations throughout Norway, and these serve for demonstration and
education rather than for commercial electricity production.

2.2 PV POTENTIAL IN SOUTHERN NORWAY

Figure 2.2: A map from the European PV geographical information system


[PVGIS]. The color code can be translated as either the annual global solar
irradiation or the yearly yield from a PV system. The star indicates the location of
the present test sites (Grimstad, at latitude 58°20’N, and Kristiansand, at 58°09N).
17

Norway’s territory (except the Svalbard archipelago in the Arctic) is covered by


the European PV geographical information system [PVGIS, Sur07] (see Fig. 2.2). The
online tool was established in the past decade to facilitate the PV community and users-
to-be in the assessment of the potential annual energy yields from fixed and sun-
tracking PV systems. Earlier studies of the solar resource in Norway revealed that the
annual horizontal irradiation generally increases with decreasing latitude and with
increasing distance from the western coast. This is somewhat difficult to see in the
PVGIS irradiation map. However, Olseth and Skartveit [Ols86] published an isoline
map showing this more clearly. That map suggests that the East Agder region receives
the highest annual irradiation, with a peak in the coastal area of Grimstad (the test site
of the present work). The latter was the location chosen for the National Center for
Renewable Energy known among Norwegians as Energiparken [NCR11]. This general
observation about the solar resource in Norway is also found in reports from domestic
companies and institutions involved in renewable energy [Kan01, NVE07]. The latter
assessed the maximum Norwegian annual horizontal irradiation at 1000 kWh/m2. Even
more solar energy is therefore expected in an optimally inclined plane facing south.

Figure 2.3: Monthly solar irradiation measured in the year 2005 in Grimstad,
Norway versus PVGIS estimates [Mid10].

A recent study of the solar resource in Southern Norway measured an annual


irradiation exceeding 1100 kWh/m2 in an inclined, south-facing plane [Mid10]. This
result exceeded by 10 % the estimate from the PVGIS tool. Furthermore, the obtained
18

monthly distribution showed much greater than expected irradiation in the spring, as
well as in January (see Fig. 2.3). Indeed, a higher uncertainty of the PVGIS estimates
can be expected close to coastlines (as in Grimstad) and in regions with lower density of
represented meteorological stations as from Norway [Sur08b, Hul10]. In addition, these
predictions are based on weather data from many years and so the year-to-year
variability (which can reach 10-15 % in some regions [The10, Ham12]) is averaged out.
Even larger variations (up to 35 %) can be expected for individual months, mainly in
winter [The10]. The results obtained for Grimstad in the year 2005 revealed that much
higher variations may be possible. However, the local study used data from a single
year, with a rather poor temporal resolution of 20 minutes. Also, the average uptime of
the data acquisition system was only about 73 %. Furthermore, the measured current-
voltage characteristics (I-V curves) contained rather small number of data points
(between 10 and 50). Consequently, the flat region and the MPP region (the knee) had
rather poor resolutions, especially at low irradiance levels (see Fig. 2.4).

Figure 2.4: PV module I-V curves from the earlier experimental setup [Mid10].
19

CHAPTER 3 TEST SETUP AND DATA SETS


3.1 CHAPTER AIMS AND OBJECTIVES
The present work aims at comprehensive characterization of the selected devices
and takes the outdoor testing approach. The aim of Chapter 3 is to review literature on
the established best practices for I-V curve measurement and to describe in detail the
test PV installation located in Grimstad, the implemented data acquisition system, the
data types being collected, as well as their pre- and post-processing. The improvements
with respect to the older experimental setup described in [Mid10] are outlined. The
present Chapter reviews the current status of the global PV market and technologies
suitable for large-scale electricity generation in Southern Norway. An important task is
the evaluation of thermal effects from a 1-minute fault of the implemented MPP-
tracking (MPPT) algorithm on the recorded performance-related module parameters.
Chapter 3, together with Chapters 4 and 5 contains most of the data theory of the
present work.

3.2 OUTDOOR VS. INDOOR CHARACTERIZATION OF PV MODULES


The energy delivered by a PV module over its lifetime depends on how efficient
the device is at different operating conditions (assuming MPP operation), on the relative
occurrence frequency of the various conditions, as well as on the mode and rate of its
long-term degradation [Sak03, Hul08, Hul10, Jor10, Mid10, Pin10, Jor11]. Knowing
the area A of the device, the in-plane irradiance G and the module’s PMAX give the
instantaneous efficiency η = PMAX/(GA). Modules are first characterized in terms of
PMAX (for given operating conditions), which is usually obtained by measuring the
complete I-V curve. This can be done indoors, using an artificial light source (suitable
for high-throughput inline characterization in the industry) or outdoors, relying on
availability of enough natural sunlight. The latter approach is used mainly by calibration
and research laboratories, but some use both [Rum06, Vir08, Jag09, Got10, Moh10].
For example, the IEC 61853-1 standard specifies a matrix of seven irradiance
levels and four module temperatures (G-T matrix), for which a PV module’s
performance must be known [Roy08, Kok11]. However, some of the desired conditions
may not occur naturally outdoors in a given location, and may be difficult to create
indoors. A possible solution is to use I-V curve translation procedures. On the other
hand, indoor testing is usually performed with decaying light and the measured curves
need correction [Mül08], whereas outdoor measurements can offer very stable and
uniform illumination. Besides for power and energy rating, I-V curves can be used also
for analysis of the physical properties of a given PV technology. Depending on the
application and on the available hardware, different level of detail and accuracy may be
targeted and eventually achieved, e.g. the total number of data points in each curve and
20

the density of points in different parts of the curve. Very noisy I-V curves are
sometimes obtained [Goo09].

3.3 BEST PRACTICES FOR OUTDOOR PV TESTING


Several laboratories (mostly European) which perform outdoor testing of PV
have decided to harmonize their procedures in order to improve the accuracy and
reliability of module characterization [Jag09]. Consequently, a better agreement has
been achieved between the partnering institutions’ round-robin results [Moh10]. The
addressed issues include preconditioning (monitoring of initial degradation and
stabilization), mounting accuracy, module and sensor cleaning (from dirt, snow and
ice), I-V curve sweeping time, irradiance stability, frequency of I-V scans, module
loading between the scans, data file format, as well as inspection, validation and
filtering of the recorded data. The data acquisition system used in the present test setup
has been built by following closely (when possible) the guidelines given in [Jag09].

3.3.1 PRECONDITIONING (LIGHT-SOAKING)


Most of the modules tested in the present work are made of c-Si, which is
characterized with very short stabilization times [Jag09]. The values found in the
literature range between 5 and 10-12 hours under STC irradiance and spectrum [Goe03,
Gos11]. All modules have been exposed to natural sunlight since early December, 2010
except for the two specimens based on Elkem Solar Silicon (ESS™), which were
installed and connected on 31 December. The monthly solar irradiation results for 2011
presented later in Chapter 4 show that the modules have received about 75 kWh/m2 of
illumination in January and February. Indeed, the direct sunlight spectrum has been
redder than the one specified for the STC, but on the other hand the c-Si modules have
received at least twice the irradiation needed for stabilization [Mun11]. Their I-V curve
analysis, power rating and performance modeling described later in Chapters 6 and 7 do
not use data recorded before March, 2011. It is therefore concluded that the c-Si
modules have undergone more than sufficient preconditioning until then.
Light-soaking behavior of CIS/CIGS modules has not been identified
unambiguously so far. Some types improve after exposure to light [Ken06, Got10,
Hul10, Dun12]. The controversial information in the literature seems to originate
partially from technological differences between devices commercialized over the past
two decades [Cue08, Jag09, Gos11]. The two strongest improvements due to light
soaking are in the module’s VOC and FF [Ken06]. The observed increase of the open-
circuit voltage has been found to result from forward-biasing of the cells rather than
from illumination. The improvements are reversible and the stabilization times of light-
soaking and dark relaxation may be comparable, ranging from minutes to hours
depending on cell temperature [Sas93, Ken06]. Such relatively short times imply that in
outdoor conditions both effects will occur repeatedly every day and night. Even at very
low irradiances, the CIGS module tested in the present work has a relatively high VOC.
21

Therefore, the increase in VOC is expected to occur even on overcast days, as long as the
module is forward-biased (i.e., if current is drawn from the module). IEC 61646
prescribes a procedure for preconditioning of thin-film PV by light-soaking them
repeatedly with 43 kWh/m2 of cumulative irradiation until the power changes become
less than 2 %, which seems optimal for a-Si only [Cue09, Cue10]. Therefore, the studies
aiming to establish preconditioning steps suitable for CIS/CIGS and CdTe have
continued [Del11]. It has been found that some CIS/CIGS need at least 100–120
kWh/m2 of cumulative light-soaking for stabilization to occur. In the context of the
present work, these findings imply that the tested CIGS module must have reached
stability by mid-March, 2011. This rough estimate excludes the module’s exposure in
December, 2010, as the typical December irradiation in Grimstad is within 20 kWh/m2
(see Figs. 2.3 and 4.34).

3.3.2 OPTIMAL I-V CURVE SWEEP TIME


Ideally, the shortest possible time should be used to scan (sweep) the outdoor I-
V curves of a PV module, so that the effects of varying irradiance are minimized
[Jag09]. The harmonized procedures recommend that the optimal time is determined
individually for each module by comparing I-V curves recorded at different speeds at
stable environmental conditions. On the other hand, too short a sweep time leads to
distorted curves. I-V curves can be swept starting from short circuit until open circuit is
reached (‘forward sweep’, not to be confused with forward-bias I-V curves), or starting
from open circuit until short circuit is reached (‘backward sweep’, not to be confused
with reverse-bias I-V curves). Taking both forward and backward curves at different
scanning speeds has been helpful in identifying the optimal sweep time for various PV
technologies [Vir08, Bel11]. For mainstream c-Si modules, sweep times of a few tens of
milliseconds do not result in notable errors in the measured PMAX or changes in the
curve shape. Some researchers tend to use an order of magnitude shorter times (1-2 ms)
in indoor measurements with pulsed solar simulators in order to minimize the light
intensity variation [Mül04]. However, some high-efficiency (heterojunction) c-Si
modules require relatively long sweep times of 120-400 ms [Vir08, Bel11]. Such is the
Sanyo HIT module tested in the present work. More recently, a way has been found to
obtain accurate I-V curves of HIT devices using pulsed light [Vir12]. For CIS/CIGS
devices, the minimum sweep time seems to be of order 5-10 ms, whereas single-
junction and multi-junction a-Si modules may need several times slower sweeping
[Kuu08, Vir08]. In the latter study, however, short-term light-soaking effect on CIS
performance has compromised the determination of a minimum sweep time. For all
times above 5 ms, the forward and the backward I-V curve have had nearly equal PMAX.
In addition, the maximum power curves corresponding to forward and backward
sweeping have not separated at lower sweep times (unlike e.g. the a-Si modules where
the capacitive effects distort the I-V curves swept too quickly). This observation
suggests that CIS/CIGS modules do not suffer from capacitive effects even at sweep
22

times as low as 1 ms. A relatively slow I/O hardware (with respect to the desired data
acquisition rate) may become a ‘bottleneck’ imposing even higher sweep times
[Yor10b]. An optimal value should therefore be chosen, one which is a good trade-off
between all applicable considerations. This is also the case in the present test setup. I-V
curves of ten PV modules are measured simultaneously in order to ease later
comparisons. Repeated sweeps are performed in moments of significant irradiance
variations or when incomplete I-V curves are recorded for some of the modules.
Measuring all ten modules in turn could be very time-consuming and therefore
impractical, given the desired sweeping frequency (every minute). On the other hand, a
rather high number of data points per I-V curve has been chosen. Eventually, the limited
sampling rate of the I/O board recording most of the data has imposed a minimum
sweep time of about 1 s. A similar value was used in outdoor measurements at the
European Solar Test Installation (ESTI) [Mül09].

3.3.3 OPTIMAL NUMBER OF DATA POINTS


The harmonized European outdoor testing procedures do not specify an optimal
number of data points per I-V curve [Jag09]. The numbers found in the literature (for
cells, as well as for modules) differ by three orders of magnitude, ranging between a
few tens or even less than ten [Mia83, Fis00, Che01, Kun04, Hao05, Mid10, Mun11,
Vir12] and several thousand [Hyv03, Mül09, Yor10b]. Indeed, the estimation of key I-V
parameters such as ISC and VOC does not require many points. In case the points around
the MPP are scarce, interpolation must be applied for accurate determination of VM, for
which a maximum uncertainty of ±0.5 % is suggested [Jag09]. On the other hand, the
density of points may differ in different parts of the curve (see Fig. 3.1). For example,
using a capacitive load for I-V curve sweeping has this side effect. When a discharged
capacitor is connected to an illuminated PV device, the voltage across it rises at a rate
defined by the capacitance and by the short-circuit current of the cell/module. As the
voltage builds up, the MPP region (the ‘knee’) of the device’s I-V curve is reached. At
even higher voltages, the current delivered by the PV device drops significantly
compared to ISC, and the capacitor is being charged at a much lower rate. If the I-V data
are sampled with a fixed sampling rate, the recorded I-V curves will have a much higher
density of points around open circuit than at higher currents. Thus, the total number of
points in such curves is less informative. The curve in Fig. 3.1 contains 101 points and
was obtained with a commercial I-V curve tracer [Ser10a].
Experimental I-V data may contain significant amount of noise [Goo09]. Fig.
3.2 shows a closer look at the points clustered near open circuit in the curve given in
Fig. 3.1. On the scale of 100 mV, the noise present in the data is clearly seen.
Nevertheless, the scatter is small compared to module’s VOC (well below 1 %). The
example is given to illustrate the typical noisiness of real data, which may require the
application of smoothing techniques. Therefore, using a high density of points along the
23

entire I-V curve may be beneficial. Automated data acquisition and the availability of
fast hardware (PC-based, as well as autonomous) allow repeated scanning of high-
resolution I-V curves. Mülleans et al. [Mül04, Mül08] measured 512 data points per I-V
curve in year 2004 and over 1000 data points four years later. Hüvärinen and Karila
[Hyv03] sampled 4096 data points in just 2 milliseconds in a study on c-Si cells. In the
present work, electromagnetic interference has been found to affect the measured I-V
curves since the early software development phase. It has been decided to use 4096 data
points per I-V curve, so that noise can be filtered out more easily on a later stage. This
resolution has revealed very clearly a sinusoidal mode with frequency around 800 Hz,
which has not been identified in I-V curves containing ‘only’ about 1000 points.

Figure 3.1: I-V curve measured with a commercial curve tracer [Ser10a]. The data
are kindly provided by Dr. Dezso Sera. Used with permission.

The I-V curves of c-Si modules can be very flat at low and intermediate voltages
(see Fig. 3.9 further in this Chapter). In the present setup, the curves are scanned with a
constant current step. At higher irradiances, this can result in greatly reduced point
density in the flat region of the I-V curve if too low a number of data points is chosen.
Consequently, the partial linear fitting used for short-circuit current determination may
24

be compromised. Another justification for the high number chosen is given by the fact
that bypass diodes may affect the curve even if the PV module is not shadowed. Minor
current mismatch between the cells can still lead to activation of bypass diodes at
intermediate and low voltages, see Fig. 3.3. The electronic loads are based on
semiconductor devices and have small but non-zero minimum (on-state) resistances.
Consequently, short circuit is not reached and ISC must be evaluated by extrapolating
the I-V curve to zero voltage. Therefore, a sufficient number of points is needed in this
part of the curve.

Figure 3.2: The I-V curve in Fig. 3.1 zoomed in around open circuit.

Parameter estimation by curvilinear fitting with the 2-exponential model was


said to require at least 100 data points per I-V curve (preferably 200) [Got97].
According to Sites and Mauk [Sit89], a large enough spacing must be maintained
between adjacent data points if differential techniques will be applied to the I-V data.
10–20 mV were recommended in the case of individual cells, which is equivalent to 30–
60 points per curve with VOC of about 0.6 V (assuming a c-Si cell at STC). In a module
with NS = 60 such cells connected in series, no more than 1800–3600 points per curve
should therefore be used. A much smaller voltage step (less than 1 mV) and curve
25

smoothing were recommended elsewhere for proper calculation of the incremental


conductance dI/dV [Wer88, Che01]. Such step size implies over 600 points for a cell
with VOC = 0.6 V, and tens of thousands of points in the case of a PV module. In the
present work, the raw I-V curves are smoothed and then any redundant (overlapping)
points are removed prior to discrete differentiation. Details on the used smoothing
algorithm are given later in Section 3.11. The processed curves contain roughly 2500
data points each. These are used to evaluate and plot the local ideality factor (the m-V
and m-I plots discussed later in Chapter 6). A similar number (2000) was used by
PHOTON Laboratory in their outdoor module tests, as well as by Pasan SA in a novel
large-area module tester [Pho11a, Pho11c].

Figure 3.3: An I-V curve with characteristic kinks due to bypass diodes activation
[Yor10a]. For greater detail, only the higher-current part of the curve is given.

3.3.4 FREQUENCY OF I-V CURVE SCANS


An interval between consecutive I-V curve sweeps no larger than 10 minutes is
recommended in the harmonized European procedures for outdoor testing [Jag09]. In a
study focused on spectral effects on module’s ISC, curves were scanned every minute
[Tsu08]. PHOTON Laboratory scans every second, whereas the sweep time is 10 ms
[Pho11a]. The sweep time becomes important when the scanning frequency is high,
26

since the modules are not operated at the MPP during the scans. If a realistic module
temperature is targeted, the sweep time must be much shorter than the interval between
consecutive I-V curve scans. This is also desirable if module’s power is fed (sold) to the
utility grid, which is not the case in the present experiment. On the other hand, a
relatively long sweep time imposed by some PV technologies like HIT or by hardware
limitations would set a lower limit on the scanning frequency. This is the case in the
present setup, in which the sweep time of 1 second requires I-V curves to be recorded at
intervals in the order of minutes. The earlier setup in Kristiansand was initially scanning
every 5 minutes [Yor10b]. As that work focused on a detailed analysis of the recorded
curves, the initial interval was found to provide insufficient amount of these.
Eventually, the frequency was increased to one curve per minute. Detailed analysis of I-
V curves is part also of the present work, and therefore the same frequency is used.

Figure 3.4: A P-I curve of the high-efficiency Sanyo HIT module from 07/01/2011.

Since no power is extracted from the modules during the measurement of I-V
curves, the main concern regarding the ratio of the sweep time and the scanning interval
is whether the modules are operated at realistic temperatures, as in the case of a
practical grid-connected PV system. The sweep time may seem too long, being 1/60
(about 1.7 %) of the scanning interval. One may think that each module operates at a
27

higher temperature compared to that in a system with constant MPP tracking, since part
of the available energy remains in the semiconductor material in the form of heat. This
expectation is ruled out by the following rough energy accounting. The two high-
efficiency modules can be close to 20 % efficient at very low temperatures in the winter.
One may roughly assume that another 10 % of the co-planar irradiance is reflected at
the interfaces between the cells and the ambient air [Mar01]. This value also accounts
for the module area which is not covered with cells. The remaining 70 % of the incident
power is lost during thermalization of excited charge carriers to the edge of the
corresponding energy band, and heats the crystal lattice. Furthermore, a constant
irradiance in the plane of the modules is assumed, as well as a constant wind speed and
direction. In case the modules are constantly operated at the MPP, they will eventually
reach a certain steady-state temperature, since the input thermal power will be
dissipated via convection and thermal radiation. If an I-V curve is recorded in 1 second
every minute, a part of the MPP power which is delivered to the load (20 % of the
normally incident solar power Pin) will remain in the cells. One needs to compare the
extra energy retained in the module in the form of heat during the curve sweeping to the
energy from carrier thermalization accumulated in 1 minute.
In 1 minute (the interval between consecutive I-V scans), the thermal energy
delivered to the cells is 60×0.7×Pin = 42Pin (in Joules, whereas Pin is in Watts). During
the 1-second sweep, module’s output power is varied between zero and PMAX = 0.2Pin
and then again to zero (see Fig. 3.4). Since the I-V curves are swept with constant time
and current increments, the horizontal axis in Fig. 3.4 is equivalent to time flow. The
extra energy remaining in the module as heat is approximately a half of the output
energy if the module was operated at MPP during the 1 second of the sweep, and equals
0.5×1×0.2×Pin = 0.1Pin (in Joules). This supplement is only about 0.2 % compared to
the main heat input from thermalization of carriers within 1 minute. This is too small to
cause any significant increase in the module’s operating temperature.

3.4 CHOOSING PV TECHNOLOGIES FOR TESTING


For greater cost-effectiveness of terrestrial PV systems, cell efficiencies of at
least 15–16 % are recommended [Que09]. This is apart from the requirement for a low
manufacturing cost. Besides module prices, there are area-related costs associated with
e.g. prices of land or roof space, hardware supports, installation services and cabling.
For lower-efficiency PV technologies such as some thin-films, there will generally be a
higher ‘area-penalty’. In addition, there can be a huge administrative cost (‘bureaucratic
penalty’) comparable with the cost of the entire PV system [Pho11b]. Insurance is an
example of operating cost which may be paid during the whole system lifetime. The
return from all these investments comes from the AC electricity sold to the utility
company (assuming that a FIT regulation is present), and the price varies between the
different states [Hir11]. The power fed into the grid by a PV system is less than the
28

instantaneous DC power of the installed modules, since there are power-conversion


losses in the inverters, as well as resistive losses in the cables. The annual energy yield
from PV is to a first approximation proportional to the yearly solar irradiation in the
plane of the modules [Hul08]. It is therefore advisable to use higher-efficiency PV
technologies in locations with relatively low solar resource, such as Southern Norway.
High efficiency becomes even more important in the case of non-optimal rooftop PV
arrays (neither facing south nor optimally inclined).
The total area of a PV module includes, in addition to cells, also the gaps
surrounding the cells and the framework (if any). Consequently, the module’s efficiency
is typically a few percentage points below the efficiency of its cells [Sun11]. Given the
minimum cell efficiency suggested by Queisser [Que09], the corresponding minimum
module efficiency is about 12–13 %. Commercially available terrestrial PV
technologies that achieve or exceed this level include various cell structures based on c-
Si, as well as highest-efficiency CIS/CIGS modules [Haa11]. Note that only non-
concentrating PV is being considered, assuming installations in Southern Norway. The
present status and recent history of the global PV market are reviewed in the following
part of the thesis.

3.4.1 CURRENT STATUS OF TERRESTRIAL PV


Solar cell production in 2010 was estimated to about 27 gigawatt-peak (GWP),
resulting in a year-to-year growth of 118 % [Her11]. The majority of the produced cells
were made of multicrystalline silicon (mc-Si). For year 2011, cell production in excess
of 50 GWP was expected. A market survey from year 2010 of PV modules suitable for
grid connection covered 3386 models announced by 165 vendors. The majority were
mono- and mc-Si modules employing cells with screen-printed front metallization. The
survey performed in 2011 covered 5252 models from 204 vendors [Haa11]. The
combined market share of monocrystalline silicon (mono-Si) and mc-Si has been over
90 % until 2007, when thin films surpassed the 10 % level. In 2009, thin films reached a
17 % share, with cadmium telluride (CdTe) alone taking 13 % [Min10]. However, in
2010 crystalline Si (c-Si) PV contributed much more to the enormous growth than the
thin films. The much lower share of the latter may be explained with their inferior
efficiencies compared to mainstream c-Si at otherwise comparable prices. The average
efficiency within the c-Si family was about 14 % as of 2011. The gap of about 1 %
absolute between mc-Si and mono-Si has remained relatively constant over the years
[Haa11]. Meanwhile, concentrating PV (CPV) with typical efficiencies of about 25 %
(far exceeding c-Si on the average) is still too expensive. CPV requires very accurate
sun-tracking devices and massive support structures, making them competitive only in
areas with annual solar irradiation exceeding 2500 kWh/m2 [Ses10]. For comparison,
the typical solar resource in Italy (presently the second largest market for PV) is about
1800 kWh/m2. Significant cost reductions may eventually make CPV more competitive.
29

3.4.2 THE No. 1 ‒ CRYSTALLINE Si


Until 1997, c-Si solar cells were made using waste Si from the microelectronics
industry. The PV makers eventually started buying expensive, electronics-grade mono-
Si wafers [Goe03, Bra08]. Casting of mc-Si by directional solidification reduced the
cost and resulted in somewhat less efficient PV cells and modules. The energy-costly
Siemens chemical process provided an ultrapure raw Si material (polysilicon) to both
industries. Shortages in its supply led to the development of alternative types of Si
feedstock (solar-grade silicon, SoG-Si) of less purity as compared to electronic-grade
silicon (EG-Si), but still sufficient for PV applications [Sin09]. An example is Elkem
Solar’s process starting with pyrometallurgical refinement of metallurgical-grade silicon
(MG-Si), followed by chemical treatment using acid solutions [Fri04]. Thus, the
required energy is 5-6 times smaller than in the classic Siemens process. Unlike
polysilicon that comes in the form of chunks of irregular size, the Elkem Solar Silicon
(ESS™) is a high-density feedstock (in the form of bricks) allowing for higher crucible
charging and growing of bigger mc-Si ingots [Chu11a]. Casting of quasi-
monocrystalline ingots (quasi-mono-Si) is a recent trend. It can be comparable in cost to
mc-Si while leading to mono-Si efficiencies for parts of the ingot [Chu11b]. However,
detailed I-V curve analyses of quasi-mono-Si modules are missing in the literature, and
could not be included in the present work. Other growth techniques result in c-Si of
much finer grain sizes leading to reduced cell efficiencies [Goe03]. Terms describing c-
Si based on grain size were proposed in [Bas94, Gre95]. Thus, sc-Si stands for ‘single-
crystal’ Si with grains larger than 10 cm; mc-Si covers grain sizes in the range 1 mm –
10 cm; pc-Si (polycrystalline Si) of grain sizes between 1 μm and 1 mm is grown by
chemical vapor deposition (CVD), whereas plasma deposition results in
microcrystalline material (μc-Si) with grains smaller than 1 μm. However, the present
author prefers using ‘mono-Si’ rather than ‘sc-Si’.

3.4.3 HIGH-EFFICIENCY c-Si PV


The high-latitude context of the present work has made the study of high-
efficiency PV a natural choice. Raising cell efficiency usually requires high-lifetime
substrates and comes at the price of more complex processing or completely different
structure. Several alternatives of the mainstream screen-printed cell design have been
developed and commercialized. These include emitter wrap through (EWT), buried
contacts (BC), heterojunction with intrinsic thin layer (HIT), and interdigitated back
contacts (IBC) cells [Gre95, Mas04, Neu07, Hez09]. The latter became the first non-
concentrating technology offering module efficiencies of 20 % [Sun11]. HIT modules
with 18 % efficiency are now commercially available [San11]. Production of modules
with BC cells had been discontinued by the time when the acquisition of devices for the
present work began [BPS10]. However, a field-aged BC module of a relatively low
efficiency (see next Section) had been available at UiA, and so this technology is also
30

analyzed in Chapters 6 and 7. Due to area and resource limitations, the only true high-
efficiency cell designs studied in the thesis are an IBC module and a HIT module.

3.4.4 MODULE LIST


The PV modules analyzed in the present work are listed in Table III.1. More
detailed information about their STC parameters ‘as new’ and the cell sizes is given in
[Ime11, Yor11a]. The mainstream screen-printed cell technology is represented by one
Chinese-made mono-Si module, three new and one aged mc-Si modules. One of the
new mc-Si generators is also Chinese-made, whereas the other two are made in Europe
and are based on Elkem Solar’s SoG-Si from their advanced metallurgical route. All the
new screen-printed modules have cells of size 15.6×15.6 cm2 with three busbars, except
the Chinese-made mc-Si module whose cells have only two busbars. The aged mc-Si
module has much smaller cells of size 10×10 cm2. It has been in the field between
summer, 2004 [Mid10] and autumn, 2008. Until the beginning of the present
measurement campaign in December, 2010, the module has been occasionally exposed
to sunlight indoors. The year of outdoor deployment of the aged BC module is
approximately 1997 [Sae12]. The author has not been able to acquire the model with
highest efficiency (12.1 %) listed in the AvanCIS PowerMax data sheet from year 2010.

TABLE III.1: Basic information about the PV modules studied in the present test setup.
Brand and model Type of the Cells Area [m2] Efficiency
RS Solar ZD300-24M Screen-printed mono-Si 1.94 15.5 %
RS Solar ZD280-24P Screen-printed mc-Si 1.94 14.4 %
BP Solar 585F* Buried-contact mono-Si (BC) 0.633 13.4 %
Gällivare PV GPV51* Screen-printed mc-Si 0.436 11.7 %
Q-Cells (ESS™)** Screen-printed mc-Si 1.71 12.6 %
Q-Cells (ESS™)** Screen-printed mc-Si 1.71 12.7 %
Sanyo HIT-240HDE4 a-Si/mono-Si (HIT) 1.39 17.3 %
Sunconnex SPR-95 Back-contact mono-Si (IBC) 0.547 17.4 %
AvanCIS PowerMax 110 Thin-film CIGS 1.09 10.1 %
*Modules are field aged. The given STC efficiencies are ‘as new’.
**Custom-built modules with 60 series-connected cells and 5 bypass diodes.

3.5 NEW VS. OLD EXPERIMENTAL SETUP IN GRIMSTAD: IMPROVEMENTS


The earlier study by Midtgård et al. [Mid10] measured and analyzed the field
operation of different types of PV modules. I-V curves of three modules were recorded
at various conditions. Parameters such as PMAX, VOC, η and the fill factor FF were
derived from the curves. The observed irradiance and temperature dependencies
resulted in annual efficiencies below those at STC, in line with the theory in [Hul08,
Hul10]. Furthermore, the FF analysis was helpful for explaining some outliers in the
efficiency plots with partial shadowing. The specimens were installed by a cottage wall,
facing south (see Fig. 3.5). On summer days, the sun in Grimstad was high enough over
31

many hours, and thus the roof shadowed the modules. In winter, surrounding objects
such as trees casted longer shadows due to much lower solar elevation above the
horizon (see Fig. 3.6). The setup suggested a somewhat higher optimal tilt angle of the
modules than that for a roof installation at the same latitude. At high latitudes, the
fraction of cloudy to clear-sky conditions is relatively high, and hence the diffuse
radiation gains in importance for the final yield and for the optimal tilt [Sur07, Hul10].
However, the cottage blocked the northern half of the sky hemisphere, and therefore a
higher tilt angle had to be used.

Figure 3.5: The old experimental setup in Grimstad used in [Mid10].

The drawbacks of the previous setup included also the added series resistance
due to the cables connecting each module with its electronic load. Its effect on the
measured power accuracy was considered to have been negligible, as the cables were
not too long. In addition, the modules were operated at open circuit in between the I-V
curve sweeps, because a MPPT algorithm was not implemented in the data acquisition
software. Low-cost, ‘home-made’ electronic loads with very poor power dissipation
capabilities were used in that setup. Since no electric power was extracted from the
modules most of the time, their temperatures must have been higher than in the case if
they were operated at the MPP between the I-V curve sweeps. Due to the negative
temperature coefficients of a PV device’s voltages and maximum power [Kin97a], the
obtained results about modules’ efficiencies were somewhat pessimistic.
The thesis is a logical continuation of the earlier research on PV at UiA. A main
goal of the present work is to build significantly improved experimental setups for
individual outdoor testing of many more modules and monitoring of key environmental
parameters (see Figs. 3.7 and 3.8). More complex data acquisition systems were
required in terms of both hardware and software. Commercially available multichannel
32

Figure 3.6: Graphs of the solar elevation in December through June for Grimstad,
Norway [UOr11]. Standard Central European Time is used for all months to ease
the comparison. Used with permission.
33

electronic loads are used. The new computer workstations have faster data acquisition
boards and mirrored storage (amounting to 2 terabytes in the setup located in Grimstad).
New LabVIEW-based code has been developed with a greatly increased functionality
and many new features added. The recorded I-V curves contain a few thousand data
points each (see Fig. 3.9). This allows efficient post-processing and noise cancellation,
reveals bypass-diode effects in much greater detail, and facilitates the analysis and the
application of differential techniques. The latter revealed an unexpected local rise of the
ideality factor in the open-circuit region for all tested modules [Yor12a].

Figure 3.7: Configuration of the present PV test installation in Grimstad.

Between the I-V curve sweeps, the modules are operated in MPP-tracking
mode, which results in realistic operating temperatures. The new experimental setup in
Grimstad and its diverse functionalities were recently reported in [Ime11], and are
described also in the present Chapter. Astronomical algorithms for calculation of solar
position and daylength are implemented [Mic88, For95], and the daylight-saving time
changes in spring and in autumn are accounted for [EUR01]. This allows calculation of
the angle of incidence (AOI) of the direct solar irradiance component on the modules.
In addition, the clear-sky air mass (AM) is calculated using an approximated equation
[Kas89, Luq03b]. Ambient and module temperatures are recorded also at night, which
allows accuracy monitoring of the sensors and transmitters. Each computer is remotely
accessible via the internet, which makes the monitoring, data downloads and software
updates available to the whole research group even from abroad.
34

Figure 3.8: Interface of the LabVIEW-based data acquisition software.


35

Figure 3.9: I-V curves of ten PV modules obtained with the new experimental
setup in Grimstad. The curves of the three aged modules are plotted with stars.

3.6 TEST SITE AND MODULE CONFIGURATION


The tested PV modules are installed in an ‘open-rack’ configuration on a flat
roof (see Fig. 3.7 and the bottom image in Fig. 3.10). The test site is located in the
coastal city of Grimstad, at latitude 58°20’N and altitude about 60 m a.s.l. The
azimuthal orientation of the modules is almost due South (deviating by about 7° towards
East). This sub-optimal orientation is due to the modules’ alignment with the building.
It has been estimated from satellite images available through the software Google Earth
[Goo12], with the help of the built-in ‘Ruler’ tool. The latitude and the approximate
altitude of the test site have been estimated using the same software. According to the
online PVGIS tool introduced in Chapter 2, the deviation of the modules from South
practically does not result in a reduced annual energy output. This can be expected,
because in the months with strongest irradiation, the sun is available at a wide azimuthal
range that exceeds the 180° field of view of the modules (see Figs. 2.3 and 3.6).
However, this consideration does not include the possibility for a greatly asymmetrical
distribution in time of cloudy conditions with respect to the solar noon (e.g. systematic
occurrence of rainy clouds in the afternoons). This is investigated in Chapter 4.
The tilt angle of a PV array is defined with respect to horizontal position. The
optimal tilt angle of fixed tilted PV plants is determined by complex models involving
36

irradiation and cost optimization. The area-related costs are a driver for placing rows
with modules close to each other. Thus, the tilt angle impacts the degree of mutual
shadowing between adjacent rows, with larger angles resulting in longer shadows (at a
given solar elevation). Eventually, a trade-off must be made between lower area demand
and yield-optimized tilt angle [Bre10a, Wei09]. In the present work, the modules are
installed in a single row, with almost no shadowing from surrounding structures.
Therefore, the tilt angle could be optimized solely for maximizing the annual solar
irradiation in the plane of the modules.

Figure 3.10: Configuration of the outdoor PV test platform in Grimstad.


37

Figure 3.11: Sunrise on 8 January, 2011. The tower was not causing shadowing.

Figure 3.12: Effect of tower shadowing on irradiance measured on 5 March, 2011.


38

Figure 3.13: Effect of tower shadowing on modules’ powers on 13 October, 2011.

The ‘rule of thumb’ that the irradiation-optimized tilt angle equals the site
latitude (‘latitude tilt’) does not apply for latitudes higher than 45°N [Bre10a]. For such
locations, the more frequent occurrence of cloudy weather results in a large fraction of
diffuse irradiation. Consequently, the tilt of a PV array with respect to the horizontal
must be substantially reduced compared to the latitude tilt, so that the modules can ‘see’
a larger portion of the sky hemisphere [Lub11]. For northern Europe, the deviation from
latitude tilt can exceed 10°, whereas for parts of the Scandinavian Peninsula (including
the present test site) it exceeds 15° [Bre10a]. The tilt angle of the modules tested in the
present work is 39°1°. The individual tilt of each module has been determined using a
smartphone equipped with accelerometer and bubble level software [HTC12, Bub12].
According to the online PVGIS tool, these angles are optimal for maximizing the annual
energy output from c-Si and CIS/CIGS modules at the given site and with the building’s
azimuthal orientation. The variation in the tilt angle between the modules is small
enough to affect their annual yield significantly [PVGIS, Mic88].
The main source of shadowing for the modules is a communication tower
located next to the campus building, roughly 65° South-East with respect to the research
platform (see Fig. 3.10). The fence surrounding the platform is lower than the tested
modules, which are in the upper part of the PV array shown in Fig. 3.7. In the summer,
39

the solar elevation is too high and the tower’s shadow is too short to affect the panels.
No shadowing occurs in the winter, since the tower is outside the range of possible solar
azimuthal angles (see Fig. 3.11). The tower causes only short-term shadowing on sunny
mornings in late February, early March and in October. This affects the measured
irradiance and the modules’ maximum powers (see Figs. 3.12 and 3.13).

3.7 INSPECTION, OBSERVATIONS AND MAINTENANCE

Figure 3.14, Top: Snow cover on the PV modules; Bottom: Wet roof reflections.
40

The operating conditions of the modules have been inspected regularly during
year 2011. In order to capture shadowing, birds, frost and snow cover, a networked TV
camera has been monitoring most of the installed modules. It takes a snapshot every
minute during daytime, but synchronization with I-V curve recording has not been
implemented. All images are stored on the computer running the LabVIEW-based data
acquisition system, and are available for later reviewing and analysis. Online video
monitoring is also possible, by means of an internet browser. In addition, on-site
inspection of the research platform has been done on a regular basis, and specific
situations have been photographed manually (see Fig. 3.14). These include drastically
increased albedo, frost or snow cover, bird droppings and soiling on the modules.
The effect of snow in the winter months has been minimized as much as possible
by regular cleaning. In many cases, perfect cleaning has not been possible, whereas frost
cleaning has not been attempted at all. Any ice remaining on the modules has eventually
melted due to solar heating. Reflections from water or ice accumulated on the roof
could not be eliminated at all. Their effects are analyzed in Chapter 4. Soiling and bird
droppings have been removed using a mop and hot water.
The research platform is elevated by about 15 m above ground level. In addition,
dust is practically absent at the test location. Soiling on the modules has originated
mostly from plant pollen brought by wind or rain. Bird droppings have been a particular
problem in the spring, as multiple seagulls have colonized (nested) the building’s roof.
In that period, much more frequent cleaning of the modules has been necessary, while
minimizing artificial shadowing as much as possible. Detailed quantification of the
energy loss caused by birds is outside the scope of this thesis. Their droppings were
shown to cause minimal yet notable effects on modules’ I-V curves and output powers
[Ime11]. The irradiance sensors have remained unaffected by this natural phenomenon.

3.8 DATA ACQUISITION HARDWARE


A commercial multichannel electronic load system, TDI Dynaload MCL488, is
used for regular I-V curves recording (see Fig. 3.15). It is controlled by custom-made
LabVIEW code via the PC-based data-acquisition boards NI PCIe-6363 and NI PCIe-
7851R. Each module’s current and voltage are monitored locally at dedicated outputs on
the electronic load. I-V curves are obtained by varying the resistance of each load. Only
backward I-V sweeps (starting from open circuit) are possible. The on-state resistance
of the loads is about 0.15  for those rated 20 A and about 0.3  for the ones rated 10
A. This prevents the PV modules from complete shorting (see Fig. 3.9). Extrapolation is
therefore needed for estimating the short-circuit current, ISC. According to the
calibration protocols and certificates, the loads and the chassis were last calibrated in
September 2010. The stated accuracy of each load’s current sample output is 0.5 % of
full scale, whereas the voltage sample output is connected directly to the power input
terminals. The stated accuracy of voltage measurement with the NI boards is very high
(38 μV for the used maximum input range of ±10 V). The open-circuit voltages of all
41

modules are much higher than the broadest input range of the NI boards. Therefore,
modules’ voltages are sampled indirectly via voltage dividers, each composed of two
resistors of 1 % precision in series connection totaling between 20.49 and 59.6 kΩ
(measured with a digital multimeter for better accuracy). These values are high enough
to ensure negligible power dissipation, and therefore practically no heating (and
temperature variation) of the elements occurs. In addition, the equipment room is air-
conditioned and the indoor temperature is maintained in a narrow range throughout the
year. The voltage-division ratios are thus determined with precision better than 0.1 %.

Figure 3.15: The main indoor data acquisition hardware.

Each module is linked to its load by means of two 6-mm2 copper solar cables.
The length of the leads is in the order of 10-15 m (one way). Some extra resistance
(between 0.06 and 0.11 Ω) is thus being added to the module’s intrinsic series
resistance. The results for modules’ resistances presented in Chapter 6 exclude these
additional resistances. Also, the measured maximum powers have been corrected by
2
adding the resistive loss I M R in the cables prior to the power rating described in
Chapter 6, and the calculation of the annual energy yields and the relative efficiency
analysis in Chapter 7. A change in RS causes mainly a voltage shift in the I-V curve,
and not a current shift. For small enough variations in RS, the current level IM at the
MPP remains practically unchanged [Pys07]. A detailed analytical proof is not given in
the thesis, but the associated variations in IM have been investigated for each module by
analyzing I-V curves recorded at nearly standard irradiance. At standard illumination
level, IM has been found to rise by 0.1 to 0.5% for the different modules after correcting
42

the I-V curve for the voltage drop –IMR. The corresponding uncertainty of I M
2
R is
within 1 %, which is negligible compared to the module’s rated power.
The resistance R of each pair of interconnecting cables has been determined by
shorting the connectors on the module’s side (after disconnecting the module) and then
running DC current through the loop by means of a regulated power supply. The voltage
drop across the cable loop was measured for four current levels in the range 1-4 A, and
then the resistance was estimated by linear regression with a precision better than 1 %.
The experiment was carried out after sunset, at an ambient temperature of about 10°C. It
should be noted that the part of the cables exposed to direct sunlight is expected to
become much hotter than the ambient air because of their black insulation. The
resistivity of copper has a positive temperature coefficient of about 0.43 %/°C in the
range between –40°C and +40°C [Dau54]. However, it is difficult to assess the variation
of cable resistances due to heating on sunny days. A substantial but unknown part of
each cable’s length is not exposed to direct sunlight, since it is either placed indoors or
being shadowed by the PV modules. In addition, it is impractical to measure the
temperature of the wires where the cables are heated by sunlight. In general, some
increase of the cables’ resistances is expected due to natural heating and changes in the
ambient temperature. This topic is not investigated in detail in the present work, but is
considered in the context of the estimated series resistances of the modules later in
Chapter 6.
In order to reduce the noise due to electromagnetic interference (EMI) in the
measured signals, these are conducted via Ethernet network cables with shielded-and-
foiled twisted pairs (SFTP) of isolated copper wires. All signal paths have been made as
short as possible. In addition, the National Instruments (NI) data acquisition boards are
equipped with shielded connector blocks and shielded signal cables. The high
availability of the measurement system is achieved by powering it with a double-
conversion (‘online’) uninterruptible power supply (UPS).
The hardware measuring solar irradiance, ambient and modules’ temperatures is
reviewed later in Chapters 4 and 5. Wind speed and direction are not measured in the
present experimental setup. These parameters are an important part of the operating
conditions of any PV system, because wind-assisted cooling of the PV modules is
beneficial for the annual energy yield. However, thermal models of PV modules
involving these parameters are not analyzed in the thesis. In addition, the geometries of
the building and of the PV array suggest significant modifications and local non-
uniformity of the natural air flow due to turbulence, as well as due to changing wind
direction. One of the data acquisition boards has vacant analog inputs and this allows
for expansion of the experimental setup with wind sensors in future. The spectrum of
the incident sunlight is not recorded at present, but a spectrometer may be added on a
later stage of the research.
43

3.9 LABVIEW CODE; DATA TYPES, FORMATS AND PRE-PROCESSING


The data acquisition software has been developed using the NI LabVIEW™
programming environment. The project consists of 24 executable files which provide
various functionalities such as: generation of control waveforms for the electronic loads;
measurement of I-V curves and environmental parameters; data quality analysis; linear
fitting and smoothing of raw data; extraction of I-V curve parameters; data recording,
etc. In order to avoid slow program execution due to e.g. antiviral software interference,
the priority of the process ‘LabVIEW.exe’ is manually changed from ‘Normal’ to
‘High’ in Windows Task Manager after each restart of the PC. This setting is of
particular importance for ‘real-time’ recording of solar irradiance and modules’ output
powers during the MPPT phases between consecutive I-V curve scans. This Section
reviews the key features of the software that are relevant for the quality and accuracy of
the data used in the thesis.

3.9.1 I-V CURVES AND IRRADIANCE STABILITY

Figure 3.16: Irradiance in the plane of the PV modules during I-V curve sweeping.

I-V curves of all modules are swept regularly at roughly 1-minute intervals as
long as any of the two irradiance sensors reads at least 30 W/m2. The in-plane irradiance
is also recorded during each sweep, and is immediately checked for instability. A
44

measure used to characterize the degree of its temporal variation is the Irradiance
Stability Factor ISF = (GMAX–GMIN)/GMEAN [Jag09]. However, inspection of these
irradiance records has revealed the presence of noise in the data dominated by a 50-Hz
mode attributed to EMI from the utility grid (see Fig. 3.16). Since the maximum and the
minimum irradiance values (GMAX and GMIN, respectively) are very often altered by the
noise, ISF usually overestimates the actual instability. Therefore, the standard deviation
σG has been used instead as a quality indicator. In case σG exceeds 3 % of the mean
irradiance GMEAN, the recorded I-V curves are disregarded, and new curves are swept
immediately. This may be repeated up to seven times in order to avoid infinite loop
cycling at very low irradiances. For these, the signal-to-noise ratio (SNR) becomes too
low, and the above condition is not satisfied even for very stable illumination. Another
condition which may trigger repeated sweeping is measurement of an incomplete I-V
curve for any of the modules, which can happen occasionally. An I-V curve is
considered incomplete when either the lowest voltage exceeds 5 V or the lowest current
exceeds 30 mA. The first criterion is a trade-off between the minimal voltages imposed
by the electronic loads (as discussed in Section 3.8) and the typical voltage ranges of the
four modules optimized for 12-Volt applications (see Fig. 3.9). The second criterion is
dictated by the typical ranges of current generated by the modules denoted ‘CIS’, ‘GPV’
and ‘a-Si 3J’ at irradiances close to the lower limit of 30 W/m2.

Figure 3.17: I-V curve affected by the shadow of a bird flying over the PV module.

The raw I-V curves are stored in text files on the PC. The files recorded on
different days are stored in separate folders. Experience with the test setup in
Kristiansand has shown that storing all data files in the same folder results in gradually
degrading software performance when the number of files reaches tens of thousands.
45

This has been avoided in the setup located in Grimstad for the sake of adequate ‘real-
time’ recording of irradiance and output powers. Later inspection of individual curves
allows the tracing some outliers in the extracted parameters (such as a module’s series
resistance and ideality factor [Yor10b]) to abnormal I-V curves (see Fig. 3.17).

3.9.2 PRE-PROCESSING OF RAW I-V DATA

Figure 3.18: A discontinuity in an I-V curve at open circuit.

Key I-V curve parameters are stored in daily text files together with module
temperatures and environmental parameters. The short-circuit current ISC is evaluated
by linear fitting of seven data points from the lower-voltage part of the curve. This small
number is imposed by the very flat I-V curves of the HIT module which can have too
few data points in this region (see Fig. 3.9). However, the region below 4 V is excluded
from the fit because very often a cluster of arbitrary shape containing many points is
formed at the end of the curve, where the electronic load enters saturation mode. The
evaluation of the open-circuit voltage VOC is similarly complicated due to the electronic
load’s behavior when it starts to conduct current (see Fig. 3.18). The curve usually
begins with a cluster with tens of points at nearly zero current followed by a
discontinuity. Outliers are often present in this part of the curve. Therefore, linear fitting
46

based on the bi-square algorithm is used, which is less sensitive to outliers in the data.
The software fits the first 250 points, corresponding to currents up to 6 % of ISC. The
fragment is short enough and therefore the linear approximation is very accurate. For
example, the curve in Fig. 3.17 is fitted between 0 and 0.43 A.

Figure 3.19: Noisy P-V curve with an outlier becoming a false MPP.

The estimation of PMAX and particularly of VM has been most challenging due to
noise present in the raw I-V curves. (Note that IM=PMAX/VM is a dependent parameter.)
An outlier in a noisy P-V curve may become a false MPP (see Fig. 3.19), unless proper
measures are taken to smooth the data by local averaging or by curve fitting. Whatever
algorithm is chosen, it should work well also with curves recorded under partial
shadowing conditions. For example, the P-V curve in Fig. 3.20 cannot be fitted
accurately in LabVIEW with a general polynomial fit. However, what is needed for
estimating PMAX and VM is partial fitting in the vicinity of the MPP. Complications
associated with non-linear curve fitting include the requirements for providing a fitting
function, making initial guesses of its coefficients, and in some cases – ascending order
of the array containing the independent variable (the voltage V), which is not fulfilled in
the present setup. For several reasons, curve fitting has not been implemented in the
data acquisition software. Instead, smoothing by means of moving average has been
47

used. In Fig. 3.19, the false MPP is at VM=15.02 V. After smoothing, the LabVIEW
code estimated the optimal value VM=14.92 V.

Figure 3.20: P-V curve of a partially shadowed PV module.

3.9.3 ASTRONOMICAL AND OPTICAL FEATURES


Solar elevation and azimuth, as well as the air mass (AM) factor corresponding
to clear-sky conditions, are also calculated and recorded at the time of each I-V curve
sweeping. The astronomical algorithm used to calculate the position of the sun has a
stated accuracy of 0.01° until year 2050 and is particularly useful for high-concentration
PV systems, which require accurate sun tracking [Mic88]. The LabVIEW code uses the
computer date and time to calculate input parameters provided to the algorithm such as
Julian date, Greenwich mean sidereal time, hour angle, solar declination, right
ascension, day number. The daylight-saving time changes implemented on the last
Sundays of March and October are also accounted for [EUR01]. In the present version
of the software, the exact dates are entered manually for each year, but full automation
may be programmed in future. An approximated formula is used to determine the air
mass (defined at 700-nm wavelength) as a function of solar elevation [Kas89]. The
maximum error is less than 0.5 % and occurs at the horizon. A correction factor for the
altitude is applied as in [Kin97c]. The latter work gives a strongly underestimated
48

sunrise/sunset AM value of about 10, whereas values around 38 were obtained in


[Kas89] and several earlier works referenced therein (see Fig. 3.21). The solar position
is also used in the calculation of sunrise and sunset times, which determine the
daylength and thus also the times when the software switches between daytime and
nocturnal measurements. Daylength can be accurately determined with a maximum
daily error of 7 min at latitude 60°N [For95]. The adopted model is applicable for a flat
terrain at zero altitude – very suitable for the present test site located at 60 m a.s.l, with
the sea occupying large part of the horizon. The model assumes clear-sky conditions.

Figure 3.21: Approximated AM [Kas89] with correction for altitude [Kin97c].

There are different definitions of sunrise/sunset and twilight. The sunrise/sunset


definition preferred in the present work is when the upper rim of the sun is apparently
even with the horizon. Refraction of sunlight in the atmosphere amounts to about 34’
when the sun is near the horizon, and causes the sun to appear on the horizon when its
astronomical position is below it. The solar radius as seen from the earth is 16’.
Therefore, the center of the solar disk is about 0.83° below the horizon when its upper
rim is apparently even with the horizon. The civil twilight definition is preferred,
according to which the day begins when the center of the sun is 6° below the horizon
[For95]. This definition resulted in the most accurate estimate of global solar radiation
49

for Toledo, Spain [Alm05]. The rotation and the orbital revolution of the earth are
modeled (using elliptical and not simplified circular orbit equation). The model is
capable of predicting the periods of continuous light and continuous darkness for
locations beyond the Arctic Circle like Narvik in Norway, and could be useful to PV
research groups based there [Goo09, Kle09].

3.9.4 MPP TRACKING ALGORITHM

Figure 3.22: Maximum power values extracted from I-V curves (discrete symbols)
superimposed on 1-second averages of power from MPPT (continuous lines).

Maintaining realistic operating temperatures requires MPP operation of the


modules between the I-V scans [Jag09]. Therefore, a MPPT algorithm of reasonable
accuracy has been implemented in the present data acquisition software. During each
MPPT phase, the output powers of all modules are recorded with sub-second resolution
aiming at very accurate energy accounting. In order to save storage space, the measured
power data are stored in text files as time series of 1-second averages. Figure 3.22
shows 1-minute PMAX data (extracted from multiple I-V curves) superimposed on the
corresponding output powers recorded during the MPPT phases. These plots show that
50

the algorithm is quite accurate for small variations of irradiance between the I-V curve
sweeps. Each module is dynamically operated at a current I(t) equal to the last curve’s
IM scaled with the ratio of the real-time irradiance G(t) and GMEAN from the I-V sweep.
(Time is t.) This approach assumes the approximation IM=µISC where µ is constant
[Ser10a, Yor10b], and also that ISC varies linearly with G. Especially for the triple-
junction a-Si module, the latter assumption translates as a requirement for unchanged
spectrum of sunlight [Kin97c, Kin00], which may not be fulfilled during transitions
between clear-sky and overcast conditions and vice versa. The spectral sensitivity of the
nine single-junction devices tested in Grimstad is considered to be small (resulting in
several per cent deviation from linearity of ISC). Spectral effects on the tested modules
are discussed in detail in Chapter 4. The a-Si module is outside of scope for Chapters 5,
6 and 7, and therefore its MPPT imperfections are not significant for the present work.
The data acquisition software allows implementation of an individual algorithm for this
device at some future stage of the present experimental setup.

Figure 3.23: Ratio IM/ISC of the ten modules on a clear-sky day (21 June 2011).

Plots of the ratio IM/ISC on a clear-sky day for the modules tested in Grimstad
are given in Fig. 3.23. The corresponding irradiance profile appears in Fig. 4.7 further in
Chapter 4. For most of the modules, the approximation of a constant ratio of the two
51

currents is reasonable for irradiances down to 200 W/m2. The mc-Si module denoted
‘RS poly’ is a notable exception at the higher irradiances, whereas the aged mc-Si
module denoted ‘GPV’ deviates below 300-400 W/m2 depending on its temperature.
The behavior of some of the plots at very low irradiances is attributed to partial
shadowing from module frameworks at AOI close to 90°. However, somewhat greater
inaccuracy of the MPPT algorithm is considered acceptable at low irradiances, at which
sunlight does not heat the modules much above the ambient air temperature.

Figure 3.24: Maximum power values extracted from I-V curves (discrete symbols)
superimposed on 1-second averages of power from MPPT (continuous lines).

The MPPT algorithm may be compromised when irradiance changes drastically


within one minute. Such possibility has been investigated by visual inspection of power
plots and by comparing the daily energy yields from 1-minute I-V curve data to the ones
calculated from real-time data. The discrepancy is usually within 0.3 % on days with
high MPPT accuracy, but can be as high as 4.5 % for some modules on days with
frequent irradiance variations of high magnitudes. Fig. 3.24 contains an example of
MPPT malfunction for the modules denoted ‘HIT’ and ‘Sunc.’. The preceding I-V curve
sweep was done at rapidly rising irradiance and resulted in incomplete curves for most
of the modules. The two mentioned above were affected most severely. Although the
52

estimated PMAX and IM were correct, they corresponded to a higher irradiance level than
the average GMEAN. Consequently, the calculated operating current for the MPPT phase
exceeded the real-time IM, and therefore the operating point was systematically being
moved close to short circuit. The next I-V curve sweep was done at a more stable
irradiance, and therefore the MPPT algorithm in the following minute performed well
for all modules.
Occasional failure of the algorithm for some or for all modules means periodic
heating above the temperature in the case of a perfect MPPT, because some or most of
the available output power is not extracted. It is however questionable if the algorithms
in all commercial PV inverters can achieve perfect MPPT at rapidly changing
irradiance. From this point of view, the algorithm implemented in the present work is
considered satisfactory. In the context of energy accounting, the monthly and annual
energy yields of the modules obtained from the real-time data set would be slightly
underestimated, perhaps not more than the losses due to inaccurate MPPT in a practical
PV system. Also, somewhat underestimated yields (as compared to a perfect MPPT) can
be expected from the 1-minute data set due to elevated operating temperatures.
A rough estimate is possible of the maximum temperature increase resulting
from a MPPT failure like that of the module denoted ‘Sunc.’ in Fig. 3.24. This involves
the assumption of a zero power output, zero wind speed and steady-state conditions (in
terms of irradiance and thermal energy flows). The initial module temperature is
[Luq03a, Kra09, Hul10, Mid10]:

T  TA  cTG (3.1)

where TA is the ambient air temperature, G is the in-plane irradiance, and cT is a


constant dependent on the module mounting and ventilation. The module’s stabilized
temperature without MPPT (at short circuit) is obtained approximately by upscaling G
in Eq. (3.1) with module’s maximum available power output (which now remains in and
heats the device) divided by the module area. The approximate temperature increase is

PMAX
T  c T  cTG (3.2)
A

where η is module’s efficiency at the MPP. Assuming a conservative (larger) value


cT=0.035°C(W/m2)-1 for glass-plastic c-Si modules in open-rack configuration [Hul10],
standard irradiance of 1000 W/m2 and 20 % efficiency, one obtains ΔT = 7°C. For a
typical temperature coefficient of the maximum power of –0.5 %/°C [Got10], such
increase in temperature translates to a 3.5 % decrease in PMAX when MPPT is restored.
This rough estimate is a pessimistic one, as the assumed conditions are hardly fulfilled
simultaneously. The MPPT failure of the module denoted ‘Sunc.’ in Fig. 3.24 lasted
only 1 minute, whereas the algorithm seems to have worked fine before and after that.
53

The increase in module temperature during that minute is determined by device’s


transient thermal response [Moh03a]. The thermodynamics of the transient is described
by the linear differential Eq. (3.3) derived from an equivalent electronic circuit, see Fig.
3.25 [Moh03a]. The current source is the electrical analogue of the heat input into the
module due to thermalization and recombination of light-generated excess carriers; the
capacitor represents module’s thermal capacitance; and the resistor is the analogue of
module’s instantaneous thermal resistance. In this representation, the voltage across the
capacitor and the resistor is analogous to the temperature difference T–TA.

Figure 3.25: Lumped-circuit analogue to PV module thermodynamics [Moh03a].

dT P  t  TA T
   (3.3)
dt C RC RC

At steady-state condition with accurate MPPT and zero wind speed, dT/dt = 0
and P(t) = (1–r–η)GA. In the latter expression, r = 0.1 can be used as a realistic value
for the reflectance at the module’s frontal interfaces [Mar01]. The solution of the
homogeneous differential equation has exactly the form of Eq. (3.1):

T  TA  1  r    RAG  TA  cTG (3.4)

which gives R = cT/[A(1–r–η)] ≈ 1.43cT/A. Module short-circuiting due to MPPT fault


would result in a step-wise increase of P(t) by ΔP = ηGA to P(t) = (1–r)GA. The time
constant of the circuit, τ = RC ≈ 1.43cTC/A, is of order 103 s for the considered module
type [Kra09]. Thus, the thermal capacitance per unit area C/A is of order 2×104
J/(°Cm2). The solution of Eq. (3.3) for the thermal transient beginning at t = 0 is
54

  t    t 
T  t   TA  RAG 1  r   exp      TA  1.43cTG 0.9  0.2 exp     (3.5)
       

At t = 0, T equals the right-hand side of Eq. (3.4). At t = ∞ (assuming permanent MPPT


failure resulting in short circuit), T reaches the asymptotic value TA + 1.287cTG. The
temperature increase would thus be ΔT = 0.287cTG which equals 10.0°C at standard
irradiance. This exceeds by 3°C the earlier rough evaluation based on Eq. (3.2). The
difference comes from the greatly simplified theoretical treatment used back then, in
which reflectance loss was not accounted for. The radiative (luminescence) losses in c-
Si PV are negligible [Ker03]. At the end of the 1-minute MPPT fault analyzed here, the
module temperature rises by roughly 0.027cTG (assuming τ ≈ 10 min), which at
standard irradiance equals about 1°C. The effect on the next I-V curve is about 0.5 %
decrease in PMAX. The above calculations assumed rather high module efficiency
(possible on very cold winter days for the high-efficiency modules ‘Sunc.’ and ‘HIT’).
For a 15 % efficient module, the temperature increase after a 1-minute operation at short
circuit would be about 0.7°C. A larger τ = RC would result in even lower temperature
increase. Thus, for the double-glass CIGS module (with STC efficiency of about 10 %),
the effect from a 1-minute MPPT fault would be negligible due to the larger C and R.
On the other hand, both R and C are very low in the a-Si module, because it has
structured plastic instead of a front glass. Consequently, the thermal time constant of
such devices is an order of magnitude lower than for generic c-Si modules [Kra09]. Its
stabilized STC efficiency is only about 6 %. This module is outside the scope of the
thesis, and the analysis of its thermal properties is left to the interested readers.
A complete, detailed quantification of the effects of imperfect MPPT on the
operating temperatures and on the real-time power output of the tested modules is
outside the scope of the thesis. Total energy losses in the order of a few per cent
(equivalent to annual MPPT efficiencies of about 97 % or more) can be inferred by
comparing the energy yields estimated from the 1-minute and the real-time data set
(which is done later in Chapter 7). The corresponding loss is intermediate between zero
(assuming a perfect MPPT on days with slowly changing irradiance) and several per
cent obtained for some days with frequent, high-magnitude irradiance variations. The
complete data sets from 2011 may be analyzed in future work in order to find a more
accurate estimate of the present algorithm’s efficiency. An important result from the
above thermal modeling is that the high-efficiency c-Si modules could be hotter by
about 10°C at standard irradiance of 1000 W/m2 and zero wind speed, if they were kept
either at short circuit or at open circuit between the I-V curve sweeps. Thus, the
importance of MPPT in outdoor performance testing of PV modules is emphasized.
55

3.10 DAILY VISUAL DATA INSPECTION


The 1-minute data set has been analyzed on a daily basis (with few exceptions)
by means of a Matlab™ script that imports and plots the various recorded parameters.
Regular monitoring of the data has allowed the timely discovery of any hardware
malfunction, unusual operating conditions, and software errors. For example, in early
April 2011, the thermocouple transmitter of the aged mc-Si module ‘GPV’ broke down
due to condensation of water inside the junction box. It was later replaced, and the box
with a desiccant bag inside was sealed with silicone. On 11 April, the thermocouple at
the back of the a-Si module detached and its readings almost equalized with these of the
ambient temperature sensor (see Fig. 3.26). Fig. 4.21 in Chapter 4 is an example of
unwanted interference in irradiance measurements from equipment belonging to another
experiment conducted on the same research platform. The daily irradiance profiles have
revealed the surprisingly high magnitudes of some bursts caused by cloud enhancement
(see Fig. 4.7). Fill factors monitoring pointed at the unusually low FF of the Chinese-
made mc-Si module at higher irradiances (see Fig. 6.1 later in Chapter 6).

Figure 3.26: The record of ambient air and module temperatures from 11/04/2011.

In spring 2011 it became clear that the output of the LabVIEW solar azimuth
sub-program was limited within ±90°. This was caused by LabVIEW’s built-in inverse
56

sine function having a default output range of ±π/2 radians (±90°). The solar azimuth is
related to the solar declination and to the hour angle through a trigonometric equation
[Mic88]. The azimuth usually ranges between 0° (corresponding to North) and 360°,
whereas 90° is East. In the present work, a modified trigonometric equation is used, so
that 0° corresponds to South, –90° is East and +90° is West. One must therefore assign
the correct quadrant by modifying the output of the inverse sine function if necessary.
The solar elevation and the air mass (AM) calculated in LabVIEW were not affected.
To solve the issue, new code was added to the Matlab™ data visualization script, which
corrected the wrong azimuthal data and the resulting angle of incidence (see Fig. 3.27).

Figure 3.27: Solar azimuth and angle of incidence (AOI) of direct sunlight on the
modules calculated in LabVIEW and after the correction in Matlab™.

A couple of errors in the implementation of the astronomical algorithm in


LabVIEW were discovered in autumn, 2011. A ‘bug’ in the program ‘Day_of_Year.vi’
caused some delay in the data acquisition onset in the mornings of 1 and 2 October (see
Fig. 3.28). Also, the calculated solar position and air mass were incorrect until the error
was fixed at 11:15 a.m. on 2 October. Consequently, the AOI computed by the
Matlab™ script were affected for little more than one day (see Fig. 3.29). A more
general programming error related to the solar position was discovered later in October,
57

2011. The solar elevation at sunrise was underestimated by 3° or more for this particular
month. Sunrise was observed in the morning of 14 October when the corresponding part
of the horizon was clear of clouds. When the lower rim of the sun was more than one
radius (about 0.25° [Mic88]) above the sea surface, the LabVIEW GUI indicated a solar
elevation of –2.7°. This programming error was corrected on 24 October. It has affected

Figure 3.28: Late onset of data acquisition on 1 Oct, 2011 (a mostly clear-sky day)
due to a programming error in one of the astronomical functions in LabVIEW.

Figure 3.29: Erroneous solar position and air mass calculated for 2 Oct, 2011.
58

all solar position data recorded in LabVIEW since 5 February (that is, during most of
2011). In order to investigate the magnitude of the errors in the calculated parameters
(and to obtain correct estimates for that period as well), all the astronomical equations
were coded also in Matlab™ (see Figs. 3.30 and 3.31). For the interested readers, the
script is given in Appendix B at the end of the thesis.

Figure 3.30: Wrong vs. true solar position, air mass and AOI for 14 October, 2011.

Figure 3.31: Solar position, air mass and AOI for 21 December, 2011 calculated in
LabVIEW (large symbols) and Matlab (small symbols).
59

Plots of the short-circuit currents over irradiance have been useful in detecting
albedo effects (reflections from water or ice on the roof beneath the modules, see Fig.
3.14, the bottom image). Because the irradiance sensors and the modules are usually
affected differently, characteristic deviations from linearity are seen in these plots (see
Fig. 3.32).

Figure 3.32: Strong deviation of ISC data from linearity hints at albedo effects.

3.11 POST-PROCESSING OF RAW I-V CURVES


The raw I-V curves contain certain amount of noise (see Fig. 3.18) and therefore
they are not very suitable for analysis, especially with differential techniques. Proper
smoothing and omission of any redundant data points must be applied prior to
parameter estimation and discrete differentiation. Such methods have been used to
obtain the results presented later in Chapter 6.
Each I-V curve is first subjected to ‘smoothing spline’ interpolation available
from Matlab’s curve fitting toolbox [Mat12]. The choice of a proper smoothing
parameter p is very important, because small variations in its value result in drastically
different fits. For I-V data, the program defaults to p=1 resulting in cubic spline
interpolation, which is not smooth at all (see Fig. 3.33). On the other hand, too low a p
makes the fit deviate too much from the true shape of the I-V curve. This is illustrated in
60

Fig. 3.33, where the smoothed data obtained with p=0.9 depart notably from the raw
data at open circuit, whereas p=0.99 gives a more accurate interpolation.
From Fig. 3.33 it is also seen that the raw I-V curves contain many redundant
data points (that is, points at the same voltage but of different current). Thus, the
smoothed curves contain overlapping points with the same voltage and current. In
addition, the arrays containing voltage and current data are not monotonic (that is, the
voltage values are not arranged in descending order and the current values are not
arranged in ascending order). These artifacts arising from data acquisition and post-
processing are problematic when it comes to discrete differentiation. In order to obtain
monotonic I-V curves, the smoothed data are cleared from the redundant data points,
and then the voltage and current arrays are re-arranged in monotonic order. In this way,
the effects of noise on the extracted curve parameters, as well as on the m-V and m-I
plots, are greatly reduced.

Figure 3.33: Fitting raw I-V data with Matlab’s curve fitting toolbox using
different values for the smoothing spline parameter p.
61

CHAPTER 4 IRRADIANCE
4.1 CHAPTER AIMS AND OBJECTIVES
The solar resource is the most important contributor to a PV system’s energy
yield (see Fig. 2.2), and therefore its detailed characterization is a major goal of the
present work. Its evaluation is the main contributor to the uncertainty in long-term PV
yield predictions [The10]. This Chapter aims at obtaining an accurate estimate of the
plane-of-modules irradiance at the installation in Grimstad over the course of one full
year, and at analysis of the obtained 1-minute and 10-ms data sets. Another task is to
study several years of global horizontal irradiance (GHI) data from the nearest
meteorological station, which is located a few km away in Landvik. In both cases, the
annual irradiation and its seasonal distribution (i.e., the monthly irradiations) are of
particular interest. The maximum year-to-year variability of the annual irradiation in
Landvik is compared to typical values from the literature, and the averaged seasonal
distribution is compared to the long-term prediction from PVGIS.
A methodology for recovery of corrupted data is proposed, which is based on
self-referenced irradiance estimates of new PV modules. The annual tilted irradiation in
Grimstad in year 2011 and its seasonal distribution are compared to those reported in
[Mid10] for year 2005. Since that study used a very coarse temporal resolution of 20
minutes, the associated uncertainty is evaluated and compared to that due to 1-minute
irradiance sampling in the present work. The sub-second data from the new
experimental setup are analyzed with a focus on the magnitudes, durations and energy
contribution of the cloud enhancement events (overirradiances). Monthly and annual
distributions of the measured irradiance values are presented and analyzed. The optimal
azimuthal orientation at the present test site, as well as the amount of irradiation lost due
to clouds is evaluated from the 1-minute data set.

4.2 SUNLIGHT AT EARTH’S SURFACE


Sunlight is the thermal radiation emitted from a thin layer at the surface of our
nearest star which is heated to about 5500°C [Luq03b, Lea05]. The photon wavelengths
range between 5 nm and 10 μm, but the photons with very high energies are completely
absorbed by earth’s atmosphere and do not reach its surface. Most of the energy flux
lies in the range between 0.2 and 4 μm [NRE12a]. At sea level, the ‘hardest’ ultraviolet
(UV) photons have wavelengths of about 0.3 μm [Lea05]. Outside earth’s atmosphere,
the irradiance varies with the sun-earth distance by about 3.4 % around the value of the
solar constant discussed later in this Chapter. Scattering, absorption and ‘reflection’
(backscattering) in the atmosphere attenuate the extraterrestrial sunlight, and this
attenuation is in general wavelength-dependent [Nan91, Kin97c, Gue02, Per07, Zeh10].
At clear-sky conditions, the light reaching earth’s surface has two main
components. The direct (beam) component is emitted from a small, circular portion of
the sky within about 5-6° around the solar disk [Kin97c, Mye02, Eme03]. The diffuse
62

component comes from the rest of the sky [Liu60]. The clear-sky irradiance is
dominated by the beam component [Liu60, Mye02]. The intensity and the spectrum of
the latter depend on the atmospheric composition and properties (e.g. water vapor and
aerosol content, pressure and temperature [Sma85]) and on the optical path length of the
sunrays through the atmosphere. The latter is often described by the dimensionless
quantity (relative optical) air mass (AM). It is the ratio of the absolute optical air mass
at a given solar elevation above the horizon to the absolute air mass at sea level when
the sun is directly overhead (i.e. for solar elevation of 90°) [Kas89, Kin97c].
At totally overcast conditions, only diffuse light reaches ground and its intensity
depends on the thickness of the cloud cover. In the present work, no data were recorded
in Grimstad on eight days in 2011, during which the irradiance in the plane of the PV
modules did not exceed the chosen threshold of 30 W/m2. At partially overcast
conditions, the clouds can boost the irradiance on an optimally inclined plane well
above the clear-sky level and even above the extraterrestrial values, which is discussed
later in Sections 4.5 and 4.8. Inclined surfaces are also susceptible to ground-reflected
light, which is added to the two main components discussed above.
The sky radiance is generally anisotropic and its distribution strongly depends
on the exact conditions, which can be described by various indices [Ska86, Per93,
Iga04, Hul08]. In addition to the often used clear sky index, another (correlated) index
is usually needed for comprehensive description of the sky conditions. The anisotropy
greatly complicates the calculation of irradiance on a slope, as well as the modeling of
angular effects on PV modules and sensors [Var00, Mar01, Abe03]. Most PV
performance models perform well for clear-sky conditions, but have difficulties for
cloudy conditions, for which the light spectrum is not understood well [Got10, Hul10].
The direct sunlight is redder at sunrise and sunset (i.e., for low solar elevation
and high air mass), and bluer at noon [Kin97c]. The spectrum of the diffuse light is
dominated by shorter-wavelength photons and is almost independent on air mass.
Spectral variation introduces a systematic influence on PV device performance, which is
time-of-day dependent. Empirically derived correction factors for spectral effects can be
used in the modeling of PV device performance [Kin97c, Abe03, Bet04, Per04, Wil04a,
Per07, Tsu08]. For clear sky conditions, the air mass may be sufficient for describing
the spectral dependencies of PV [Kin96, Kin00].
Clouds can cause significant temporal and spatial variability of solar irradiance.
In large PV systems, the big size can smooth the fluctuations in the energy input, and
therefore hourly irradiance data may be sufficient for predicting their electricity
production. However, very high temporal resolution of irradiance measurements (1-
minute or even 1-second) is needed to characterize the operating conditions of smaller-
area PV systems [Ham12]. This is particularly important at partially overcast conditions
with frequent overirradiance events, whereas the hourly averages smooth out the
extreme values. A recent trend in the PV power field is the simulation of a system’s
output based on irradiance data with time steps of 1 minute or less [Gue09].
63

Microclimates can have a significant impact on the solar resource in some


locations. For example, several microclimates take place in the San Francisco area,
which are expressed as distinct differences in weather across short distances [Mit10].
The maximum difference in the annual irradiation was estimated to be almost 20 %. The
authors of the study concluded that having site-specific solar resource data was critical
for assessing the performance of a PV system. In addition, coastal regions (especially
those with complex sea-land patterns) seem to have a greater year-to-year variability of
the annual solar irradiation [The10]. The present experimental setup is located in such a
region, among the entry of the Grimstad fjord to the east, which is surrounded by
multiple islets, and several fresh-water lakes to the west and north-west [Map12].

4.3 MEASUREMENT METHODS AND SENSORS; DATA SOURCES


Measurement of global horizontal irradiance (GHI) is an established practice of
most meteorological stations [Ols86, Hul08]. The data available to the solar community
usually have a very low temporal resolution (e.g. daily global irradiation). Satellite
images have been used to obtain 30-minute and hourly irradiance values [Esp09,
Mye09b, Hul10, Sat12]. In order to determine the performance of a PV system with
particular configuration (fixed or with sun tracking), models are used that estimate the
instantaneous direct, diffuse and ground-reflected irradiance on a tilted plane [Gue09].
Using averaged irradiance values in performance models is not recommended, because
a PV module’s efficiency has a non-linear dependency on irradiance [Kin04, Hul08,
Hul10, Mid10, Hul11]. Satellite images also provide much better spatial resolution (5-
10 km) compared to ground-based stations, which can be very sparsely distributed in
certain parts of Europe like in Norway. In order to assess the PV potential of a particular
location, spatial interpolation of data must also be performed [Sur07]. All these steps
introduce uncertainties in the final estimates in addition to measurement uncertainties.
Furthermore, spatial interpolation cannot distinguish local microclimates, except those
originating from shadowing in mountainous regions and deep valleys in models that
also use high-resolution terrain elevation data. The contribution of sea surface
reflections in coastal locations like the present one (see Fig. 4.22) is normally not
accounted for. Long-term ground-based and satellite observations can otherwise be very
useful in studies of the year-to-year variability of the solar resource.
Studies of the spatial variability of irradiance caused by clouds involved the use
of multiple sensors installed a few hundred to a few thousand meters apart from each
other. Short-term temporal variability required resolution in the order of seconds. For
PV applications, some of the sensors were measuring the irradiance in tilted planes
[Ham12]. Unlike the standard horizontal irradiance measurements, tilted irradiance data
generally cannot be directly compared (even for the same location) due to possible
differences between the chosen tilt angles and azimuthal orientations. For example, the
seasonal irradiation distribution from year 2011 obtained in the present work can be
64

compared only qualitatively to that from 2005 in [Mid10], since the two data sets were
obtained for different inclinations and azimuthal orientations of the irradiance sensors.
There are multiple solar databases for Europe at present, which cover different
periods of time and may result in different estimates for a PV system performance in a
particular location [Hoy08]. For example, the PVGIS tool allows the selection of two
solar databases for the region of Veliko Tarnovo in Bulgaria, where two 20-MW PV
systems have begun operation recently. In the plane of an optimally inclined c-Si PV
array, the long-term irradiation estimates based on the two databases differ by 8.7 %
[Hul12, PVGIS]. It is possible that this difference is due to an actual short-term climatic
change, because the data from ground-based observations covers the period from 1981
through 1990, whereas the satellite data are from the years between 1998 and 2010. For
the region of the present test setup, the average annual horizontal irradiation of several
databases has an uncertainty of about 15 % [Hoy08, Sur08b]. The discrepancies for the
two example European locations are within the 10-15 % possible year-to-year
variability found in the literature, and therefore it is not possible to attribute them to any
particular cause. However, from an economic perspective, a long-term uncertainty of ±5
% may be detrimental for the profitability of big PV projects, and therefore this is the
prediction accuracy level sought by investors [Wil06, Gue09]. The uncertainty of the
resource in the coastal regions of Southern Norway is well above this value. On the
other hand, both the earlier studies and the present work suggest local boost of the
annual irradiation [Ols86, Mid10]. Therefore, the establishment and the future
maintenance of the present solar research platform are of high significance for
prospective local PV projects. At present, there are plans for complementing it with a
horizontal irradiance sensor.

4.3.1 GROUND-BASED MEASUREMENTS: TYPES OF IRRADIANCE SENSORS


The instantaneous solar resource available at the earth’s surface cannot be
estimated reliably from satellite imagery analysis without calibration and validation by
means of accurate surface-based measurements [McA04]. Ground-based measurements
of the direct, diffuse and global irradiance, as well as of the solar spectrum, have been
performed by numerous research labs and meteorological stations. In addition, ground
albedo has been estimated by means of coupled irradiance sensors – one facing up and
recording the global horizontal irradiance, the other facing down and registering the
ground-reflected light. The present work uses only tilted irradiance data measured in the
plane of the tested PV modules. In addition, the short-circuit currents of the latter are
used in the estimation of their self-referenced irradiances and for obtaining general
information about the spectrum of light. Therefore, this part of the thesis provides only a
brief review of the instruments relevant for the present work, focusing on their typical
uncertainties and response times. Separate measurements of direct and diffuse sunlight,
as well as detailed spectrum characterization, are outside the scope of the present work.
65

Thermopile pyranometers have been used for measurement of global and diffuse
horizontal irradiance for decades. They are integrators of radiation power density with a
broad spectral response ranging between about 300 and 3000 nm [Mye89, Glo08]. In
addition to the classic all-black pyranometers, also black-and-white pyranometers are
used solely for measurement of diffuse horizontal irradiance [Nas83, Mye02]. In 1965,
the World Meteorological Organization (WMO) classified the pyranometers in three
groups according to their accuracy and time of response to a step input [Gar93a,
Gar93b]. The response time is defined as the interval in which the output signal reaches
either 90 % or 95 % of the final steady-state signal, and may differ for rising and falling
step inputs (rise and fall times). This parameter was within 25 s for 1st class
pyranometers, within 1 min for the 2nd class, and within 4 min for the 3rd class.
Measurement errors can originate from zero offsets, angular dependency of
responsivity, non-linearity, long-term instability, limited sensitivity, variations of the
ambient temperature, deviations from desired tilt, as well as from spectral effects
[Nas83, WMO08]. The total uncertainty of sub-hourly global irradiance measurements
can approach ±5 %, whereas results from laboratories using different pyranometers may
disagree by up to 10 % [Mye89, Mye09a]. Averaging of data from multiple sensors may
cancel some of the random variability between them, but such redundancy can be very
costly due to the high prices of thermopile pyranometers, and differences exceeding 5 %
are still possible (especially in winter). Devices with a single responsivity (defined at
solar zenith angle of 45°) can disagree by more than 20 % at zenith angles exceeding
70° due to various cosine responses. There is not a typical response curve, and the
response for a given solar zenith angle may differ in morning and in afternoon [Nas83,
Mye02]. Azimuthal angle dependency causes variations of responsivity between 3 %
and 15 % (highest for black-and-white pyranometers). These are in most cases
negligible because they are only important at very large angles of incidence of the beam
component resulting in low irradiances. The responsivity gradually degrades over time
[Nas83].
Thermopile pyranometers have been refined significantly in the past decades and
may attain accuracies of about 2 % in steady-state solar radiation measurement, whereas
the typical time constants are between 1 and 10 s. For a pyranometer with a time
constant of 4 s, inaccurate readings will be made for about 15 s following a rapid drop
in irradiance. Dynamic response models can be used for almost complete elimination of
such errors. The errors due to slow response time tend to cancel with integration of
irradiance [Sue90]. Modern pyranometers are classified according to more recent
standards and guides [Geu06, Kip12a, Kip12b]. The WMO guide dated 2008 also
classifies the pyranometers in three groups: high quality, good quality and moderate
quality [WMO08]. The best quality instruments have response times within 15 s,
achievable uncertainty of 3 % of the hourly totals and 2 % of the daily sums. Accuracy
improvements have been achieved by detailed uncertainty analyses, accounting for
known bias errors, and the introduction of angle-dependent responsivity [Nas83,
66

Mye02]. For solar zenith angles exceeding 80°, the typical correction needed is about 15
% of reading. All-black pyranometers have an inherent negative site-dependent bias
error in the field. In an improved calibration procedure, bias errors due to diffuse
measurement of the order of 20 W/m2 were removed from the reference irradiance, and
another 15-30 W/m2 were removed by accounting for zenith-angle dependency of
pyranometer responsivity [Mye02]. The best measurement accuracy for broadband
radiation (about 1.8 %) is achieved using the responsivity for the zenith angle at the
time of the measurement. Using a single responsivity results in an uncertainty of 3-5 %
or larger for the global horizontal radiation [Mye04]. Secondary standard pyranometers
have a daily irradiation error below 3 % and less than that for longer time series
[Got10]. When tested in a single location, the thermopile pyranometers used by several
European laboratories disagreed by 3.7 % (4.5 % when also the reference sensors were
considered). The implication for PV energy modeling was that prediction could not be
more accurate than 4.5 % [Got10].
Broadband sensors (thermopile pyranometers) cannot accurately determine the
energy yield of a PV system also due to the spectral and angular sensitivity of PV
devices. For such purposes, a calibrated solar cell of the same technology as the PV
system (a reference cell) is more suitable [Glo08, Jag09]. The high cost of classic
pyranometers has led to the introduction of the more affordable Si-photodiode
pyranometers and PV-based sensors. With correction for the systematic influences,
photodiode-based sensors achieved the same accuracy of the total irradiance
measurements as the thermopile pyranometers [Kin97b]. The two newer sensor types
have a nearly instantaneous response. The typical response of c-Si cells is about 10 μs
[Sue90]. PV-based sensors usually have a planar glass front surface, thus being very
sensitive to reflectance losses at ‘flat’ angles of incidence exceeding 60°. The
temperature variation of responsivity of c-Si PV devices is less than 0.1 %/K.
Acceptable accuracy (±3 %) is achievable for low-cost PV-based irradiance sensors by
applying empirical corrections that compensate for the systematic influences of the solar
spectrum, AOI and operating temperature [Kin98a, Jag09]. Some devices have a
temperature compensation implemented on a hardware level [Sol12].
Measurement of inclined (tilted) irradiance in the plane of PV modules has
advantages over GHI recording. The typical uncertainty in the prediction of the long-
term energy yield of a PV system is 3 % for measurements in the plane of the array
versus 5 % for measurements of GHI [The10]. Sensors facing up do not account for
albedo effects. However, some types of thermopile pyranometers are not suitable for
inclined positions because of changes in the convective flow inside the glass
hemispheres. Black-and-white pyranometers were found to suffer most from tilt-related
errors [Nas83]. In a recent study, 13 types of solar cell sensors and 4 types of Si-
photodiode pyranometers were tested in 2 sites in Germany [Zeh09]. All sensors were
mounted at the optimum inclination angle for the particular site. Measurements were
taken at 15-second intervals. The sensors’ readings were compared against a reference
67

thermopile pyranometer, for which detailed information was available about the
angular, temperature and spectral dependency of responsivity. It was found that Si-
photodiode pyranometers could overestimate the integrated solar energy by up to 6 %,
but underestimation by –10 % was found possible for individual models. Most of the
non-amplified cell-based sensors tended to underestimate the irradiation by –3 % to –6
%, but individual specimens deviated by –10 %. The instantaneous deviations of some
sensors varied significantly with irradiance level. Most of the devices over-estimated
irradiance at levels below 200 W/m2. In general, typical sensor behavior for a given
sensor type could not be defined. The authors stressed on the importance of knowing the
operational behavior of each individual irradiance sensor [Zeh09]. The main conclusion
from that work was that irradiance measurements with an individual low-cost sensor
could be very inaccurate (well outside the acceptable ±3 % accuracy range). Therefore,
at least two instruments of different type should be used in order to improve the
accuracy. Such a redundancy approach is taken in the present experimental setup in
Grimstad, which uses an amplified all-black thermopile pyranometer and an amplified
mc-Si PV cell with hardware compensation for the temperature effects. Details about
the sensors are given further in Section 4.5.

4.4 ANALYSIS OF LOCAL METEOROLOGICAL DATA

Figure 4.1: Annual global horizontal irradiation data for Landvik from [Bio12].
68

The meteorological station which is closest to the present test site is located a
few km away in Landvik. The data recorded there is available via the Bioforsk
AgroMetBase website [Bio12]. Hourly averages of global horizontal irradiance (GHI)
sampled every 5 seconds are available since 1987 [And12]. However, the data from
1987 through 1994 do not exceed 100 W/m2, which indicates a problem with the
database or with the recorded data. The data from 1995 and 1996 have too many gaps
(fields marked ‘NULL’), which add big uncertainty to the annual irradiation. In 2005,
the sensor downtime has been 3 % (mostly data from the 2nd half of May are missing).
Figure 4.1 shows the yearly global horizontal irradiations for years 1997 through 2011.
The mean equals 972 kWh/m2 which is about 7.6 % below the Norwegian maximum
estimated in [Ols86] for the present region. The maximum year-to-year variability is
about 200 kWh/m2 (20.5 %), whereas the standard deviation is 5.5 %. This variability
exceeds the typical values found in the literature. Note that the irradiation for 2005 is
underestimated due to missing data from the late spring (see Fig. 4.2). May is a strong
contributor to the annual irradiation (see Fig. 2.3 in Chapter 2, as well as Fig. 4.34
further in this Chapter), with a share of over 11 % in 2005 [Mid10] and 13 % in 2011
according to results from the present work. These shares are for tilted irradiation, in
which the diffuse component has a lower share than in the horizontal irradiation. The
direct component has a lower share in the horizontal irradiation because of the large

Figure 4.2: GHI evolution in 2005 for Landvik [Bio12].


69

cosine loss at low solar elevations, whereas the ground-reflected component does not
contribute at all. On the other hand, direct irradiance is available to horizontal sensors
from wider range of solar azimuthal angles in the summer than to tilted sensors. As a
result, the monthly horizontal irradiations in the summer may exceed the corresponding
tilted irradiations (see Figs. 4.3, 2.3 and 4.34). Note that the value for May, 2005 given
in Fig. 4.3 is underestimated due to missing data. Assuming about 50 kWh/m2 for the
skipped time, the irradiation in 2005 is approximated to 1020 kWh/m2. Consequently,
the mean for all 15 years becomes 975 kWh/m2, 7.2 % below the Norwegian maximum
given in [Ols86]. The year-to-year variability is practically unaffected by the correction.

Figure 4.3: Seasonal distributions of the global horizontal irradiation in Landvik.

The seasonal distribution study allows deeper analysis of the big difference
between 1997 and 1998 seen in Fig. 4.1. In the latter year, less irradiation was received
in most of the months from April through September (see Fig. 4.4). The reduction was
particularly strong for April (50 %). For all the 15 years, August had the lowest
irradiation variability with respect to the mean (6 % in terms of standard deviation; 20
% maximum variation). Significant variability is seen for the spring and autumn months
(reaching 65 % of the mean), and even higher for the winter months. However, the latter
contribute very little to the annual energy yield of a fixed, optimally inclined PV array.
70

Figure 4.4: Seasonal distributions of the global horizontal irradiation for years
1997 and 1998. The data are from [Bio12].

Figure 4.5: Comparison of the average monthly global horizontal irradiations from
[Bio12] versus those from [PVGIS].

The PVGIS tool estimates an average annual global horizontal irradiation of 889
kWh/m2 for the present test site and 885 kWh/m2 for Landvik – nearly 10 % below that
estimated from the Bioforsk data set and over 15 % below the estimate in [Ols86]. This
may indicate a systematic error in the solar data used by [PVGIS] for the region of
interest. Comparison between the average monthly irradiations estimated from [Bio12]
71

and [PVGIS] is given in Fig. 4.5. The data sets agree very well in autumn and winter,
but the latter one seems to underestimate the solar resource available in spring and
summer by 10-15 %. A possible implication from this is a slight underestimation of the
optimum tilt angle calculated by the PVGIS for a fixed, south-facing PV array due to
the lower expected share of direct versus diffuse irradiation.

4.5 IRRADIANCE MEASUREMENTS IN THE PRESENT WORK

Figure 4:6: The irradiance sensors used in the experimental setup in Grimstad.

For accurate real-rime monitoring of irradiance, a temperature-compensated mc-


Si PV-based sensor, SolData 80spc, is used in combination with an isolation amplifier
[Sol12, PRE12]. The total calibration uncertainty of the cell is 3 %, whereas ±2 %
uncertainty under “ideal conditions” is suggested by the manufacturer [Sol10]. The
stated response time of the amplifier to a step increase (to 90 % of the final steady-state
value) is below 25 ms. In addition, a secondary class (according to ISO 9060:1990)
thermopile pyranometer is used together with a current loop amplifier [Kip12c,
Kip12d]. The calibration uncertainty of the pyranometer is 3.4 %; whereas the amplifier
error due to non-linearity, inaccuracy and temperature dependency of gain is an order of
magnitude lower [Kip10, Kip12e]. The spectral response is between 300 and 2800 nm.
The response time is within 18 s for the pyranometer and 0.44 s for the amplifier. The
output of the latter is in the range 4–20 mA, which is converted to 2–10 V by means of a
0.5 kΩ resistor and corresponds to an irradiance range of 0–1600 W/m2. Both
instruments are installed at the same tilt angle, co-planar with the array of PV modules
under test (see Figs. 4.6, 3.7 and 4.25).
72

A surprising result from the present work is the measurement of unexpectedly


high peak irradiance values at such high latitude [Ime11]. On clear-sky summer days,
the local irradiance (in an optimally inclined plane) reaches about 1000 W/m2 at noon.
By chance, this equals the level defined in the standard testing conditions (STC) for PV
modules. In the presence of clouds, however, multiple short bursts close to 1400 W/m2
were found possible (see Fig. 4.7). The peak irradiance values reported in the earlier
study in Grimstad were only about 1150 W/m2, which can be attributed to the poor
temporal resolution of the recorded data set [Mid10]. Surprisingly, the maxima
measured in the present work exceed the extraterrestrial solar irradiance (which is not
attenuated by the atmosphere) at altitude of only about 60 m a.s.l. In year 2011, the
annual maximum of 1413 W/m2 was recorded in May, but in September the irradiance
reached 1390 W/m2. In fact, all months from April through September had peaks
exceeding 1300 W/m2. In summer 2012, bursts exceeding 1500 W/m2 were recorded.

Figure 4.7: Plane-of-module irradiance profiles from a partially overcast day and a
clear-sky day measured in Grimstad, Southern Norway (at latitude 58°20’N).

The supplier of the PV-based sensor did not anticipate as strong overirradiances
as the ones obtained in the present work, and therefore its range was limited up to
1354.3 W/m2. This has been revealed by analysis of irradiance data recorded during an
73

I-V curve sweep that occurred at conditions of very strong cloud enhancement (see Fig.
4.8). Otherwise, the very short response time of this sensor (although impaired by the
amplifier) makes it very suitable for overirradiance studies, whereas the pyranometer
reacts with significant delays and smooths out important details (see Fig. 4.9).

Figure 4.8: Upper cut-off of the PV-based irradiance sensor.

Figure 4.9: An overirradiance event as recorded by each of the two sensors.

The pyranometer is more difficult to clean from snow and ice in the winter. A
rubber tool marketed for cleaning of glass windows is used to clean both the sensors and
the PV modules. The PV-based sensor has a flat geometry and is heated by sunlight. Ice
74

that cannot be removed completely eventually melts at sufficient irradiance. This is not
the case with the pyranometer. Its curved glass dome can retain certain amount of ice,
and seems to heat very little – indirectly, via convection from the heated black absorber
disk. Such situation occurred on 7 January, 2011. On the following days, the author
removed any ice from the dome by heating it with the flat of a hand. However, on that
particular day the dome was not cleaned completely, which resulted in a bias with
respect to the other sensor (see Fig. 4.10). Therefore, using two sensors is advantageous.

Figure 4.10: Pyranometer bias on 7 January, 2011 caused by ice on its dome.

Figure 4.11: Irradiance dependency of ISF on a clear-sky day (discrete symbols).


75

As already discussed in Chapter 3, irradiance stability during I-V curve sweeps


is determined from the reading of the PV-based sensor. It is used later on as a quality
indicator in data filtering prior to deriving PV module parameters, e.g. the coefficients
of the relative efficiency model described in Chapter 1. Unfortunately, the SNR of that
sensor degrades significantly at lower irradiances due to EMI from the utility grid (see
Fig. 3.16). As a result, even on clear-sky days, the ISF estimated and recorded in
LabVIEW is irradiance-dependent (see Fig. 4.11). The data were fitted versus irradiance
with the function 0.0076+0.169exp(–0.0102G), where all coefficients are empirical and
G has units of W/m2. All I-V curve parameters recorded at ISF exceeding 1.2 times the
above value are disregarded. This irradiance dependency is attributed to the range of the
amplifier’s output voltage which is 0–10 V. At low irradiances, the signal level falls
well below 1 V, whereas the noise magnitude is relatively constant. Consequently,
poorer SNR is obtained, which leads to artificially high ISF even at constant irradiance.

Figure 4.12: Power spectrum of irradiance data recorded during an I-V curve
sweep (shown up to 0.1 W2/m4). The peak at 0 Hz corresponds to the useful signal.
Noise modes oscillating at 50 Hz and 665 Hz are also seen.

According to the data sheet, the amplifier allows one to offset the output signal
level corresponding to zero irradiance by 20 % of full range (to 2 V) [PRE12].
However, setting a new input range (in order to record correctly also the strongest
overirradiances) would not be that straight-forward, because neither the data sheet nor
the manual gives the necessary DIP-switch combinations. According to the
manufacturer, non-standard input- and output-signal ranges can only be implemented on
a software level by an authorized agent [PRE11]. It seems that even the standard output
range 2–10 V cannot be implemented on a hardware level with the present configuration
76

of the amplifier, which does not correspond to any DIP-switch combination given in the
data sheet. As of 9 March, 2012 the amplifier is programmed with the desired ranges,
just in time for this year’s overirradiance measurement campaign [Yor12b].
Overirradiances up to 1600 W/m2 can now be recorded with the fast sensor. Preliminary
results from 2012, together with the results from 2011 presented later in part 3.5.3 of the
thesis, were presented recently in [Yor12b, Yor12c].

Figure 4.13: Same as Fig. 4.12, but at smaller scale revealing other noise modes.

Figure 4.14: Physical ISF estimation after cancelling the noise in irradiance data.
77

As of 5 March, 2012, the irradiance recorded during the I-V curve sweeps is
filtered by applying a moving average smoothing, with a window width of 81 data
points. This number corresponds approximately to one complete period of the main
periodic mode of the noise which is 50 Hz (see Fig. 4.12). Modes of lower magnitudes
have been identified at frequencies 150 Hz, 0.7–0.8 kHz and 1.7–1.9 kHz (see Fig.
4.13). Only the third harmonic of the grid frequency is fixed, whereas the frequencies of
other modes vary. Detailed analysis of the noise’s power spectrum is outside the scope
of the thesis, but may be performed in future. In the present work, the identification and
filtering of the main mode has been sufficient for revealing the actual degree of stability
of irradiance during each I-V curve sweep (see Fig. 4.14). ISF is usually within 0.5 %
on clear-sky days, except for irradiances below 100 W/m2, at which it rises due to the
very low level of the useful signal, whereas some noise still remains after the filtering.
The ‘real-time’ irradiance data set recorded in the present work (amounting to
nearly 200 megabytes for a single summer day, 38 gigabytes in total for 2011) is also a
serious contribution to the power electronics community, in particular to designers of
MPP tracking (MPPT) algorithms for PV inverters. Also other studies have revealed the
benefits of recording irradiance with very high temporal resolution. Sampling every
second was found advantageous over using 15-second or 1-minute resolutions, as
changes happened on sub-minute time scale [Vod04]. The resolution was limited by the
relatively high response time (0.1 s) of the reference cell. Very few laboratories in the
world were reported to record irradiance with resolution below 1 minute. A good
weather data set is needed in order to identify the typical operation conditions of PV
inverters at a given location. In-plane irradiance and module temperature over a whole
year need to be recorded with a high resolution [Ble08]. In the latter study, data sets
with resolutions between 1 s and 5 min were used. Another study found that resolution
better than 10 min was generally recommended [Zhu11]. 10-second series of global
horizontal irradiance were used from which condensed subsets of lower resolution were
created. The main effect of reducing the resolution was seen at high irradiance levels
(above 750 W/m2). Decreasing the resolution from 10 s to 1 min did not affect
significantly the calculated optimum inverter sizing ratio, but a further reduction to 10
min or 1 hour had a strong impact. A different research group suggested 1 minute as the
optimum resolution for irradiance [Che10]. A poorer resolution might overlook high
irradiance peaks, while a better resolution (such as 10-second) often did not provide
further improvement. However, 25 % of the overirradiance events in the example US
location lasted less than 1 minute. The location of the present test setup is a coastal one,
and therefore the typically high wind speeds are expected to cause more short-lived
irradiance bursts. Therefore, the high resolution of 10 ms used in the present work is
very beneficial for such analysis.
78

4.6 EFFECTS OF LOW TEMPORAL RESOLUTION OF IRRADIANCE DATA

Figure 4.15: Two 20-minute subsets created from the same 1-minute data set.

The optimal temporal resolution of irradiance measurements can differ


depending on the research task. One goal of the present work is to compare the seasonal
distribution of solar irradiation in 2011 to that from 2005 given in [Mid10], see Fig. 2.3.
That study used 20-minute sampling, which is expected to have added extra uncertainty.
This Section investigates the magnitude of the additional uncertainty of the monthly
irradiation introduced by sampling irradiance with finite temporal resolution. This is
79

done for each month in year 2011 by creating subsets of lower resolution from the two
irradiance data sets recorded in the present work. From the 1-minute data set, 20 subsets
with resolution of 20 minutes are created (see Fig. 4.15), and the monthly irradiation
calculated from each of these is compared to that obtained from the full data set. In this
way, the experiment conducted back in 2005 is simulated for the two irradiance sensors
used in the present work. This approach is applied also to the 10-millisecond ‘real-time’
data set. For each month, about 6000 subsets with 1-minute resolution are created, and
the corresponding estimates of the monthly irradiation are compared against that from
the complete data set. This gives insight into the magnitude of the extra uncertainty
contained in the estimates from the recorded 1-minute data set. The 10-ms data set
recorded with the fast sensor is considered to be ‘real-time’, because drastic changes of
irradiance do not occur on a time scale finer than this. Therefore, it is not associated
with any additional uncertainty due to discrete sampling. This is not the case with the
slow sensor due to its long response time (in the order of 20 seconds).

Figure 4.16: Comparison between the two 20-minute subsets that gave the
maximum and the minimum monthly irradiation estimate for January, 2011.

The monthly irradiation estimate is obtained from each subset by integration


over time. However, the irradiance profile of a given day can look very different
between any two subsets (see Fig. 4.16). These differences propagate into the monthly
irradiations. Because a 20-minute time step results in coarse integration, the estimates
are generally different from that obtained from the full (1-minute) data set. Fig. 4.17
shows a frequency distribution of the deviations (errors) for the 20 subsets for January,
2011. What is important in this statistical analysis is the maximum error for this month
for the sensor under consideration. This error should be interpreted as the added
80

uncertainty due to discrete sampling with 20-minute resolution compared to sampling


every minute. For example, the uncertainty corresponding to Fig. 4.17 is about 1.3 %.
The obtained error distributions differ a lot between the different months. They are
generally asymmetric with respect to zero. The same conclusion is valid also for the 1-
minute subsets created from the real-time data set (see Fig. 4.18), but the maximum
errors are of much smaller magnitude. The results from the statistical analysis of the
error distributions are summarized in Tables IV.1 and IV.2.

Figure 4.17: Frequency distribution of the monthly irradiation errors of all the 20-
minute subsets for January, 2011 for the fast irradiance sensor.

The maximum monthly irradiation uncertainties due to 1-minute sampling are


obtained for the months with lowest shares in the annual irradiation – February,
November and December (see Fig. 4.34 further in Section 4.8). This observation is valid
for both the fast and the slow sensor, but the uncertainties are much larger for the latter,
which is attributed to its long response time. The added uncertainty for the irradiation in
2011 is estimated by summing the monthly uncertainties expressed in units of kWh/m2.
This is done by multiplying the corresponding maximum errors in Table IV.1 with the
monthly irradiations based on the fast sensor (respectively, the data in Table IV.2 is
used in the case of the slow sensor). The uncertainty due to 1-minute sampling for 2011
81

is 0.26 % of the annual irradiation for the fast sensor, and 0.33 % for the slow sensor.
These values are an order of magnitude below the calibration uncertainties of the
sensors. Interestingly, the numbers are practically equal for both sensors, despite their
very different (by 4 orders of magnitude) response times. Therefore, it is concluded that
for the local climatic conditions and irrespective of the sensor type, irradiance sampling
with 1-minute resolution does not add significant uncertainty to the annual sum. On the
other hand, such discrete sampling shows a clear trend towards overestimation of the
monthly irradiations relative to the ‘real-time’ measurement.

Figure 4.18: Frequency distribution of the monthly irradiation errors of all the 1-
minute subsets for January, 2011 for the fast irradiance sensor.

Much larger are the uncertainties added due to 20-minute sampling. Again, the
highest values are obtained for the three months with the lowest contribution to the
annual irradiation. For these months, the extra uncertainty is comparable to the
instrumental uncertainties of the sensors. The corresponding error spans in Tables IV.1
and IV.2 are more symmetric with respect to zero (compared to the case of 1-minute
sampling). The additional uncertainty of the annual irradiation due to such coarse
sampling is 1.6 % for the fast sensor and about 1.4 % for the slow sensor. The first
number is therefore the approximate uncertainty of the annual irradiation obtained in
[Mid10] for 2005 solely due to using 20-minute sampling. Note that this estimate is
82

based on the present 1-minute data set obtained with system uptime practically equal to
100 %. It excludes the uncertainty due to inferior availability (uptime) of the old data
acquisition system.

Figure 4.19: Correlation of the monthly sampling uncertainties of the two sensors.

TABLE IV.1: Monthly irradiation errors of irradiance subsets for the PV-based sensor.
Month Error Span [%], 20-min Subsets Error Span [%], 1-min Subsets
January –0.3/+1.3 0.02/+0.38
February –1.1/+3.5 –0.09/+0.62
March –1.2/+1.6 –0.15/+0.21
April –0.8/+1.0 –0.05/+0.07
May –1.9/+1.7 –0.14/+0.25
June –0.5/+1.4 –0.07/+0.21
July –0.8/+0.8 –0.03/+0.25
August –0.6/+1.7 –0.08/+0.23
September –1.1/+2.0 –0.26/+0.40
October –1.0/+2.0 –0.11/+0.23
November +0.5/+3.1 –0.13/+0.98
December +0.8/+3.7 +0.10/+0.67
83

TABLE IV.2: Monthly irradiation errors of irradiance subsets for the pyranometer.
Month Error Span [%], 20-min Subsets Error Span [%], 1-min Subsets
January –0.5/+1.1 +0.08/+0.51
February –1.3/+2.9 +0.29/+1.23
March –1.1/+1.3 –0.12/+0.19
April –0.7/+0.9 –0.02/+0.08
May –1.5/+1.7 –0.06/+0.25
June –0.4/+1.4 –0.05/+0.20
July –0.7/+0.7 +0.03/+0.23
August –0.9/+1.4 0/+0.32
September –1.4/+1.7 –0.05/+0.49
October –1.4/+1.4 +0.14/+0.56
November +0.3/+2.6 +0.30/+1.20
December +0.3/+2.7 +0.39/+1.17

Figure 4.20: Fast sensor sampling uncertainties versus monthly solar irradiations.

There is a very good and positive linear correlation of the monthly sampling
uncertainties between the two sensors (see Fig. 4.19). The correlation coefficients are
0.91 for 1-minute sampling and 0.98 for 20-minute sampling, both being very close to
unity. On the other hand, the correlation is much weaker between the sampling
84

uncertainties and the corresponding monthly irradiations (see Fig. 4.20; the plots for the
other sensor look qualitatively very similar). Therefore, the magnitudes of the sampling
uncertainties of the sensors in a particular month can be attributed to the shape of the
corresponding real-time irradiance profile. All monthly profiles can be analyzed in the
frequency domain by means of their power spectra. This computationally-intensive and
time-consuming task is beyond the scope of the present work, but may be attempted in
future.

4.7 DATA RECOVERY AND SELF-REFERENCED IRRADIANCE

Figure 4.21: Plane-of-module irradiance measured on 23 January, 2011.

Albedo effects can have a significant impact on irradiance measured in an


inclined surface. Unlike in the case of global horizontal irradiance (GHI)
measurements, ground-reflected light is well within the field of view of sensors installed
co-planar with PV modules tilted with respect to the horizontal. Snow and ice during
winter, as well as rain water in all seasons can increase the ground albedo and add
another major component to the direct and diffuse irradiance. Light reflected from snow
is diffuse in its nature, but ice and liquid water can cast specular reflections on inclined
PV modules and irradiance sensors. In such situations, the measured irradiance profiles
85

on clear-sky days can differ significantly from the typical bell shape (e.g. the red curve
in Fig. 4.7). Fig. 4.21 shows an example from a mostly clear-sky winter day in January,
2011. On that day, overcast weather took place until 10:30 a.m. and after 4 p.m., while
clear sky conditions prevailed in between. The readings from the two irradiance sensors
disagreed notably for more than two hours about noon. The strange shape was being
caused by specular reflection of sunlight from a solar thermal module installed in
horizontal position in front of the PV test installation located in Grimstad (see Fig.
4.22). The solar thermal module was later covered to prevent reflections from its front
glass. However, some reflections could not be avoided like those from ice on the
building’s roof, its metallic elements, as well as from the sea surface. The latter could
generally benefit an on-shore PV system, but its contribution would be rather small due
to the flat angle of incidence for optimally inclined modules and the associated cosine
and reflection losses. The reflections from nearby objects, however, can cause uneven
illumination on the PV arrays and lead to current mismatch between the modules. In the
present work, such reflections made the accounting of solar energy very challenging
during the winter months, whereas some of the recorded module data were strongly
contaminated. In general, such non-optimal configurations should be avoided in outdoor
characterization of PV modules [Got10], but this was not possible in the present study.
Therefore, techniques were needed for recovering the lost irradiance information, and
for filtering out the erroneous module and sensor data.

Figure 4.22: Southern view in front of the irradiance sensors on 20 January, 2011.

Fig. 4.23 illustrates a different situation which results in biased readings from
the pyranometer (see Fig. 4.10). It is possible to use the short-circuit current ISC of a PV
86

device as an alternative measure for the in-plane irradiance after calibration against a
reference PV device [Jag09]. The self-referenced irradiance (self-irradiance) derived
from ISC has several advantages when estimating various PV module parameters. It is
equivalent to using an irradiance sensor which is perfectly matched to the tested PV
device in terms of spectral, reflection and temporal response, and provides good
correspondence between indoor and outdoor characterization [Bet09, Jag09, Zhu09,
Moh10]. The short-circuit current ISC of a c-Si PV module changes almost linearly with
the in-plane irradiance, except for flat angles of incidence at which the reflectance loss
is greatly increased [Kin96, Mar01]. Its temperature dependency is usually very small
[Kin97a]. Another issue resulting in deviation from linearity is the spectral dependency
of the device [Kin97c, Per07]. Under some conditions, a PV module’s ISC can be used
as a reasonably accurate measure for the in-plane irradiance. Even in some cases of
partial shadowing, ISC still equals that of the non-shadowed module [Mid10]. This
Section of the thesis describes the self-referencing approach used to recover irradiance
data contaminated by severe albedo effects. Fig. 4.24 shows the calculated self-
referenced irradiances from the eight c-Si modules together with the readings of the two
sensors around noon on 23 January, 2011. For seven of the modules, the estimates show
a very good agreement, whereas the eighth module seems to have been strongly affected
by reflection, similar to the two sensors. That module was installed in a close proximity
to the sensors (see Fig. 4.25), which explained its strongly biased self-irradiance.

Figure 4.23: Frost covering the pyranometer’s dome on a sunny winter day.
87

Figure 4:24: Application of self-referenced irradiance in data recovery.

Figure 4.25: Position of the two irradiance sensors and the eighth module
represented in Fig. 4.24. The lower two modules are not part of the present work.

For all tested modules except the triple-junction a-Si module, ISC scales almost
linearly with the reading of the mc-Si irradiance sensor (see Fig. 4.26). The hysteresis
88

seen in some plots may have originated from the small but non-zero temperature
coefficient αSC of this module parameter, according to the following expression [Jag09]:

G
I SC  I SC,STC 1  SC  T  T0  (4.1)
G0
where ISC,STC, G0 and T0 are the values of the short-circuit current, the in-plane
irradiance and the module temperature at STC, respectively. Significant spectral
mismatch between the a-Si module and the mc-Si sensor can take place, and therefore
this module is disregarded in the present context. For the c-Si modules and the CIGS
module, the values of αSC and ISC,STC estimated from linear regression as in [Jag09] are
listed in Table IV.3. Only data corresponding to irradiance between 950 and 1050 W/m2
(from the pyranometer) have been used, whereas the equivalent cell temperature (ECT)
has been used instead of the back-of-module temperature. The bracketed numbers are
the uncertainties of the preceding two digits at 95 % confidence level.

Figure 4.26: Short-circuit currents of the tested PV modules versus irradiance


measured with the mc-Si sensor. Data are from a clear-sky day (21 June, 2011).

The effects of using irradiance data from the mc-Si sensor are twofold. First, the
estimated values for ISC,STC are about 2-2.5 % larger which is attributed to calibration
89

uncertainties of the two sensors. Second, the estimated temperature coefficients are
generally lower than those in Table IV.3. The differences range between 8 and 20 % for
the c-Si modules, but for the CIGS module a remarkable 50 % lower αSC is obtained. In
addition, larger uncertainties are obtained due to a poorer linear correlation in the data
(up to a factor of 2). In the context of self-referenced irradiance calculation from ISC, the
associated temperature correction would be relatively small, and therefore the
uncertainty in αSC does not affect the overall uncertainty too much. The self-irradiance
definition given below is based on Eq. (4.1):

G0 I SC
GS  (4.2)
ISC,STC 1   SC  T  T0  

Table IV.3: Short-circuit current at STC and its temperature coefficient.


Module ISC,STC [A] αSC [mA/K] αSC [%/K]
Aged BC (‘BP’) 4.706 8.44(10) 0.179(2)
IBC (‘Sunc.’) 5.948 6.29(16) 0.106(3)
‘RS mono’ 8.542 7.88(27) 0.092(3)
‘RS poly’ 8.398 8.46(24) 0.101(3)
‘ESS 1’ SoG-Si 7.861 13.09(21) 0.167(3)
‘ESS 2’ SoG-Si 7.909 11.89(22) 0.150(3)
HIT 7.406 6.95(20) 0.094(3)
Aged mc-Si (‘GPV’) 2.954 6.38(10) 0.216(4)
‘CIS’ 3.094 1.45(8) 0.047(3)

The major uncertainty of GS is that associated with the irradiance measurements


used to determine ISC,STC. Similarly, the label value of this parameter provided by the
manufacturer is affected by the accuracy of the irradiance sensor used in the module
rating procedure. Figure 4.27 contains self-irradiance plots based on label data of the
eight c-Si modules tested in Grimstad, along with measurements from the two sensors
from a sunny day in April, 2011. For simplicity, no temperature correction has been
applied, but very similar operating temperatures of the modules are assumed. The
shapes of all curves are very similar, but there are systematic discrepancies between the
estimates. As expected, the self-irradiances of the two aged modules are the lowest ones
because their ISC,STC values have degraded during outdoor operation for several years.
Averaging of the self-irradiances of the new modules together with the readings
of the two sensors may allow one to determine the real-time irradiance with a reduced
uncertainty, well below the ±3 % typical for good-quality instruments [Jag09]. This
approach is similar to testing at one site the irradiance sensors used for module rating by
the different manufacturers. In an international module intercomparison, agreement of
about ±3 % was achieved of the measured ISC,STC for most of the represented PV
90

technologies [Rum06]. Since ISC scales almost linearly with irradiance, and because
module rating is performed at standardized light spectrum and module temperature, the
differences in ISC,STC obtained by the different laboratories can be attributed mainly to
the calibration errors of their irradiance sensors. Electric current measurements typically
have much smaller uncertainties [Jag09]. One can assume a normal distribution with a
mean equal to 0 for the calibration biases of all irradiance sensors with accuracy ±3 %.
Thus, the standard deviation σ of the population (in a statistical sense) equals (3 %)/1.96
≈ 1.5 % [Hin75]. For a large sample of N sensors, the mean of their biases would be
within about 1.96 N   3 % N around zero with 95 % confidence. Therefore,
achieving accuracy in the order of ±1 % or even better would require data averaging
from N≥9 sensors (or a combination of irradiance sensors and PV modules totaling 9 or
more devices). Exact accuracy evaluation of such averaging is beyond the scope of the
thesis. Instead, the author aims to show the potential for improving the accuracy of
irradiance measurements by using self-irradiances of multiple new PV modules.

Figure 4.27: Self-referenced irradiances of the c-Si modules based on label values
of ISC,STC, along with data from the two sensors. The data correspond to air masses
of about 1.5 (AM1.5).

The averaging concept described above has been used to derive calibration
correction factors applicable to the label values of ISC,STC of the eight c-Si modules and
91

the CIGS module. Data from a clear-sky day (18 August, 2011) recorded at air masses
of 1.5±0.05 and angles of incidence within 20° have been used, and the measured ISC
values have been adjusted to 25°C. The obtained correction factors are listed in Table
IV.4. For the two aged modules, the degradation of current is obvious. Correction
factors have also been evaluated for the mc-Si sensor and for the pyranometer. Their
irradiance readings have been multiplied respectively by 1.027 and 0.997 in order to
obtain agreement with modules’ self-irradiances. These values reveal that the PV-based
sensor has been biased by –3 % with respect to the thermopile instrument, whereas the
readings of the latter have practically coincided with the mean from the sample (of
irradiance-measurement devices). Surprisingly, the correction factors for the two SoG-
Si modules denoted ‘ESS 1’ and ‘ESS 2’ differ by about 1 %. They come from the same
German manufacturer (Q-Cells) and are made of the same Si feedstock, but probably
have been rated using distinct irradiance sensors having different calibration errors.

Table IV.4: Factor applied to the labeled ISC,STC used in self-irradiance calculations.
Module Correction
Aged BC (‘BP’) 0.948
IBC (‘Sunc.’) 1.024
‘RS mono’ 0.982
‘RS poly’ 0.989
SoG-Si ‘ESS 1’ 1.000
SoG-Si ‘ESS 2’ 1.012
HIT 1.016
Aged mc-Si (‘GPV’) 0.936
CIS 0.990

The averaging procedure has been repeated for a clear-sky day in April, 2011.
Almost unchanged correction factors have been obtained for all modules and for the
pyranometer (0.999). However, a notably lower correction (1.011) has been necessary
for the mc-Si sensor. These findings suggest that the responsivity of the pyranometer
has been relatively stable between April and August, whereas that of the PV-based
sensor has degraded by about 1.6 %. Significant degradation of the modules’ ISC,STC
values during that period cannot be expected if one assumes a year-to-year degradation
in the order of 0.5 % [Sak03, Jor10]. This expectation is supported by the almost
constant correction factors that have been obtained for each module in both months.
The present work has indicated that averaging of self-irradiances by taking the
mean may not always be practical. Fig. 4.24 in depicts a situation in which one of the
mc-Si modules and both the irradiance sensors are strongly affected by albedo effects.
Taking the mean of the ten irradiance estimates would result in a strongly biased
determination of the true real-time irradiance. Assuming that the majority of the devices
92

are unaffected by reflections or significant snow cover, the median irradiance estimate
would be much more accurate, like in the case illustrated in Fig. 4.24.

Figure 4.28: Reflectance loss model relative to normal incidence of light for a c-Si
PV module [Mar01]. An average value of 0.164 is used for the shape parameter ar.

Obtaining accurate self-irradiance estimates for all operating conditions is not at


all straight-forward. This would require accounting for reflectance losses (compared to
normal incidence of light) with separate modeling for the direct, diffuse and ground-
reflected (and sea-reflected) components. A simple Fresnel model or a more advanced
model can be used for the direct component, see Fig. 4.28 [Mar01, Bet09]. However,
the empirical parameters of the model must be identified separately for each module,
which is difficult for fixed tilted modules tested outdoors. Also, reflectance modeling
becomes very sensitive to errors at flat AOI. Furthermore, reflectance losses are very
difficult to model for the diffuse component due to its distributed nature (angles of
incidence between 0° and 90°) and its general anisotropy. In addition, the individual
spectral response of each module and the instantaneous light spectrum must be known.
Fig. 4.29 shows the effect of increased reflectance losses on self-irradiance at
flat angles of incidence of direct sunlight. The mean self-irradiance of the nine modules
(with temperature correction) is plotted together with the readings of the two irradiance
sensors from a clear-sky summer day. In the early morning, as well as in the evening,
both the mean self-irradiance and the reading of the PV-based sensor are much lower
than the reading of the pyranometer, while being practically equal to each other during
the whole day. The PV devices have flat geometry, and therefore one should expect
very similar angular dependencies of their reflectance losses for all the components of
sunlight. For clear-sky days, one could model these losses as a function of the AOI
93

calculated with an astronomical algorithm. However, the relative contributions of the


direct and diffuse components on a partially overcast day are difficult to model, unless
they are measured individually.

Figure 4.29: Mean self-irradiance of the eight c-Si modules and the CIGS module
on 21 June 2011, along with irradiance measured by the two sensors.

Figure 4.30: Partial shadowing of the modules on 12 January 2011, when the snow
was cleaned off the modules shortly before 1 p.m. The image is from 12:42 p.m.
94

Figure 4.31: Readings of the irradiance sensors and modules’ self-irradiances from
12 January 2011, along with a part of PV-based sensor’s reading from 7 January.

Modeling complications arise also from the spectral dependency of a PV


module’s current response. Due to diurnal variations of the solar spectrum, the ratio
ISC/G of a PV module is not constant. For mono-Si modules installed in Madrid, this
ratio varied by about 3 % during winter and within 2 % in summer [Per07]. However,
variations exceeding 10 % were reported in [Kin97c] for individual c-Si devices. This
PV technology performs better under redder light associated with low solar elevations,
and the latter occur more frequently in Norway than in Spain. Therefore, the spectral
effects on PV have greater significance in the present test setup than at lower latitudes.
In general, irradiance sensors based on PV cells would require spectral corrections in
order to improve their accuracy. The air mass (estimated from the solar elevation) can
be used for applying a spectral correction to the module’s ISC (given the module’s
spectral response), but the procedure requires knowledge of the direct and diffuse
irradiance, plus two calibration constants [Kin97c]. For the present test location, the
possibility for significant sea-reflected irradiance component makes the modeling even
more complex. In the present work, the application of module self-irradiances is limited
to determination of module parameters from I-V curves and restoring irradiance data
95

corrupted due to reflections, as well as due to snow or ice cover. Another example of a
situation requiring data recovery is illustrated by Figs. 4.30 and 4.31.
On 12 January 2011, the snow that had accumulated on the modules and on the
sensors was cleaned off shortly before 1 p.m. instead of in the morning. Images from the
TV camera revealed that the conditions had been sunny until about 12 p.m. and that the
snow on some of the modules had started to slide down after 11 a.m. (see Fig. 4.30).
The module denoted ‘RS poly’ had its top substring of cells snow-free in a short while,
followed closely by the adjacent module ‘RS mono’. This is evident also from their self-
irradiance plots in Fig. 4.31, which overlapped as of 11:20 a.m. Later on, the module
denoted ‘Sunc.’ was uncovered sufficiently and its self-irradiance equalized with those
of the former two modules. About the same time, overcast conditions occurred. The
aged BC module denoted ‘BP’ was not sufficiently freed of snow until it was manually
cleaned by the author in 12:45 p.m. Artificial shadowing from the cleaning tool resulted
in an incomplete I-V curve, a strongly over-estimated ISC, and a wrong self-irradiance
exceeding 1200 W/m2. For that particular day, the self-irradiance of the module ‘RS
mono’ is considered as representative for the true in-plane irradiance after 11:20 a.m.
The earlier irradiance profile is approximated with mc-Si sensor’s data from the clear-
sky morning of 7 January. The so recovered irradiance data correspond to irradiation of
about 0.7 kWh/m2 that had not been accounted for by the pyranometer due to snow.

Table IV.5: Uncertainties of the modules’ self-referenced irradiances GS introduced by


the electronic loads’ current samples.
Module ΔGS [W/m2]
Aged BC mono-Si (‘BP’) 21
Aged mc-Si (‘GPV’) 17
Back-contact mono-Si (‘Sunc.’) 8
RSS mono-Si (‘RS mono’) 12
RSS mc-Si (‘RS poly’) 12
Sanyo HIT (‘HIT’) 13
SoG-Si 1 (‘ESS 1’) 13
SoG-Si 2 (‘ESS 2’) 12
AvanCIS (‘CIS’) 16

The additional uncertainty of current measurements introduced by the electronic


load’s current sample (0.5 % of full scale) becomes significant at low irradiance levels,
and affects the self-irradiance accuracy. The modules denoted “RS mono”, “RS poly”,
“HIT”, “ESS 1”, “ESS 2” and “BP” are equipped with loads rated 20 A, whereas for
“GPV”, “Sunc.” and “CIS” the full current scale is 10 A. The corresponding
uncertainties of 0.1 A and 0.05 A, respectively are added to the instrumental uncertainty
of the NI PCIe-6363 board, which is negligibly small. In terms of self-irradiance, these
96

uncertainties translate to the values listed in Table IV.5. However, such large errors
have not been observed in plots of the measured ISC, which are nearly perfectly fitted
with a linear irradiance dependency of type ISC=kSCG, even at low irradiances.
Therefore, it is concluded that the instrumental uncertainties of the self-referenced
irradiances are much smaller than the corresponding values in Table IV.5.

Figure 4.32: Short-circuit currents of the nine modules corrected to 25°C. Data are
from a clear-sky day (21 June, 2011).

Plots of the nine modules’ ISC corrected to 25°C have been created for several
clear-sky days. Their slopes are slightly reduced as expected, but the hysteresis seen
before the temperature correction is still present (see Figs. 4.32 and 4.26). The
difference between morning and afternoon outputs is probably caused by angular effects
due to misalignment of some modules with respect to the mc-Si irradiance sensor. At
flat angles of incidence, a few degrees change in AOI affects the reflectance loss a lot
(see Fig. 4.28). The change of the cosine loss (1–cos(AOI)) equals +∆AOIsin(AOI) and
is also much higher at AOI closer to 90°. For the module with the widest hysteresis in
the plot of ISC over G (the one denoted ‘ESS 2’), the ratio of the afternoon output to that
in the morning is given in Fig. 4.33. The relative difference increases strongly at low
irradiances, which are associated with flat AOI and significant reflectance loss for
97

direct sunlight, and drops abruptly with the onset of pure diffuse-light conditions when
AOI>90° (i.e., when the sun is behind the plane of the modules). This observation
supports the misalignment hypothesis, thus suggesting that the module’s azimuthal
orientation is a few degrees to the west from that of the mc-Si irradiance sensor. In
March 2012, the two modules based on ESS™ were replaced by devices of the same
size and make which were so far tested in Kristiansand. This allowed the author to
investigate whether the observed hysteresis was module-specific or was due to
imperfect alignment of the module. Eventually, the second explanation was confirmed
when the most pronounced hysteresis in the ISC-G plots was obtained for the module
that replaced ‘ESS 2’.

Figure 4.33: Ratio of the afternoon and the morning current output (corrected to
25°C) of the module denoted ‘ESS 2’. Its data represented in Fig. 4.32 are used.
98

4.8 SOLAR AND CLOUD RESOURCES MEASURED IN THE YEAR 2011

4.8.1 SEASONAL DISTRIBUTION OF TILTED IRRADIATION IN 2011

Figure 4.34: Monthly irradiations in the plane of PV modules measured in 2011


compared to results from 2005 presented in [Mid10] and the PVGIS estimates.

The monthly solar irradiations and the annual irradiation in year 2011 estimated
from 1-minute pyranometer data are given in Table IV.6, together with the long-term
prediction from [PVGIS]. Data corrupted due to snow and ice cover, as well as by
specular reflections, have been corrected using the median self-irradiance of nine PV
modules, as described in the previous Section. Irradiance levels below 30 W/m2 have
been disregarded during the integration. The total irradiation obtained in the present
work exceeds the long-term estimate of the PVGIS tool by 15 %. This is a remarkable
difference, but one should take into account the even larger year-to-year variability of
the local solar resource discussed earlier in Section 4.4. However, the analysis of local
meteorological records covering 15 consecutive years has revealed that the PVGIS tool
underestimates the local horizontal irradiation by about 10 %. Also, the horizontal
irradiation in 2011 was below the average for the past 15 years. These findings indicate
that the solar resource available in the present geographical region is indeed strongly
underestimated by the solar radiation database used in [PVGIS].
99

Table IV.6: Monthly tilted irradiations (in kWh/m2) in year 2011 from the present test
configuration in Grimstad versus long-term predictions from [PVGIS].
Month Present Work PVGIS
January 48.1 20.8
February 28.0 42.9
March 118 79.8
April 166 119
May 156 152
June 169 156
July 149 154
August 145 127
September 101 92.5
October 69.4 54.8
November 26.8 25.8
December 23.2 15.2
TOTAL 1200 1040

Figure 4.35: Differences between the monthly irradiations measured with the mc-
Si PV sensor and the thermopile pyranometer.

The two seasonal distributions in Table IV.6 are represented in graphical form in
Fig. 4.34, along with the distribution for year 2005 obtained in [Mid10] for a tilt angle
of 50°. Due to the significant difference of the two experimental setups, only a
qualitative comparison is possible between the distributions obtained for years 2005 and
2011. Nevertheless, in both of them the irradiations in January, March and April are
much higher compared to the corresponding ones estimated from [PVGIS]. The setup
used in 2005 had a larger tilt angle, and therefore the spring irradiation was emphasized,
100

whereas the summer resource was suppressed due to high solar elevations and the
associated higher cosine losses. The lower tilt angle used in the present test setup
emphasized the summer irradiation as compared to that in the spring. Despite this, high
irradiation was obtained in March and April also in 2011. This similarity between 2005
and 2011 seems to had been a coincidence, because the horizontal irradiations in the
two spring months were among the highest when considering their variability from 1997
through 2011 (see Fig. 4.3). The study of the local solar irradiation initiated by the
present work should continue for at least seven more years in order to determine the
long-term annual average with less than ±5 % uncertainty [Ham12].

Figure 4.36: Irradiance profiles recorded by the two sensors on 01 January and 21
November, 2011.

The annual irradiation registered by the PV-based sensor is 3.6 % below that of
the pyranometer. However, only part of the difference can be attributed to reflectance
losses at flat angles of incidence. Such losses are expected to be strongest for the
summer months, during which the sun is available from a much wider azimuthal range
than in the other seasons. However, the difference between the monthly irradiation
estimates from the two sensors peaks in November, reaching almost 8 % (see Fig. 4.35).
Indeed, there has been a hint for some degradation of the responsivity of the PV-based
sensor in the course of the year. In the preceding Section, the readings of the two
irradiance sensors are compared against the self-referenced irradiances of nine of the
tested PV modules, and correction factors are derived for each device at AM1.5
conditions. The correction factors of the pyranometer and most of the modules are
practically unchanged in August as compared to April; moreover, almost no correction
is needed for the thermopile sensor. On the other hand, the mc-Si PV sensor had to be
101

corrected by +1.1 % as of April and by +2.7 % as of August. Analysis of irradiance


profiles recorded throughout the year also reveals a gradually changing bias of the PV-
based sensor with respect to the pyranometer around noon, when the reflectance losses
are minimal (see Fig. 4.36). The bias has been positive in January, but has become
negative afterwards, peaking in November and then decreasing slightly in December.

Figure 4.37: Irradiance profiles recorded on 27 November, 2011. Note the artifact
in the reading of the PV-based sensor between 4 and 5 p.m.

This decreasing responsivity of the mc-Si sensor greatly exceeds the typical rate
of aging of PV modules based on the same material. It cannot be attributed to initial
light-induced degradation (LID) either, because the latter occurs on much shorter time
scales. Also, the long-term degradation of PV devices in the field is usually not
reversible, whereas in the present case some recovery of the sensor’s responsivity took
place. Eventually, the hardware and the software that are behind the mc-Si PV cell in
the data acquisition chain have been identified as the main suspects for this behavior.
The irradiance profiles recorded by both sensors on 27 November are given in Fig. 4.37,
in which an artifact is seen in the mc-Si sensor’s reading well after the sunset. The data
acquisition system has attempted to record I-V curves of the PV modules at twilight
conditions, and thus also the sensor’s signal can be investigated (see Fig. 4.38). The
typical magnitude of the 50-Hz noise contained therein (at the amplifier’s output, which
102

is connected to an analog input of the PC), is within 1 W/m2 (see Fig. 3.16). However,
in the abnormally high signal recorded on 27 November, this mode has a far greater
magnitude, thus suggesting malfunction of either the amplifier or the NI data acquisition
board. The input and output ranges of the former are programmed on firmware level
(and not with the onboard DIP switches), and software is prone to errors. This issue
deserves further investigation and should be addressed in future. If the amplifier has
caused the abnormal data, this would imply that the EMI resulting in the 50-Hz noise
occurs inside the device, and not in the wiring between its output and the PC input.

Figure 4.38: The signal of the mc-Si sensor corresponding to one of the anomalous
data points in Fig. 4.37.

4.8.2 STATISTICS OF IRRADIANCE VALUES


The statistical distribution of irradiance values is important when the AC power
delivered by a PV system is investigated [Zhu11], and forms the basis for defining
standardized efficiencies of PV inverters. The European and the Californian efficiency
are calculated by weighting the inverter’s efficiency at different input power levels
[Hae01, Ble08]. The weighting factors used in the two geographical locations generally
reflect the local irradiance distributions, and thus the different climates. Indeed, the
power from a PV array does not depend linearly on the in-plane irradiance, partly due to
the involved temperature effects. Also, inverters and module arrays in a PV system are
103

rarely 100 % matched in terms of rated power. Nevertheless, irradiance distributions


represent local climates, which are expected to differ between e.g. Greece and Norway.
Consequently, the overall performance of the same PV inverter may differ significantly
across Europe. In this context, the annual and monthly irradiance statistics presented
further in this Section provide valuable information about the local climate in Southern
Norway. The relative energy contribution of the overirradiance events is also seen in
these plots.

Figure 4.39: Probability distribution of solar irradiance values recorded in 2011.

Figure 4.39 shows the probability of occurrence of different irradiance levels in


2011 as measured with the two sensors with 1-minute resolution. The two distributions
are practically identical except at the lowest irradiances, where reflectance losses affect
the mc-Si cell most significantly, and for irradiances around 1000 W/m2, at which cloud
enhancement is more probable and the slow response of the pyranometer often leads to
erroneous data. Also, the gradually decreasing responsivity of the PV-based sensor has
biased its readings by a few per cent towards lower values, which has affected its
distribution mainly at the high irradiances. In general, the shape looks quite similar to
the one presented in [Zhu11] for the GHI at Loughborough, UK, which is expected
because both locations are at high latitudes and have maritime climates. Therefore, the
inverter sizing of future PV systems in Southern Norway should not follow blindly any
104

guidelines developed for continental climates, but instead, optimization study should be
carried out and local guidelines should be developed. The high-resolution irradiance and
module temperature data obtained with the present experimental setup can serve as a
basis for such a study.

Figure 4.40: Energy distribution of the irradiance values recorded in 2011.

The distribution of the relative energy contributions of the different irradiance


levels looks very different from the probability distribution (see Fig. 4.40). For example,
during almost 30 % of the daytime, the local irradiance does not exceed 100 W/m2. The
actual percentage is even higher than this, since the data represented in Fig. 4.39
exclude illumination levels below about 30 W/m2. Despite the high probability, these
low levels of illumination contributed less than 5 % of the annual tilted irradiation. On
the other hand, irradiances between 600 and 1000 W/m2 contributed half of the received
solar energy. Thus, eventual significant inverter undersizing of local PV systems may
result in less electricity production as compared to sizing ratios closer to unity, which
seem more appropriate for optimally oriented PV arrays [Zhu11]. However, the
recommended values should be obtained from detailed calculations accounting for the
physical and economical characteristics of the different system components, as in
[Zhu11, Che10]. Such analysis is beyond the scope of the thesis.
105

Energy distributions of the different irradiance levels have been plotted also for
each month in 2011. These can differ significantly from each other, even between two
consecutive months. For example, the distribution for April is very different from that
for May, with a much larger contribution of the high irradiances, which is attributed to
the lower probability of cloudy conditions (see Figs. 4.41 and 4.42). On that particular
year, January was quite sunny and its distribution looks somewhat similar to that for
April, but due to the low solar elevations it peaks between 500 and 700 W/m2.

Figure 4.41: Energy contribution of the irradiance values recorded in April, 2011.

The monthly distributions of measured irradiance values can be useful for


identification of cloud enhancement events and their peak magnitudes in a given month.
The next part of this Section presents a more detailed analysis of those events which
resulted in overirradiances (values well exceeding 1000 W/m2). For example, the
longest bar in Fig. 4.41 corresponds approximately to the peak irradiance level at clear-
sky conditions in April, whereas the relative contributions of the higher irradiances drop
exponentially with magnitude. This behavior corresponds well to the location-specific
probability distribution of overirradiances which is discussed in the next part, and to
similar distributions for other locations found in the literature [Han10]. Above 1000
W/m2, the distributions for all months from March through September are qualitatively
very similar to the corresponding part in the distribution for April. The distribution of
106

October peaks slightly above 800 W/m2 which is attributed to the lower solar elevation,
whereas cloud enhancement has caused bursts in excess of 1000 W/m2. However, cloud
enhancement can also happen at much lower irradiance levels, and does not necessarily
lead to overirradiances. For January and November, the distributions reveal bursts
barely reaching about 800 W/m2. On the other hand, no evidence of cloud enhancement
is evident in the distributions for February and December.

Figure 4.42: Energy contribution of the irradiance values recorded in May, 2011.

4.8.3 OVERIRRADIANCES: MAGNITUDES, DURATIONS AND ENERGY SHARE


The so called ‘solar constant’ is defined as the extraterrestrial solar irradiance at
the earth’s mean distance from the sun, normal to sun’s rays. Different values of this
constant can be found in the literature, ranging between 1353 and 1394 W/m2 [Liu60,
Tul76, And00, Luq03b, Nel03a, Bet04, Lea05, Sze07b]. The values measured by the
aerospace community varied between 1353 and 1372 W/m2 [Eme03]. The estimates
considered as the best are 1367 W/m2 and 1366.1 W/m2 recommended by the World
Radiation Center and the American Society for Testing and Materials (ASTM),
respectively [Kal09]. The extraterrestrial irradiance varies by about ±3.4 % around the
mean due to the variation of the sun-earth distance (closest on 3 January, furthest on 4
107

July) [Liu60, Kal09]. A net increase of solar brightness of barely 0.1 % was measured
during activity maxima associated with the ‘11-year’ solar cycles [Lea05].

Figure 4.43, Top: interaction of sunlight with clouds (often called ‘silver lining’);
Bottom: strong forward scattering of sunlight in a thin cloud covering the sun.

The present work reveals the possibility for overirradiances well exceeding the
extraterrestrial values even at latitudes close to 60°N (at sea level), in good agreement
with prior studies and observations. A short burst exceeding 1800 W/m2 was observed
108

in May, 2011 in Kisumu, Kenya (at altitude 1131 m a.s.l., very close to the Equator),
whereas the peak clear-sky irradiance was about 1200 W/m2 [Bus11]. The specific
conditions involved clouds surrounding the sun. Similar observations were made in the
present research for the overirradiance events in Grimstad (as can be deduced from Fig.
4.7). Such cloud enhancement events were recently reported in [Zeh10, Zeh11] as
“unexpectedly high irradiances” due to “cloud reflection at the sharply defined edges of
cumulus clouds”. However, the present author does not agree with the offered
explanation. He would rather attribute these overirradiances to strong forward scattering
(Mie scattering) within a narrow angle inside the clouds, not necessarily only at the
edges. The Mie phase function has a strong peak in the forward direction where 51 % of
all scattered photons are concentrated within just 5° [Bou06]. It is responsible for the
silver lining which is sometimes seen at overcast conditions. For example, in the upper
photo in Fig. 4.43 the sun is covered by a cloud. Strong light is emitted by a narrow
region near the upper right cloud edge, which cannot be attributed to reflection since the
sun is behind the cloud. ‘Refraction’ (by means of forward scattering) is a more
plausible description. This is confirmed by the lower photo, where the sun is covered by
a much thinner cloud transmitting light over a much greater and continuous part of its
area.
A situation is possible, in which the unobstructed sun appears (to an irradiance
sensor) in a narrow gap between clouds which are thin enough within 5° or less around
the solar disk, and act a bit like a magnifying lens. The result would be a natural small
concentration of sunlight. On a partially cloudy day with many clouds moved by the
wind, multiple overirradiance events would be possible, with magnitudes and durations
depending on the geometry and velocity of clouds. The present work shows that this is a
very probable situation in this coastal location of Southern Norway [Ime11].
Overirradiances have been largely underestimated in PV system sizing and
design. They are associated with low module temperatures, and therefore result in high
voltages. Consequently, the power conversion electronics connected to PV arrays (e.g.
inverters or battery chargers) experience bursts of high power. Their duration may
exceed 5 minutes in locations with low wind speeds [Zeh10]. For several locations in
Germany, the duration distribution of global horizontal irradiances exceeding 900 W/m2
was presented. These could last over 30 minutes in some places. That study used
meteorological records with high (1-second) temporal resolution. It suggests that in
coastal locations with relatively high wind speeds, overirradiance events would have
little contribution to the annual solar irradiation, which is confirmed by this thesis.
In the present work, a second, ‘real-time’ irradiance data set of 10-millisecond
resolution is being recorded for the purpose of detailed overirradiance studies. In this
way, important questions can be answered, such as what the peak irradiance values in
Grimstad are, how long they typically last, and what their cumulative energy is. This is
another valuable contribution to the PV community, because the detailed and accurate
statistics of irradiance values are important for the proper inverter sizing in a PV system
109

(the ratio of inverter’s and PV array’s STC rating) [Bur05]. Sizing based on hourly
averages hides the important peaks and leads to energy losses due to inverter
undersizing, usually by 30 % or more below the PV array in terms of nominal power.

Figure 4.44: The strongest overirradiance event recorded in 2011. Note the cut-off
level of the fast sensor and the lag of the slow sensor.

Cloud enhancement of the instantaneous solar irradiance has been introduced in


Section 4.5. This Section contains results from the statistical analysis performed on the
real-time irradiance data set recorded in 2011. The latter consists of time series of
simultaneous readings by the two sensors with a main time step of 10 ms. There are
gaps in this data set corresponding to I-V curve sweeps performed on the PV modules
roughly every minute. About 1.7 % of the relevant daytime is therefore not represented
in the real-time data set. The study of overirradiances (here defined as periods with in-
plane irradiance well over 1000 W/m2) is further complicated by two main issues
associated with the two sensors. These complications are illustrated by Fig. 4.44
showing the strongest irradiance burst recorded in 2011. The amplifier of the mc-Si PV
cell limited the reading up to about 1354 W/m2. This proved insufficient for real-time
study of the highest possible overirradiances. The annual maximum of 1413 W/m2 was
therefore recorded with the pyranometer. Because its response time is several orders of
110

Figure 4.45: Probability distribution (top) and relative energy contribution


(bottom) of overirradiances with different magnitudes.

magnitude longer, this value in fact underestimates the actual peak level. Even after
accounting for the calibration uncertainties of the sensors, it is still surprising that extra-
terrestrial irradiances can occur near the sea level at such high latitude. Manufacturers
of irradiance sensors seem to be aware of this fact, which explains the upper range limit
of 1600 W/m2 of the thermopile instrument. However, even this range would be
insufficient to register the peak overirradiances near the Equator [Bus11]. Furthermore,
the slow response of thermopile pyranometers makes them inappropriate for such
111

studies, because many bursts can last shorter than the instrument’s response time. On
the other hand, PV-based and photodiode-based sensors give very high level of detail on
sub-second time scales (in combination with a fast enough data acquisition system).
The provider of the fast sensor used in the present work warns the user for such
events in the calibration protocol, as well as in the online data sheet, and attributes these
to forward scattering in cumulus clouds [Sol10, Sol12]. However, he expects
enhancement (with respect to STC) by no more than 20 % according to the data sheet,
and up to about 30 % according to the calibration protocol. The calibration has been
done in outdoor conditions in Silkeborg, Denmark (at approximate latitude 56°10’N). In
the present work, peak enhancement of about 50 % was registered in Southern Norway
(at latitude 58°20’N) in year 2012, after re-programming the sensor’s amplifier.

TABLE IV.7: The monthly overirradiance maxima registered by each irradiance sensor
in 2011. The upper cut-off for the mc-Si PV cell is indicated in blue. Corrections of the
cell’s sensitivity are not applied. The maximum annual record is shown in red.
Peak reading, mc-Si cell Peak reading, pyranometer
Month, 2011 2
[W/m ] [W/m2]
January - -
February - -
March 1293 1202
April 1354 1368
May 1354 1413
June 1354 1357
July 1325 1341
August 1331 1335
September 1354 1390
October 1057 1067
November - -
December - -

The real-time data set has been condensed by disregarding all the readings of the
fast sensor which are below 1000 W/m2. However, the resulting subset still contains
slow sensor readings below the standard irradiance level because of its long response
time. The remaining data are analyzed with respect to the maximum monthly
irradiances, the durations of overirradiances exceeding 1100 W/m2 (10 % or more above
the maximum clear-sky level), and their contribution to the annual irradiation. The
monthly maxima recorded by each sensor in year 2011 are listed in Table IV.7, in which
the cut-off of the fast sensor is marked with a blue color. Values of about 1300 W/m2
and even higher are possible during most of the year, excluding the winter months from
November through February. Contrary to the intuitive expectations, the two strongest
112

monthly maxima did not occur in June or July, but during May and September. This
observation should of course be associated only with that particular year, and should not
be taken as a general rule due to the inherently (from cloud cover) stochastic nature of
overirradiance events.

Figure 4.46: Probability distribution of overirradiance durations (top) and


frequency distribution of the longest events (bottom; note the limitation of the
vertical axis which cuts off the three left-most bars).
113

The probability distribution and the relative energy contribution of the different
overirradiance levels are shown in Fig. 4.45 and are practically identical. The
cumulative energy contribution of irradiances above 1000 W/m2 in year 2011 was 50
kWh/m2, or about 4 % of the annual irradiation. Only a fifth of this amount was due to
events exceeding 1100 W/m2, which numbered about 13’000 in year 2011. Only 100 of
these lasted above 1 minute, whereas the vast majority (96.4 %) lasted less than 15
seconds (see Fig. 4.46). The probability for an overirradiance exceeding 1100 W/m2 to
last more than 2 minutes was very small, yet individual events with duration of up to 6.5
minutes were found possible. These results correspond well to the coastal location of the
present experimental setup in Grimstad. Coastal locations are characterized with higher
wind speeds. Consequently, the clouds move quickly, and the resulting overirradiance
events last shorter compared to those occurring inland [Zeh10].

Figure 4.47: A very strong overirradiance recorded on 11 May, 2012.

At the present test site, measurements with high temporal resolution continue
using a fast irradiance sensor of range made wide enough to accurately characterize the
maximum possible cloud enhancement. Fig. 4.47 shows a burst recorded in May 2012
reaching 1521 W/m2, preceded by another peak of about 1350 W/m2. A burst of 1528
W/m2 has been registered later in June. This is the maximum overirradiance measured
by the time of the thesis finishing. Similar and much higher values have been reported
114

in the literature. Bursts exceeding the clear-sky GHI were seen in profiles of 1-minute
resolution presented in [Sue88]. Clearness indices greater than 1.0 were attributed to
cloud reflection, and the extra irradiance due to cloud enhancement was classified as
diffuse. However, this is not the case when strong forward scattering occurs within 5°
around the solar disk, because this portion of the sky is included in the direct (beam)
irradiance [Eme03, Kin97c, Mye02]. Multiple bursts were seen in a daily GHI profile
from February recorded with 1-minute resolution in Townsville, Australia [Wal01]. The
peak value exceeded 1300 W/m2. Overirradiances of up to 1500 W/m2 (in terms of
GHI) were recorded in June, 2009 in Albuquerque, New Mexico [Han10]. The high
values were attributed to reflection from clouds, following conclusions in earlier works.
Bursts exceeding 1500 W/m2 are seen in the presented tilted irradiance profiles of 1-
second resolution from San Diego, California [Luo12]. The maximum possible AC
energy production of a PV system required an inverter sizing ratio of 1.22. It was
concluded that losses during cloud enhancement events could only be quantified by
using 10-second or finer resolution.
GHI values exceeding 1800 W/m2 were registered in the equatorial Andes
[Emc08]. These imposed revision of the instrumentation and the database. The authors
termed the recorded excessive values “superirradiance”. The maximum enhancement
factors (with respect to clear-sky irradiance) reached 170 %. A photograph was given
which showed “refraction” of sunlight by clouds surrounding the sun. That study may
have missed many short-lived bursts, because the temporal resolution (5 min) was rather
low for focused studies of the phenomenon, whereas the sensors’ response time was too
long (18 s). A very detailed analysis of the phenomenon was implemented. Empirical
maxima were shown to scale linearly with altitude. The authors concluded that bursts as
high as 2000 W/m2 could be expected at altitudes of about 5 km a.s.l. A correlation was
found between the monthly overirradiance maxima and the modeled site-specific clear-
sky irradiance. The relative enhancement factor was shown to be practically
independent on the solar elevation. However, the absolute maxima were recorded at
elevations of about 80°. The phenomenon was attributed to diffuse reflection of
sunlight, as well as of ground-reflected light, from the clouds. Cloud enhancement was
found to occur on a daily basis. The authors believed that clouds scatter light
isotropically, despite the fact that the Mie phase function shows strong anisotropy and
has a high, narrow peak in the forward direction [Dei64, Bou06].
Wen et al. [Wen01] analyzed the possible cloud enhancement effects on
downwelling solar radiation in the gaps between cumulus clouds. Using a simplified
model with a plane parallel cloud and an overhead sun seen through a small gap (of size
equal to that of the solar disk, 0.5° in diameter), they obtained a maximum enhancement
of 1.82 times the clear-sky irradiance at an optimal cloud optical depth of 3. According
to the present author, a weak point of that model had been the assumption that the gap
could be neglected when integrating the diffuse irradiance scattered in the cloud. Only a
half of all photons undergoing single scattering would be contained within 5° from the
115

vertical [Bou06]. Therefore, adding cloud at the sun’s position (i.e., filling the gap)
would lead to some underestimation of the actual enhancement.
In a recent work, a fractal model was used to generate a synthetic cloud shadow
pattern [Ham12]. The obtained time series of global horizontal irradiance with 1-second
resolution did not contain any overirradiance peaks. Obviously, the cloud enhancement
effect was not included in the simulation model. There are many grid-connected small-
area PV systems in Germany. Strong, long-lasting irradiance bursts may affect the AC
output of such systems, as well as the operating mode of their inverters. It is therefore
advisable that all synthetic irradiance data should include cloud enhancement events of
various durations, with peak magnitudes of up to 1.5-1.6 times the corresponding clear-
sky irradiance. In this way, the simulations performed with these data would become
much more realistic. The present work provides a good starting point for such modeling.
The equation for maximum global irradiance suggested in [Hoy08] for quality
control of ground measurements leads to an upper limit of 2150 W/m2 for overhead sun
conditions (applicable to e.g. equatorial locations), but gives only about 1450 W/m2 for
Southern Norway:
W
G MAX  G O 1.5  cos  Z   100 2
1.2
(3.3)
m
GO is the normal irradiance at the top of the atmosphere, whereas θZ is the solar zenith
angle. The limits specified above are calculated assuming that GO equals the solar
constant, and seem reasonable in the context of the overirradiance results from the
present work and the value reported in [Bus11] for Kisumu, Kenya. However, the direct
irradiance upper limit given in [Hoy08] equals the extraterrestrial level GO which may
appear too conservative even for Southern Norway. For overhead-sun conditions
occurring near the Equator, the direct irradiance limit is definitely too conservative. A
more physical upper limit is therefore desirable, one which quantifies the magnitude of
the strongest overirradiance events possible in a given location. Their theoretical
quantification or numerical simulation is beyond the scope of the thesis. Such
significant cloud enhancement seems possible only for a very narrow angle around the
solar disk (within a few degrees), and so it is part of the direct irradiance component.
A major fault possible in PV modules is arcing, which may lead to fires and loss
of real estate [Rei12, Woh12, Joh12a]. Series arcing in c-Si modules occurs at weak or
broken parts of the front metallization, often at interconnection points between adjacent
cells or at microcracks in the wafer due to mechanical stress. Stress-testing of such
modules involved running a forward current of up to two times the rated ISC [Rei12].
From the results obtained in the present work, as well as from the literature review on
overirradiances given above, it becomes clear that similar stressing conditions can occur
naturally anywhere at latitudes within 60° and maybe even beyond that range.
Therefore, the use of arc-detection equipment is highly recommended in all PV systems
116

with voltage ratings above a certain (rather low) level. Such provisions have been
included in e.g. Article 690.11 of the American National Electric Code [Joh12b].
Apart from the PV context of the thesis, extreme overirradiances may have so
far unsuspected but quite significant impacts on other areas of human life, e.g. they may
have acted as fire starters in arid regions covered with dry vegetation. Therefore, the
latter hypothesis should be investigated carefully by the relevant research communities.
Damaging effects of cloud-enhanced UV on living organisms such as melanomas and
cataracts were discussed in [Par04a, Par04b, Tur06]. Overirradiance conditions can last
many minutes, and therefore any calculation of UV doses and the UV index should
account for cloud enhancement events. Regions in the southern hemisphere such as
Australia, Brazil, Chile and South Africa have a potential for particularly strong
overirradiances, because a combination of low air mass (i.e. overhead sun) and shortest
sun-earth distance occurs there in the summer.
To conclude, overirradiance events in this part of Southern Norway are quite
frequent, which can be explained by the high probability of partially cloudy conditions.
Their contribution to the annual irradiation is very small, and is negligible for
overirradiances exceeding 1100 W/m2. The majority of the latter typically lasts below
15 seconds. Thus, the expectation of short-lived overirradiances due to higher wind
speeds in coastal locations, which is expressed in Section 4.5, is confirmed. It is of little
importance whether the MPPT algorithms implemented in the present commercially
available PV inverters can adjust faster than this. However, individual peaks can last for
several minutes and may exceed 1500 W/m2, thus putting some robustness requirements
on PV inverters and range requirements on irradiance sensors. Given the high latitude of
the present test location, such requirements are generally valid for all locations where
PV generation is relevant. Even higher overirradiance values can be expected at the
lower latitudes, exceeding 2000 W/m2 for high-altitude locations near the Equator. Such
possibilities should be taken into account by designers of PV inverters and systems, as
well as by PV module manufacturers. Future studies of overirradiance events would
require a photodiode- or PV-based sensor with a response time in the order of tens of
milliseconds or less, and with a sufficient range depending on the latitude and the sensor
orientation. Designers of PV inverters and systems must take into consideration realistic
values of the possible input currents, voltages and powers due to peak overirradiance
events. A safety issue associated with the strongest bursts is the possible initiation of
series arcing in c-Si modules with faults in the front metallization or with cracked cells.
For Southern Norway, further work is necessary in order to fully analyze statistically the
durations and the magnitudes of such events. The present work provided gigabytes of
irradiance data with 10-millisecond resolution, which are a good starting point for such
analysis. However, the range of the PV-based sensor was insufficient by a factor of
about 1.3 until 9 March 2012, when its amplifier was re-programmed. With the
increased range (of up to 1600 W/m2), the study of overirradiances can be repeated for
year 2012 with a real-time data set of significantly improved quality.
117

4.8.4 OPTIMAL AZIMUTHAL ORIENTATION OF A PV ARRAY


In some geographic locations, systematic occurrence of cloudy conditions during
a particular part of the day (e.g., rainy weather in the afternoons) can result in an
optimal azimuthal orientation of a PV array which is different from the south. For the
present test site, such possibility has been investigated by averaging all daily irradiance
profiles from year 2011 (see Fig. 4.48). The resulting shape is quite symmetric with
respect to noon, suggesting a relatively uniform distribution of cloudy weather over all
daytime hours. The present azimuthal orientation of the tested PV modules (about 7°
east from south) belongs to a range of optimal angles around the south resulting in about
the same annual yield from an optimally inclined PV array [PVGIS]. This can be
explained with long-lasting high solar elevation angles combined with a wide range of
solar azimuthal angles in the months with highest contributions to the annual irradiation.
The ‘center of mass’ of the profile in Fig. 4.48 is at 13:05 p.m., which is after the
average solar noon at the given location (see Fig. 3.6). This suggests an optimum angle
of about 10-15° west from south. The exact evaluation is outside the scope of the thesis.

Figure 4.48: Averaged daily irradiance profile for year 2011.

4.8.5 SOLAR ENERGY LOST DUE TO CLOUDS


A question that is worth answering is: How much solar energy was lost due to
clouds in the year 2011? In other words, what would be the annual irradiation in the
plane of the PV modules in a hypothetic completely cloud-free year? The hypothetic
cloud-free irradiation is supposed to be relatively invariant over the years (varying only
slightly due to changes in the atmospheric water vapor concentration, aerosol content
and stratification profile). What varies between different years is the cloud cover in a
given location, and its analysis allows one to look at the annual irradiation from a
118

different perspective. In this way, X % year-to-year variation of the annual irradiation


can be translated to Y % year-to-year variation of the occurrence of cloudy conditions.

Figure 4.49: Curvilinear fitting of the irradiance profile of a mostly clear-sky day.

Answering the above questions is possible by analyzing the irradiance data from
2011 obtained in the present work. From the recorded daily irradiance profiles, one may
be able to model the profiles for all days in a hypothetic cloud-free year. Since a rough
first approximation is sought here, a rather simple tilted irradiance model would be
preferable – one involving only the angle of incidence AOI and the air mass AM as
main time-dependent variables. In addition, the small variation of the extraterrestrial
irradiance throughout the year can easily be modeled with a sine function. A key
challenge in such modeling would be the separate description of direct normal (beam)
and diffuse irradiance on a tilted surface. Simplification is possible by assuming that the
diffuse component GD is a small, fixed percentage of the beam component GB, which
depends on the extraterrestrial irradiance GO, as well as on AM. The total irradiance on
a tilted plane equals [Kin97c]:

G  G B cos  AOI   G D (3.4)


119

Figure 4.50: Fitting of the clear-sky daily irradiation versus the day of the year.

However, an even simpler approach is taken in the present work, because only a
rough first approximation to the problem is targeted. The daily irradiance profiles from
25 completely or mostly clear-sky days from the years 2010, 2011 and 2012 are
subjected to piecewise polynomial fitting (named ‘Smoothing Spline’ in Matlab’s
curvilinear fitting toolbox). Pyranometer data from the 1-minute data set are used. This
approach avoids the complexity of classic functional fitting, which involves initial
guesses for the values of multiple parameters. Instead, a single smoothing parameter p
is supplied for each daily profile. Experience shows that ‘longer’ profiles (from spring
and summer days) require a lower parameter, whereas ‘shorter’ profiles from winter
days need a higher parameter. In general, values much smaller than unity are preferable
in order to obtain a smooth fit also for days which have not been completely cloud-free
(see Fig. 4.49). For spring and summer irradiance profiles, the value 0.05 is found to
work best, whereas several times higher values are needed for the winter profiles. The
proper smoothing parameter is modeled with the simple equation p = 0.05(888/L)2,
where L is the number of data points in a given irradiance profile, whereas 888 is the
number of data points (minutes) in the profile from 21 June 2011 (see Fig. 4.7).
The fit of each daily irradiance profile is then integrated over time to obtain the
corresponding clear-sky daily irradiation. The values obtained for the 25 representative
days are plotted over the number of the corresponding day of the year (the discrete
120

symbols in Fig. 4.50). It is assumed that the clear-sky daily irradiation varies smoothly
throughout the year, with a minimum on 21 December and a maximum on 21 June, and
that it is periodic in nature. Trigonometric functional fitting is indeed available in
Matlab™, but again the simpler ‘Smoothing Spline’ method is used (with an optimal p
= 0.0003). The optimal value of the smoothing parameter is chosen by the trial-and-
error approach. A higher value results in a less smooth fitted curve, whereas a lower
value results in less accurate fit at the lower irradiations. Another issue with this method
is that it does not give a periodic fitted curve if data from only one year are fitted. This
is solved by fitting three consecutive years while using the same data for each year (see
Fig. 4.51). The fitted curve from the intermediate year is then considered to be an
acceptable periodic fit to the data (the continuous line in Fig. 4.50). Finally, the fitted
curve is integrated over 365 days, which gives the sought annual irradiation on a
hypothetical cloud-free year. For the degree of accuracy of this greatly simplified
approach, the result is practically the same for a leap year, since a clear-sky leap day (29
February) would add less than 0.3 % more irradiation.

Figure 4.51: Curvilinear fitting in Matlab™ of the clear-sky daily irradiation.

Thus, on a hypothetical cloud-free year, the PV modules tested in Grimstad


would receive 2130 kWh/m2 of in-plane solar irradiation. This rough estimate is of little
practical significance for electricity generation with PV in Southern Norway. It is done
121

to satisfy the pure scientific curiosity by answering the questions, “How much is out
there?”, and “What is lost due to clouds?” In the context of the actual irradiation
received in 2011 (1200 kWh/m2), it becomes clear that nearly a half of the sunlight that
was generally available was blocked by the clouds. Therefore, a maximum year-to-year
irradiation variability of, say, 20 % would translate as roughly the same variability in
the local ‘cloud resource’ (in terms of solar energy prevented from reaching ground
solely due to clouds, and not in terms of sky clearness index). It is worth noting that the
above assessment accounts not only for reduction of the clear-sky irradiance at overcast
conditions, but also for the enhancement of sunlight by clouds discussed in Section 4.5
and earlier in this Section. The derivation of a more accurate irradiation estimate for the
hypothetic cloud-free year is beyond the scope of the thesis. Future work may include
the application of the simplified approach presented above to the global horizontal
irradiance data available from Bioforsk [Bio12] analyzed in Section 4.4.
122

CHAPTER 5 TEMPERATURES
5.1 CHAPTER AIMS AND OBJECTIVES
The goal of this Chapter is to present the temperature measurement methods
used in the present work and to analyze their accuracy. The Chapter focuses mainly on
module temperatures, but ambient temperature results from Grimstad from the year
2011 are also considered. The operating temperature of a PV device has a very strong
influence on its power output. Therefore, the correct characterization and modeling of a
PV module’s performance depends on the accurate determination of this parameter. The
Chapter begins with a review of the best practices used to sense a PV module’s
operating temperature and the associated uncertainties. Then, the back-of-module and
ambient temperature measurements done in the present work are described. Later on,
the accuracy of the measured back temperatures is investigated by comparing these to
equivalent cell temperatures (ECT) calculated from module’s open-circuit voltage. A
detailed error analysis of ECT is then presented, and the influence of each parameter’s
accuracy is analyzed. Thereafter, the implications of a variable ideality factor for
calculating ECT are discussed, and a new equation is derived for PV devices with
significant (linear) ideality factor dependency on irradiance and temperature. Results for
the temperature coefficients of the open-circuit voltages and for the thermal resistances
of the studied modules are then presented. The Chapter ends with a discussion of the
methods, the obtained results, and some open questions which remain to be answered.

5.2 BEST PRACTICES IN MODULE TEMPERATURE MEASUREMENTS


Module’s operating temperature is the second most relevant parameter (after
solar irradiance) that strongly affects the output power and the I-V curve. Temperature
sensor accuracy within ±1°C is usually required in the characterization of PV modules
[Jag09, IEC11]. However, much larger uncertainties in module’s temperature can be
expected due to its non-uniform distribution across the module’s area [IEC93, Mül09,
Got10]. Another source of significant uncertainty is the fact that direct access to the
cells is not available for a typical commercial module. The module’s temperature is
therefore measured at one of its surfaces, most often at the back. At high irradiance
levels and low wind speeds, the cells are typically much hotter than the ambient air
[Kin98b, Kra09]. Consequently, a temperature gradient takes place across the
encapsulation of the module, and its surfaces are cooler than the cells. One way to avoid
this is to have a temperature sensor incorporated within the module, in direct contact
with an individual cell [Pro08, Kra09, Mid10]. Such customized module designs are not
possible when comparing modules from different manufacturers purchased on the
global market. Moreover, temperature non-uniformity is not accounted for when using a
single probe. The average module temperature can be measured using multiple
thermocouples attached to the rear surface [Kin97a]. This procedure (using film
thermocouples) was recommended for modules operated at the MPP between I-V curve
123

measurements [Whi01]. At near STC irradiance, an average cell-to-back temperature


difference of 2.5±1°C was adopted by the latter authors.
When testing many modules simultaneously, the use of multiple temperature
sensors for each module is quite resource demanding. This is because both
thermocouples and Pt100 sensors are used together with active electronics for signal
amplification. The use of a single temperature sensor for each module is much more
practical, but information is thus lost about the non-uniformity. The latter was estimated
as 3°C for modules tested outdoors at the European Solar Test Installation (ESTI), with
an assumed standard deviation of ±1.73°C [Mül09]. A Pt100 sensor attached at the
center of the back surface of each device was used. Other leading research groups
preferred Pt100 sensors as well [Pro08, Got10]. Foiled sensors were shown to measure
more accurately than ones placed in a ceramic case [Pro08], whereas tubed Pt100
resulted in a severe underestimation [Kra09]. More generally, the total uncertainty in the
module’s effective cell temperature is 3-5°C (at STC) that is equivalent to rated power
uncertainty of 1.5-2.5 % for c-Si modules [Got10].
A potentially very accurate approach to estimating module’s averaged operating
temperature is based on the temperature dependency of the open-circuit voltage VOC
[IEC93, Kra09, IEC11]. The latter reacts instantaneously to changes in irradiance and
cell temperature [Pro08], whereas the back-of-module temperature is affected by
thermal transients (delays). The equivalent cell temperature (ECT) was defined by an
international standard as the uniform p-n junction temperature at which the measured
electrical output of the PV device would be reproduced [IEC93, IEC11]. The standard
offers a procedure for ECT determination (for non-shadowed PV devices) from the
measured VOC and the in-plane irradiance G. The 1st edition of the standard uses the
definition
1  G1 
ECT  T2  T1   VOC2  VOC1  DNS ln   (5.1)
  G2 

where D=nk(T2+273)/q is the “diode thermal voltage” at T2 (including the ideality


factor n); k is Bolzmann’s constant; q is the elementary charge; and NS is the number of
serially-connected cells in a module. Prior knowledge is required of the open-circuit
voltage VOC1 at reference testing conditions (a set of irradiance G1 and cell temperature
T1 which may or may not equal the ones defined at STC). In addition, the temperature
coefficient of VOC, β (in units V/K), and the “diode thermal voltage” D must be known
in advance. Since D is linearly related to T2, Eq. (5.1) is easily solved for T2.
Furthermore, the ratio of the short-circuit currents can be used instead of G1/G2, if the
short-circuit current of the PV device scales linearly with irradiance.
The IEC procedure assumes: 1) a linear dependency of VOC on temperature; 2)
that the device’s I-V curve is governed by the 1-exponential equation; 3) that it has a
124

constant n; 4) a constant β; 5) an Ohmic series resistance; and 6) a reasonably high


shunt resistance. The last assumption is considered valid for most of the commercially
available PV devices [IEC11]. The procedure is however not recommended for
irradiances below 0.2 kW/m2, where the temperature coefficient is said to drop rapidly.
In general, the involved assumptions need verification for each device type, given the
short innovation cycles of some PV technologies [Hul10]. The IEC procedure has been
verified experimentally for individual glass-glass and glass-plastic c-Si modules [Pro08,
Kra09], but extensive testing with a wide range of technologies seems to be missing in
the literature. One possible explanation is the prerequisite for acquisition of several
other IEC standards distributed on a commercial basis (two in the 1st edition and eight in
the recent 2nd edition).
The voltage temperature coefficients of “typical” flat-plate modules are said to
be constant within 5 % for irradiances between 0.1 and 1.0 kW/m2 [Kin97a]. However,
it is not clarified whether this generalization applies to β when expressed in units V/K as
in [IEC93] or to its value in K-1 (normalized with respect to VOC) used in the recently
updated formulae [IEC11]. However, this ambiguity is unimportant in the case of ECT
calculation after IEC. The procedure first translates VOC2 to the reference irradiance, G1,
by means of the logarithmic term in Eq. (5.1). Then, the voltage difference is translated
to temperature difference by dividing the former by the temperature coefficient. In this
way, β needs to be known only at the reference irradiance, G1. If the latter defaults to
the standard irradiance of 1 kW/m2, the behavior of β at other illumination levels is
unimportant. In the 2nd edition of the standard [IEC11], the mathematical approach is
basically the same, and therefore the same conclusion is equally valid for the updated
procedure. From the equations given therein it becomes clear that β is normalized with
respect to the reference voltage, VOC1 (defaulting to VOC,STC).
The Sandia module database lists constant diode ideality factors for the
represented c-Si modules [Pra10]. The module’s ideality factor n is one of the input
parameters (constants) used in the Sandia PV array performance model [Kin04]. The
present author showed in an earlier work that, for several types of c-Si modules, n can
vary to a different degree with irradiance and temperature [Yor11a]. Although explicit
dependencies were not specified, this finding (if independently confirmed) can possibly
make IEC 60904-5 obsolete, and may impose a major revision. Moreover, its 2011
edition uses confusing terminology and perhaps an erroneous equation (to be discussed
shortly). In the present work, the formulation given in the 1st edition of IEC 60904-5 is
preferred as a starting point for further developments of the ECT calculation procedure.
This choice is justified by the following reasons. First, the mathematical concept is
basically the same (namely, linearization of partial derivatives). Second, two different
parameters (a and D) are termed as thermal voltages in [IEC11], whereas from Eqs. (2-
5) therein it is clear that both quantities must be dimensionless. Third, there seems to be
a technical error in Eq. (5) in the 2nd edition, namely the ‘plus’ sign before the
125

logarithmic term. Unless a negative “thermal voltage D” is assumed in the cited IEC
60891, such formulation is physically incorrect. However, using opposite signs for a
and D makes little sense since both are termed (dimensionless) thermal voltages, thus
confusing the user even more. By the time of the thesis finishing, the present author has
not been contacted by IEC’s customer service department. Finally, the commercial
acquisition of all the linked IEC standards (eight for the last edition) is not considered
economically reasonable. In Section 5.4, the ECT estimation for eight c-Si modules of
different type and make installed in Grimstad, is described. A detailed error analysis is
presented with respect to the input parameters. Most importantly, a modified formula is
suggested for modules with irradiance- and/or temperature-dependent ideality factors.

5.3 BACK-OF-MODULE AND AMBIENT TEMPERATURES

Figure 5.1: Sensing of a module’s temperature at the back surface.

The back-of-module temperature TMOD of all studied modules is measured using


type-K thermocouples sealed in epoxy. Each thermocouple is attached at the center of
the module’s back surface by means of thermally conductive paste and three layers of
insulator tape (see Fig. 5.1). The middle of the corresponding cell in the module has
been targeted. A 10-meter cable links each thermocouple to its own transducer (a
Phoenix Contact MINI MCR-SL-TC-UI). The measuring range is between –50°C and
+100°C, corresponding to output voltages between 0 and 10 V (identical with the
configured input range of the I/O board’s analog inputs). Thus, the transmission error is
1.1 %. This amounts to 0.54°C at module temperature of 0°C, but rises to 1.1°C at a
temperature reading of 50°C, which is closer to the nominal operating cell temperature
126

(NOCT) for most PV module types [Luq03a, Kra09, Mid10, Kui11, Mul12]. The
transmission error for the ambient temperature sensor (a tubed 4-wire Pt100 probe
placed in a radiation shield, see Fig. 5.2) has a similar magnitude.

Figure 5.2: The ambient temperature sensor is placed inside a radiation shield.

The daytime ambient temperatures measured during most of year 2011 are
presented in Fig. 5.3. Data are unavailable after 11 December due to transmitter
malfunction caused by condensation of water in the outdoor junction box. The test
location in Grimstad is characterized by a rather narrow span of the air temperature of
about 35°, as well as by low ambient temperatures in the summer. The latter are
beneficial for the performance of PV modules, which is discussed in detail in Chapter 7.
The temperatures of most modules are being recorded immediately after each I-
V curve measurement (see Fig. 5.4). On each channel, 100 samples taken in 25 ms are
averaged in order to cancel the noise. Only the temperatures of one module and the
ambient are being recorded during the I-V sweep time, due to the limited bandwidth of
the I/O hardware. Similarly, averaging of multiple samples is applied. In addition, all
temperatures are recorded at night at 1-min intervals (see Fig. 5.5). This allows the
periodic checking of all temperature channels for bias errors (see Fig. 5.6). It is assumed
that, in the absence of sunlight, all modules are in thermal equilibrium with the ambient
air during most of the time. This assumption is supported by the observation that the
bias errors measured at night can be very stable over many hours, whereas the
thermocouples’ readings vary in a similar way. In general, thermal transients can be
expected due to changing weather conditions, radiative exchange with the sky (e.g.,
127

with moving clouds), etc. Phase change effects are also possible due to freezing,
melting, condensation or evaporation of water on the modules’ surfaces.

Figure 5.3: Evolution of the daytime ambient air temperature in year 2011.

The bias errors at night are not constant on the longer term. Differences
exceeding 4°C have been observed between the thermocouples’ readings. This indicates
that the instrumentation error of a module’s temperature can exceed ±2°C (assuming
that the mean of the nocturnal temperatures equals the true ambient temperature). Under
the same assumption, a bias of the ambient temperature of order ±1°C is considered
possible. The latter measurement can often contain some noise (short-term variations
with a peak-to-peak magnitude of about 0.5°C). Corrections between –1.3°C and
+1.0°C have been applied to the measured temperatures since early 2011 as an attempt
to cancel the bias errors. However, this is considered unsuccessful in the long term (see
Fig. 5.7).
128

Figure 5.4: Daytime temperatures measured on 3 March, 2011. The delayed


cooling of the ambient air in the evening is attributed to the thermal inertia
(capacitance) of the radiation shield (see Fig. 5.2).

Figure 5.5: Temperature measurements at night, 9-10 January, 2011.


129

Figure 5.6: Deviations of the nocturnal thermocouple readings from their mean, 9-
10 January, 2011. The deviation of the Pt100 ambient sensor is also shown.

Figure 5.7: Deviations of corrected nocturnal thermocouple readings from their


mean, 8-9 June, 2011. The deviation of the Pt100 ambient sensor is also shown.

5.4 EQUIVALENT CELL TEMPERATURE (ECT) VS. BACK TEMPERATURE


It is demonstrated in Section 5.3 that the instrumentation used in the present
experimental setup can add ±2°C to the module temperature uncertainty discussed in
Section 5.2. A combined uncertainty of 5-7°C is considered very unacceptable, given
130

the strong temperature dependency of the PV modules’ output powers. Therefore, ECT
estimation has been undertaken for the eight c-Si modules and for the CIGS module as
alternatives to their back temperatures. This Section gives details on the calculations
used and discusses the challenges faced. A detailed analysis is performed that links the
error of the calculated ECT to each of the input parameters. An advanced formula is
proposed for modules with variable ideality factors, in which case the standard IEC
procedure is not applicable. As discussed later in Section 5.5, different degrees of
ideality factor variation are revealed for most of the modules tested in Grimstad.
For two of the mc-Si modules studied, the open-circuit voltage at STC, VOC,STC,
and the temperature coefficient of VOC at standard irradiance, β, are available from an
independent, certified laboratory. ECT calculation is implemented for these two
modules in the LabVIEW code, in parallel with the estimation of I-V curve parameters.
For this purpose, Eq. (5.1) has been modified to the form

ECT  25C 
1

V
OC  VOC ,STC  DNS ln G  (5.2)

where the plane-of-module irradiance G is in units kW/m2 (or suns, 1 sun = 1 kW/m2).
Note the changed sign of the logarithmic term compared to Eq. (5.1).

Figure 5.8: Comparison between ECT calculated in LabVIEW and the back-of-
module temperature. The open symbols indicate the maximum difference for
various irradiance levels. Values below –16°C are not shown.

The “diode thermal voltage” D (including the diode ideality factor, as defined in
[IEC93]) is determined for each measured I-V curve with the method described in
131

[Yor10b, Yor12a] and discussed later in Chapter 6. However, the slight temperature
dependency of the module’s series resistance, RS, is not accounted for in LabVIEW, but
a fixed value is used. Figure 5.8 compares ECT to the measured back-of-module
temperature, TMOD, for one of these two mc-Si modules discussed so far. The plot
combines data from twelve mostly clear-sky days in March and April, 2011. Irradiance
is expressed in terms of module’s short-circuit current, ISC, which equals about 8 A at
standard irradiance of 1 kW/m2 under the standard solar spectrum. As of the other
module, the plot looks qualitatively very similar. The open symbols mark the maximum
difference between the two estimates at different irradiance levels. The maximal values
of ECT–TMOD have an almost linear dependency on irradiance above about 0.2 kW/m2
(i.e., for ISC > 1.6 A). At thermally steady conditions, a linear dependency can be
expected between ECT and TMOD of the form

ECT  TMOD  cG (5.3)

where G is the in-plane irradiance and c is proportional to the combined thermal


resistance of all the media between the thermocouple and the cell in front of it. Eq. (5.3)
has the same form as Eq. (3.1), which relates module’s operating temperature to the
ambient temperature and the irradiance in the plane of the module. A similar linear
relationship between cell temperature and back-of-module temperature is used in the
Sandia model description [Kin04], which gives a difference of 2-3°C at standard
irradiance. It should be noted that not the entire solar power density G incident on the
module flows through its back. First, there are nonzero reflectance losses even at a
normal angle of incidence [Mar01]. Second, some percentage of the power absorbed by
the semiconductor material is delivered to the load connected to each PV module.
Furthermore, there are gaps between the cells in the module, which may reflect most of
the incident light if they are white (which is usually the case in c-Si modules). More
importantly, part of the thermal energy dissipated by the cells flows via the front
encapsulation layer and the front glass. The corresponding fraction is expected to be
about 1/2 for a glass-glass module structure and much smaller for glass-plastic modules,
in which the media behind the cells are much thinner compared to those in the front.
The linear fitting of the open symbols in Fig. 5.8 with Eq. (5.3) results in a
vertical axis intercept of –0.7°C, whereas ECT should equal TMOD at zero irradiance.
This deviation can be attributed to an instrumentation error (bias) as discussed in
Section 5.3, but a part of it may be due to an error in the VOC,STC estimated by the
independent lab. After correcting the fitted line for the bias, a difference of 7.0°C
between ECT and TMOD is obtained at irradiance of 1 kW/m2. For the other module, the
bias at zero irradiance is –2.2°C, whereas ECT exceeds TMOD by as much as 9.8°C at
132

standard irradiance. Both discrepancies are too big compared to the typical ones found
in the literature [Kin04], and are discussed later in this Section.
The difference ECT–TMOD varies by a few degrees at all irradiance levels, with
a slightly larger spread at high irradiances. Possible causes include: changing weather
(e.g., wind speed and direction); thermal transients; varying temperature distribution
across the module’s surface; and instrumentation errors. For example, southern winds
will improve cells cooling through the front glass by reducing the corresponding
thermal resistance. At a given irradiance, this will result in less heat flow through the
media behind the cells, and so the temperature drop ECT–TMOD across the associated
thermal resistance will be reduced. However, the almost constant spread for all
irradiance values indicates that most of it is due to instrumentation error. The ±2°C
scatter seen in the nocturnal temperatures of the modules corresponds well to the spread
seen in Fig. 5.8.
The quick drop at the low irradiances may be attributed to ECT being ill-defined
below about 200 W/m2 [IEC93, IEC11]. A detailed error analysis of the calculated ECT
is done later in this Section, which traces this low-irradiance behavior to
underestimation of D. The method used for estimation of D is optimal for higher
irradiances, at which the ideality factor is practically constant over a broad range of
module currents [Yor12a].

Figure 5.9: Comparison of two ECT calculation approaches for a mc-Si module.

Two conclusions are drawn from the above analysis of ECT calculated in
LabVIEW for two of the mc-Si modules. First, the implemented measurement of the
back-of-module temperatures is not at all satisfactory. The associated errors are typical
133

for the worst methods studied in the literature [Kra09]. Second, ECT must be used in
the analysis of all the modules instead of the measured back-surface temperatures.
For the two mc-Si modules discussed so far, an alternative approach to ECT has
been tried out, which uses a fixed ideality factor n. The latter is determined offline in
Matlab™ from I-V curves recorded at irradiances of about 0.9 kW/m2 [Yor11a] using
the method described in [Yor10b, Yor12a]. Details about the refined method used in the
present work are given in Chapter 6. The values of n for these two modules show very
slight variation with irradiance. Details about the nature of the observed variations are
given later in Chapter 6. In the ECT estimation described here, the values at standard
irradiance are used. Slightly different β values are used, taken from the most up-to-date
report of the independent laboratory. A comparison of the two approaches is shown in
Fig. 5.9 for one of the modules (using the data in Fig. 5.8). Irradiance readings of the
mc-Si reference cell are used without sensitivity correction, as the focus here is on the
overall performance of both ECT calculation methods. The effects of irradiance’s and
other parameters’ uncertainty on the ECT accuracy will be discussed shortly.
Since the same TMOD data set was used in the two plots, all differences between
them should be attributed to the ECT calculation methods alone. The similar spreads
seen in both plots are attributed to the measurement error of TMOD. Consequently, linear
fitting with Eq. (5.3) must be done for all data points in each plot (except for the
obvious outliers), and not only for the maximum values at a given irradiance (the open
symbols in Fig. 5.8).
Both approaches give very similar results at the higher irradiances. The slight
discrepancy close to standard irradiance is attributed to the difference of about 4 %
relative between the values used for β. However, in terms of linearity, the second
approach approximates much better Eq. (5.3) at very low irradiances. Due to a
somewhat steeper slope, this approach underestimates ECT–TMOD at zero irradiance
even more. In order to identify the possible cause, an error analysis of ECT is
undertaken (described in the following paragraphs), which focuses in turn on each of the
input parameters VOC, G, n (in the first approach, D), and β. Apart from these, a low
value of module’s shunt resistance will strongly affect VOC (as well as VOC,STC)
resulting in additional error of ECT. In an earlier work, the author found that the shunt
resistance is in the order of several kΩ for modules of the same type (one of which was
based on SoG-Si from a metallurgical route) [Yor11a]. Therefore, this parameter is
considered unimportant to VOC (and ECT), in line with the assumption made in the
updated IEC standard [IEC11], and is not considered in the following error analysis.

5.4.1 ECT ERROR ANALYSIS


The error of ECT estimated in LabVIEW can be expressed by differentiating
Eq. (5.2) and then simplifying by substitution in the term proportional to Δβ:
134

 1 G 
ECT    ECT  25C   VOC  VOC ,STC  DNS  NS D ln G  (5.4)
  G 

The quantities Δβ, ΔVOC, ΔVOC,STC, ΔG and ΔD denote the deviations of the associated
input parameters from their true physical values. The contribution of one parameter’s
error to ΔECT can be assessed independently on the other parameters by setting their
errors to zero (i.e., by assuming them absolutely correct). This is equivalent to
performing a partial differentiation on Eq. (5.2) with respect to the parameter of interest.
Thus, an error in the measured VOC would result in the following deviation from
the true ECT:
1
ECTVOC  VOC (5.5)

For the two mc-Si modules considered so far, an overestimation of VOC by 0.1 V
(roughly 0.3 % of VOC,STC) will underestimate ECT by about 0.85°C. The errors have
opposite signs because β is always negative. For VOC,STC, an overestimation by 0.1 V
would overestimate ECT by 0.85°C (note the sign of the former in Eqs. (5.2) and (5.4)).
Both errors are independent on irradiance, thus adding a constant bias to ECT. This is
illustrated in Fig. 5.10. For a specific value of VOC,STC (0.2 V above the value provided
by the independent laboratory), the vertical-axis intercept of the linear fit to ECT–TMOD
turns to zero (for this particular module).

Figure 5.10: Effects of erroneous VOC,STC on the calculated ECT.

The error due to a wrong estimate of the in-plane irradiance G will be


135

1 G
ECTG   DNS (5.6)
 G

Thus, a 5 % overestimation of G (a somewhat high but possible error in the measured


instantaneous irradiance) will overestimate ECT by about 0.8°C at module temperature
of 25°C, for any irradiance level. This value will vary with module’s temperature
because of D. At the higher irradiances (where the ideality factor is found to be
relatively constant for the two mc-Si modules considered so far), D will vary mainly
due to temperature variations. Consequently, the error in ECT due to typical irradiance
measurement errors will vary weakly between 0.7-1.0°C for cell temperatures in the
range –20 to +100°C. However, increased ideality factors can be expected at very low
irradiances due to the I-V curve’s deviation from the classic 1-exponential model. For
similar modules studied in an earlier work [Yor10b], the ideality factors at open circuit
were found to increase by 20 % at irradiance of 30 W/m2. Since D is proportional to n,
ΔECTG will be similarly increased at very low irradiances. The methods used for local
ideality factor analysis (developed by the present author) are described later in Chapter
6. Their basics were recently reported in [Yor12a].
An erroneous estimate of the thermal voltage D will lead to an error

1
ECTD   NS D ln G (5.7)

At standard irradiance (one sun), the logarithm becomes zero, and so does ΔECTD. At
lower irradiances the logarithm will be negative, with largest significance at very low
irradiances. As β is always negative, ΔECTD and ΔD have equal signs. Furthermore,
significant errors in the estimation of D in LabVIEW can be expected at very low
irradiances. Therefore, an underestimation of D of up to 30 % seems a logical
explanation of the negative errors in ECT seen in Figs. 5.8 and 5.9 at the lowest
irradiance levels.
In the 2nd ECT approach, Eq. (5.1) is solved for T2≡ECT as in [IEC93] (since D
is temperature-dependent), whereas the conversion to absolute temperature units is done
using the more accurate value of 273.15 K corresponding to 0°C. The solution is
subjected to error analysis with respect to n. The detailed calculation is given in
Appendix C. The error in ECT due to an erroneous n reads

 N k  ECT  273.15 ln G 
ECTn   n  S  (5.8a)
 q  nkNS ln G 

At standard irradiance, the error will be zero due to the logarithm, independently on the
values of β, n, ECT and NS. At lower irradiances, the logarithm will be negative and
136

consequently both the numerator and the denominator will be negative as well (since β
is always negative). At such conditions, there will be a sign difference between the
errors in ECT and n. However, Eq. (5.8a) is difficult to analyze with respect to ln(G).
The analysis is much easier when using the following form (applicable for G≠1
kW/m2):
 
 ECT  273.15 
ECTn   n   (5.8b)
 n  q 
 NSk ln G 

Note that for irradiances below 1 kW/m2, the 2nd term in the denominator is always
positive. If the logarithm is close to zero (i.e., at irradiances close to the standard
irradiance), the denominator will be very large, thus making ΔECTn very small in
magnitude. (This estimate is also valid for irradiances slightly above 1 sun.) At much
lower irradiances, the smaller denominator will lead to a much larger error. The idea is
illustrated in Fig. 5.11, where the 2nd ECT approach is also applied with erroneous
ideality factors. The zero-irradiance limit of the bracketed term in Eq. (5.8b) equals
(ECT+273.15)/n, which is of order 102 K. This explains the behavior of the two
erroneous plots at very low irradiances.

Figure 5.11: Effect of erroneous n on the calculated ECT. The T MOD data set is not
changed.

An erroneous β can only affect ECT at cell temperatures different from the
standard temperature:
137


ECT    ECT  25C (5.9)

The error does not depend directly on irradiance. Depending on the environmental
conditions, ECT may reach 25°C at some irradiance level, in which case the error due
to Δβ will be zero. Example plots with erroneous β are given in Fig. 5.12.

Figure 5.12: Effect of erroneous β on the calculated ECT. The TMOD data set is not
changed.

5.4.2 ESTIMATING ECT OF MODULES WITH VARIABLE IDEALITY FACTORS


For the remaining six c-Si modules and the CIGS module tested in Grimstad, the
label value of VOC,STC given by the manufacturer is used, whereas its temperature
coefficient is determined from measured data as described later in Section 5.5. Real-
time ECT calculation in LabVIEW is not implemented, but the measured data are
processed off-line. The ideality factor of each module is determined from multiple I-V
curves, as described later in Chapter 6.
A key finding of the present work is that, for some c-Si modules, as well as for
CIGS modules, the ideality factor can vary with irradiance. Preliminary results for the c-
Si modules were published in [Yor11a]. In the course of a clear-sky day, module I-V
curves recorded at different irradiances and module temperatures have a different value
of the ideality factor. The observed variation is not equal for all modules. For the two
mc-Si modules discussed earlier, the ideality factor varies by up to 4 % for irradiances
between 0.6 and 1.0 kW/m2, generally increasing with irradiance. The same pattern is
seen also for most of the remaining c-Si modules. The Chinese-made mc-Si module
shows the largest variation (about 8 % for the same irradiance range, supposedly much
138

more with a lower limit of 0.1 kW/m2). However, the observed ideality factor profiles
are not smooth like the corresponding irradiance profile. This observation suggests also
the presence of a temperature dependency. It is not straight-forward to determine the
irradiance- and temperature-dependency separately, as the two parameters are usually
(but not perfectly) correlated in outdoor measurements. With the increase of the wind
speed, the operating temperature of a module would be reduced. The coastal and
elevated location of the test site logically results in ribbed profiles of the module
temperatures, even on the most clear-sky days. The analysis of the irradiance- and
temperature-dependencies of modules’ ideality factors is described later in Chapter 6.

Figure 5.13: ECT versus back-of-module temperature comparison for the c-Si
modules. Fixed ideality factors have been used for the ECT calculation. Data from
22 March 2011 have been used.

A variable ideality factor is a big challenge to the ECT evaluation, as the case is
not covered by the standard procedure [IEC93, IEC11]. As a first approximation, the
value of n at standard irradiance has been used for each module. The so obtained ECT
values have been checked against the corresponding back temperatures, as for the two
mc-Si modules analyzed earlier. Plots of data from 22 March, 2011 (a mostly clear-sky
day) for all c-Si modules are given in Fig. 5.13. At the lower irradiances, the plots of
three of the modules are significantly curved. In accordance with the error analysis
139

performed earlier (see Fig. 5.11), this behavior is attributed to an overestimated n used
in the ECT calculation. Indeed, the Chinese-made mc-Si module (denoted ‘RS poly’)
has shown the biggest ideality factor variation among all the c-Si modules. Its much
lower values of n at lower irradiances are therefore overestimated by calculating ECT
with the ideality factor’s value at standard irradiance. This is exactly what is seen in Fig.
5.13 for this module and for the module with back-contact cells (denoted ‘Sunc.’), in
line with the expectations stemming from the error analysis. For the aged mc-Si module
(denoted ‘GPV’), the curving indicates ECT overestimation due to underestimation of
n at very low irradiances. On the other hand, a much smaller variation of n is obtained
for this module at the higher irradiances, where its ECT–TMOD plot shows a good
linearity. The high n at very low irradiances suggested by the plot corresponds well to
the much lower fill factor (FF) values of this module at low light (see Fig. 6.1 in
Chapter 6).

Figure 5.14: ECT–TMOD of the module denoted ‘RS poly’ for constant and for
variable n. The label value of VOC,STC has been used in both plots.
140

Figure 5.15: Fitting ECT–TMOD vs. irradiance with Eq. (5.3) in the case of the
Chinese-made mc-Si module (‘RS poly’). The vertical-axis intercept is used to
determine the necessary correction of the label value of VOC,STC.

For the same three modules, the fixed-ideality-factor ECT is nearly equal to the
back temperature at the high irradiances. As discussed earlier, an erroneous n would
have very little effect on ECT close to standard irradiance. Furthermore, the used fixed
ideality factors are namely those at 1.0 kW/m2. On the other hand, the similar slopes of
all curves in Fig. 5.13 at the higher irradiances suggest that the coefficient c defined in
Eq. (5.3) has the same order of magnitude for all the modules. Consequently, a
reasonably high difference ECT–TMOD (at least 6-8°C) will take place near standard
irradiance for all of them, which is not seen in the plots of the three discussed above.
The small spreads seen in their plots for all irradiance levels indicate that β has been
determined rather accurately in the present work (see Figs. 5.12 and 5.13). The voltage
measurements (and the derived values of VOC) are very accurate as well. From the ECT
error analysis performed earlier, it eventually becomes clear that there are only two
possible causes for the too low values of ECT–TMOD near standard irradiance. The
possibility of large (6-8°C) positive biases in the measured back temperatures TMOD of
the three modules is ruled out, since such are not observed at night (see Fig. 5.5). The
only remaining cause possible is therefore a deviation of the label value of a module’s
141

VOC,STC (used in the ECT calculation) from its true physical value (see Fig. 5.10). The
resulting error in ECT can be evaluated by substituting –ΔVOC,STC for ΔVOC in Eq.
(5.5). In the case of the three modules considered, ECT is significantly underestimated,
meaning a negative ΔVOC,STC (i.e., the actual value of VOC,STC is higher than the label
value given by the manufacturer).
The deviations ΔVOC,STC are not known a priori, but the analysis of ECT–TMOD
with Eq. (5.3) has pointed at a path to its determination. First, the standard ECT
calculation procedure has been advanced to include modules with variable ideality
factors. The proposed solution will be described shortly. Depending on its success, the
plots of ECT–TMOD of all modules are expected become more or less linear (see Fig.
5.14) in accordance with Eq. (5.3). Then, a linear fit can be applied to each plot (see
Fig. 5.15), and then the vertical-axis intercept can be multiplied with the module’s β to
give the necessary correction of the label value of VOC,STC. Subsequently, the corrected
VOC,STC of each module can be used in all further ECT calculations, resulting in the
closest possible compliance with Eq. (5.3). Table IV.1 lists the values of the parameter c
obtained from the linear fits (equal to ECT–TMOD at G = 1 kW/m2), and the corrections
applied to VOC,STC for each module.
It should be noted that a biased TMOD measurement has the same effect on the
ECT–TMOD plot (a vertical shift) as does an erroneous VOC,STC (see Fig. 5.10).
However, the corrected values of the modules’ VOC,STC are estimated using multiple
daily data sets, and therefore the method averages out the random errors in TMOD (see
Fig. 5.15). Therefore, the present estimates of VOC,STC are considered quite accurate.

TABLE V.1: Estimates of the parameter c in Eq. (5.3) and the corrections applied to the
label value of VOC,STC for the purpose of accurate ECT calculation.
Module c [K/(kW/m2)] Correction of label VOC,STC [V]
Aged BC mono-Si (BP) 9.5 –0.23
Aged mc-Si (GPV) 7.1 +0.15
Back-contact mono-Si (Sunc.) 12.7 +0.67
RSS mono-Si (RS mono) 6.9 +0.02
RSS mc-Si (RS poly) 11.8 +1.24
HIT 9.8 +0.14
ESS1 mc-Si 7.4 +0.14
ESS2 mc-Si 8.6 +0.40
CIS 10.6 –1.42
142

Suggested ECT Calculation Procedure for Modules with Variable Ideality Factors
The last term in the brackets in Eq. (5.1) represents a translation of VOC from
irradiance G2 to irradiance G1 at a fixed module temperature. The quantity D therein is
proportional to the ideality factor n, which is assumed constant in [IEC93, IEC11]. The
term can be regarded as the integral between G2 and G1 of the partial derivative
∂VOC(G,ECT)/∂G. The latter can be expressed from the 1-exponential I-V curve model
applied at open-circuit conditions:
 V 
I  VOC   0  I SC  I 0  ECT  exp  OC  (5.10a)
 nNS VT 
I  k G
VOC  nNS VT ln  SC   nNS VT ln  SC  (5.10b)
 I0   I0 
The short-circuit current ISC in the latter equation is assumed to vary linearly with the
irradiance G (in line with [IEC11] and with the experimental results presented in
Chapter 4). For a constant n,
VOC I k SC nNS VT
 nNS VT 0  (5.11)
G k SCG I 0 G

Figure 5.16: Synthetic J-V curves (at T=25°C) of the module denoted ‘RS poly’.
143

Integration of the derivative between G2 and G1 gives

VOC G  G 
G1 G1
dG
G G  dG  nN S T 
V
G2
G
 nNS VT ln  1   DNS ln  1 
 G2   G2 
(5.12)
2

which is exactly the last term in the square brackets in Eq. (5.1).
From Eq. (5.10b) it is clear that for a constant n and a fixed device temperature,
the semi-logarithmic ISC-VOC curve is linear, with a constant slope equal to 1/(nNSVT).
Aberle et al. [Abe93a] argued that, if the saturation current and the ideality factor did
not depend on the device voltage, the ISC-VOC curve would be identical with the RS-
corrected I-V curve. The analysis of I-V curves of the modules studied in the present
work (presented later in Chapter 6) shows that n and I0 are constant at the higher
voltages, where the RS-corrected semi-logarithmic I-V plots are practically linear. As
the module temperature varies with irradiance in outdoor test setups, ISC-VOC curves at a
fixed module temperature cannot be obtained directly. However, n is found to vary with
irradiance. One would therefore assume a curvilinear semi-logarithmic ISC-VOC curve,
and examples of such curves were given in the literature [Abe93b]. In the present work,
it is assumed (partly following [Abe93a]) that, at a fixed device temperature, the slope
of the tangent to that curve equals 1/(nNSVT), see Fig. 5.16. Consequently, Eq. (5.11) is
also valid for a device with a variable n(G,ECT), but the integration of the derivative
over irradiance must differ from the one performed in Eq. (5.12).
The assumption of a curvilinear semi-logarithmic ISC-VOC curve may seem
incompatible with the observed linearity over a wide current range of the RS-corrected
semi-logarithmic I-V plots analyzed later in Chapter 6. This apparent contradiction is
ruled out by considering the major difference between the two types of curves. As
pointed out in [Abe93a], the excess minority carrier density (which defines their level of
injection) varies much slower along an illuminated (semi-logarithmic, ISC-shifted) I-V
curve than along an ISC-VOC curve. As already discussed in Chapter 1, injection-level
dependency of the Shockley-Read-Hall recombination in c-Si devices can result in
ideality factors exceeding the ‘raw’ theoretical value of 1. For a one-sun illuminated
ISC-shifted I-V curve, the excess carrier density at the MPP is only 3-5 times less than at
open circuit. However, the corresponding point in the ISC-VOC curve (in semi-
logarithmic axes) is obtained for, say, 40-times lower irradiance. At open circuit, the
level of injection is determined mainly by irradiance, and therefore the excess carrier
density decreases about 40 times in the ISC-VOC curve [Abe93a]. In thin-film devices,
depletion-region recombination is significant and the resulting ideality factor values will
generally differ from those typical for c-Si. Nevertheless, the same considerations are
considered equally valid for the CIGS module studied in the present work.
144

The ideality factor variation is modeled (as a simple first approximation) with a
linear relationship to the in-plane irradiance and back-of-module temperature given in
Eq. (5.13). The value n0 corresponds to standard irradiance G0 = 1000 W/m2 and
temperature T0 = 25°C. The derivation of the coefficients a and b for the different
modules is described later in Section 6.6.
n  G ,TMOD   n0  a  G  G 0   b  TMOD  T0  (5.13)
For the purpose of ECT calculations, Eq. (5.13) is substituted in Eq. (5.11),
which is then integrated over irradiance (at a fixed module temperature):
G1
VOC
G
1
n  G ,TMOD  dG
G G  dG  N S T 
V
G2
G
2

 G1
dG
G1

 NS VT  n 0  aG 0  b  TMOD  T0     a  dG  (5.14a)
 G2
G G2 
 G  
 NS VT  n 0  aG 0  b  TMOD  T0   ln  1   a  G 1  G 2  
  G2  
The above equation can be simplified by setting G1 equal to the standard irradiance G0,
and substituting G for G2 (all in units kW/m2, whereas a must be in (kW/m2)-1):


VOC  NS VT n0  a  b  TMOD  25C  ln G  a 1  G  (5.14b)
In the advanced ECT formulation suggested in the thesis, the above expression replaces
the last term in the brackets in Eq. (5.2), which is based on the standard procedure
[IEC93]. Note that VT is proportional to ECT+273.15 K and requires that the advanced
equation be solved for ECT, similar to what is done in the standard procedure [IEC93].
For some of the modules, the ECT evaluated with the new equation is
significantly overestimated at irradiances below about 0.1 kW/m2 (see Fig. 5.17). For
these modules, the ideality factor at very low irradiances seems to differ from the values
modeled with Eq. (5.13). The physical nature of this behavior is outside the scope of the
present discussion, which is focused on improving the accuracy of ECT calculated with
a variable n. For this purpose, the values obtained with the advanced equation are
corrected by adding an exponential term of the form
ECTLOW _ G   Aexp  B  G  g   (5.15)
where g ranges between 0.018 and 0.042 kW/m2 for the different modules. Correction is
applied for all the nine modules (including the CIGS). For the module represented in
Fig. 5.17, a value of A as high as 8.2°C is used, whereas B ranges between 5.6 and 67.5
(kW/m2)-1 for the different modules. The optimal values of the coefficients are obtained
by fitting the plot of ECT–TMOD–cG over the irradiance G with the negated Eq. (5.15).
Example is given in Fig. 5.18. The latter equation is preferred to Eq. (5.8b) as much
simpler, yet sufficient for most of the modules. However, for two of the modules a two-
145

exponential version of Eq. (5.15) must be applied, which is omitted for brevity. The two
modules are the Chinese-made mc-Si (denoted ‘RS poly’) and the HIT module. The
need for more complex corrections is attributed to a less accurate fitting of their variable
n with Eq. (5.13).

Figure 5.17: ECT obtained with a corrected VOC,STC vs. back temperature of the
aged mc-Si module denoted ‘GPV’.

As a result of the advanced formulation and the corrections described above, the
final values of ECT show very good compliance with Eq. (5.3) also at the lowest
irradiances (about 0.03 kW/m2). The outcome is considered a success for the attempted
determination of modules’ operating temperature from their VOC. By means of a simple,
quasi-static thermal model, ECT is linked to the measured back-of-module temperature.
The analysis reveals that the latter underestimates significantly the module’s effective
temperature (which determines its output power) at the higher irradiance levels. It has
been confirmed that the ECT is the more accurate representative of the effective
operating temperature. A modification is proposed to the standard procedure [IEC93,
IEC11], which allows also the ECT determination for modules with irradiance- and
temperature-dependent ideality factors. The advanced method proposed in the present
146

work extends the applicability of ECT to irradiances as low as 30 W/m2 (versus 200
W/m2 in the standard procedure).
The tenth module tested at the installation in Grimstad (an aged triple-junction a-
Si module) cannot be modeled with the standard 1-exponential I-V model, in agreement
with the findings in [Mer10]. Due to time constraints, a more complex modeling of its I-
V curves (as in [Mer98]) is not attempted. The work presented in this Section shows the
importance of accurate I-V curve modeling for a reliable ECT determination, and some
possible complications arising from elaborate models. Therefore, ECT calculation has
not been attempted for the a-Si module. In addition, its much lower efficiency (as
compared to those of c-Si and CIS/CIGS devices) has naturally excluded it from the
high-latitude context of the thesis.

Figure 5.18: Fitting the residual error of ECT–TMOD–cG with Eq. (5.15) in the case
of the aged mc-Si module (‘GPV’).

5.5 TEMPERATURE COEFFICIENTS


For the modules studied in the present work, temperature coefficients of the
open-circuit voltages and of the maximum powers are derived from experimental data.
The temperature coefficient of a module’s short-circuit current is very small, and its
147

derivation from outdoor data is therefore not recommended [Moh10]. Nevertheless, this
has been done in Section 4.7 for the c-Si modules and the CIGS module. The
temperature coefficient of PMAX stems naturally from the ‘southern’ method for module
power rating [Moh10] applied later in Chapter 7. The methods used for estimation of
the temperature coefficient β of VOC are further outlined.

5.5.1 CLASSIC LINEAR FITTING


Fig. 5.19 illustrates the classic way of deriving β by linear fitting [Kin97a,
Jag09, Moh10, Mid10]. However, different approaches and degrees of scrutiny are
reported in the literature. Within the European outdoor testing network, the harmonized
procedure prescribed the use of data recorded at a rather narrow self-irradiance range
(between 995 and 1005 W/m2) [Moh10]. A much broader range, between 950 and 1050
W/m2, had been in use just one year earlier [Jag09]. A recent work used all data
recorded at irradiances above 950 W/m2 (measured with a co-planar mono-Si sensor)
[Mid10]. It was assumed that the effect of varying irradiance on VOC was negligible
compared to the impact of changing module temperature. This assumption would have
been acceptable if the peak irradiances did not exceed about 1050 W/m2. In fact, the
highest recorded values were about 1150 W/m2. The study of overirradiance events
presented in Chapter 4 of the thesis reveals that short bursts exceeding 1500 W/m2 are
possible in Southern Norway. However, in the previous study, the irradiance was
measured with 20-minute temporal resolution, which had greatly reduced the
probability of detecting the absolute maximum. In the present work, the narrow self-
irradiance range recommended in [Moh10] is used in the derivation of β.

Figure 5.19: Determination of the temperature coefficient β of the open-circuit


voltage VOC of the Chinese-made mono-Si module (denoted ‘RS mono’).
148

Two of the modules studied in [Mid10] had thermocouples built-in by the


manufacturer. For the third module, the thermocouple was attached at the back sheet
and tightened with a piece of styrofoam measuring roughly 9×9×5 cm3. One would
expect that that geometry had resulted in higher operating temperatures of the
corresponding part of the module due to an efficient thermal insulation. Indeed, despite
the absence of a front glass in that module’s design, its measured peak temperatures
exceeded those of the two glass-plastic c-Si modules whose backs were cooled by
natural convection. Consequently, the obtained value of β for that module had been less
accurate.
The standard linear fitting procedure was put to scrutiny in [Kin97a] with a
focus on temperature non-uniformity and its possible avoidance. The average module
temperature was obtained by using multiple thermocouples. Another major difference in
that work was the shadowing of the modules prior to the outdoor measurements.
Subsequently, the voltages and temperatures were measured during a thermal transient
occurring at stable solar irradiance. In addition, comparison was made between identical
modules with standard open-back configuration versus thermally insulated back surface
and edges. It was demonstrated that the values of β obtained for modules with open
backs were strongly influenced by temperature non-uniformities. The latter were caused
by the module’s frame and junction box(-es). On the other hand, the thermal insulation
resulted in more uniform back temperatures, which were in addition closer to the actual
cell temperatures. It was concluded that the true temperature coefficients could be
obtained only with thermal insulation, whereas the open-back approach was shown to
give about 10 % larger apparent coefficients [Kin97a]. That difference would be
smaller for glass-glass modules. However, in the present work, the thermal insulation
approach is considered unpractical, partly due to the large number of tested modules.
More importantly, operating conditions typical for a real-world, open-rack (and not
building-integrated) installation, have been targeted.
In the present experimental setup, the cells within a module are expected to have
somewhat different operating temperatures. Observations with a thermovision camera
(carried out at the sunny noon on 08 May, 2011) showed that each module was hotter in
front of the junction box. Cooling effects due to module framing were not observed.
However, the aluminum supports to which the modules were mounted had caused a
pronounced module cooling along their lengths. This was somewhat surprising, as these
metallic beams did not touch the back sheets of the modules. Due to a non-operational
floppy-disk controller of the aged camera used back then, none of the above
observations is documented in the thesis.
The linear fitting of VOC vs. TMOD can alternatively be applied for a limited
range of module temperatures. An example is given in Fig. 5.20, where different values
of β are obtained below 25°C and above 40°C (–0.123 V/K and –0.108 V/K,
respectively). The latter value is much closer to the one obtained from fitting the whole
temperature range (–0.110 V/K). The ‘low-temperature’ and the ‘high-temperature’ β
149

differ by 14 % for that particular module, but the difference reaches 26 % for the HIT
module. For two of the modules, the difference is negative. However, the applicability
of this approach is questionable, given the big spread in the plots. Therefore, only the
value obtained from fitting the whole TMOD range is used in the ECT calculation.
A major issue in the present work is the significant deviation of the measured
back temperatures from the calculated ECT at standard irradiance (as seen from Table
V.1). This may have additionally affected the accuracy of the temperature coefficients
obtained by linear fitting. However, after considering Eq. (5.3) and Fig. 5.9, that
deviation is assumed approximately invariant with cell temperature at standard
irradiance level (for a given module). Indeed, the spread in ECT–TMOD seen in Fig. 5.9
can be considered as quite small compared to the wide range of operating temperatures
(see Fig. 5.20). It is therefore concluded that the rather high values of the coefficients c
in the present test setup have not affected significantly the accuracy of β obtained from
linear fitting. The spread of the data points seen in Fig. 5.20 seems typical for all
outdoor measurements on modules in open-rack configuration [Kin97a, Jag09, Moh10,
Mid10]. It should rather be attributed to thermal transients than to the generally lower
back-of-module temperatures.

Figure 5.20: Fitting VOC of mc-Si module ‘ESS 2’ for different temperature ranges.
150

Following the conclusions in [Kin97a], the accuracy of β values obtained with


the classic method seems insufficient for the ECT calculation. The only exception is the
CIGS module, for which the fitting may have resulted in a more accurate β due to its
double-glass structure. However, one should first answer the question: Which value of β
is relevant for ECT calculation of modules with open backs? The updated IEC
procedure [IEC11] required a value determined after IEC 60891, but none of the eight
procedures linked therein has been available to the present author. Therefore, the answer
of the above question has not been obvious. The I-V curve model used for ECT
calculation assumes (according to the very definition of ECT) that the cells within the
module are identical (also in terms of their operating temperatures) [IEC93, IEC11]. On
the other hand, non-uniform module temperatures are observed in the present test setup,
and the measured VOC (the ‘cornerstone’ of ECT) is a superposition of the (non-
identical) voltages of the individual cells. One can therefore expect that the ‘apparent’ β
obtained from fitting VOC vs. TMOD is the one relevant to the ECT calculation, rather
than the ‘true’ β obtained with thermal insulation as in [Kin97a]. Indeed, at standard
irradiance (at which the fitting must be performed), Eq. (5.2) is reduced to

ECT  25°C   VOC  VOC,STC 


1
(5.16)

where the only variables are VOC and ECT. Consequently, β must be defined as the
derivative of VOC with respect to ECT. Taking into account that ECT–TMOD is
practically invariant with temperature at standard irradiance (as shown in Section 5.4), it
is concluded that the plot of VOC vs. ECT has the same slope (that is, β) as the plot of
VOC vs. TMOD.

5.5.2 DERIVING β FROM ECT ERROR ANALYSIS


An alternative method for deriving β (called the ‘RMSE vs. β’ method) is
developed on the basis of the ECT error analysis described earlier in Section 5.4. In
particular, Eq. (5.9) and Fig. 5.12 show that an erroneous β will increase significantly
the spread in the plot of ECT–TMOD over irradiance. As shown earlier, that plot obeys a
simple linear relation given by Eq. (5.3). Therefore, the goodness of fit of that plot can
be used as a measure of the accuracy of the value of β used to calculate ECT. The root
mean square error (RMSE) is found to be a suitable measure of the goodness of a linear
fit, as it is quantitatively related to the spread of the plotted data. Ideally, the plot will
have zero spread (i.e., a perfect line), and the RMSE of the linear fit will be zero as
well. In the real situation, the RMSE has a minimum (RMSE0) for the ECT calculated
with the most accurate value of β (assuming that all the other input parameters are
correct). One can create numerous plots of ECT–TMOD over irradiance by varying β
over a certain range of values, and then compare the RMSE values of the corresponding
linear fits. The method is illustrated in Fig. 5.21 and is somewhat similar to the method
151

for determination of a module’s series resistance RS proposed by the author in [Yor10b]


(discussed later in Chapter 6).
The method is applied to two data sets containing data from mostly clear-sky
days. In the first data set, 11 days from March and April, 2011 are represented. The
second data set initially contained data from 20 days in the period April-October 2011,
which later have been reduced to 14. The ‘RMSE vs. β’ method is applied to the
different sets, and somewhat different optimal values of β are obtained for each module.
The optimum RMSE values differ as well. The achieved minimum RMSE0 is about
0.6°C for most of the modules, and about 0.8°C for the CIGS module and that with
back-contact mono-Si cells (denoted ‘Sunc.’). The optimum β values obtained with this
method are listed in Table V.2, together with the values obtained with the classic
method. In addition, the measures of goodness of fit for both methods are given. For the
new method, those β values are chosen that resulted in a minimal RMSE0 at the
optimum.

Figure 5.21: Goodness of the linear fit of ECT–TMOD vs. irradiance as a function of
the value of β used to calculate the ECT. Data of five c-Si modules are shown.
152

The following observations are made on the results in Table V.2. For modules
for which the classic fitting has a goodness r2 of about 0.99, a quite similar β estimate is
obtained with the new method. However, for most of these modules a second, somewhat
different β estimate is also obtained with the new method from a different data set, also
with a good (i.e., low) RMSE0 at the optimum. For the module denoted ‘ESS 2’, the
estimates from both methods are practically equal, but differ by about 6.5 % from the
value provided by an independent laboratory. For the remaining modules, the β
estimates from the two methods differ notably, by up to 8 % (except for the CIGS
module, for which the difference is only 1.4 %).

TABLE V.2: Temperature coefficients β of VOC obtained with two different methods.
Measures of goodness of fit are also given for each method. For some modules, the 2nd
method yields two different values from two different data sets with a comparable
optimum goodness of fit (RMSE0).
Module β [V/K], ‘VOC vs. TMOD’ r2 β [V/K], ‘RMSE vs. β’ RMSE0 [°C]
‘BP’ –0.0725 0.993 –0.0755/–0.0723 0.5/0.6
‘GPV’ –0.0738 0.988 –0.0755/–0.0732 0.5/0.6
‘Sunc.’ –0.0591 0.949 –0.0566 0.8
‘RS mono’ –0.1402 0.992 –0.140/–0.137 0.6
‘RS poly’ –0.1394 0.974 –0.1293 0.6
‘HIT’ –0.1105 0.975 –0.1042 0.6
‘ESS 1’ –0.1277/–0.123* 0.979 –0.1208 0.6
‘ESS 2’ –0.1104/0.118* 0.988 –0.1110 0.6
‘CIS’ –0.2115 0.976 –0.2145 0.8
*Values provided by an independent laboratory.

To conclude, both the classic ‘VOC vs. TMOD’ method and the ‘RMSE vs. β’
method (proposed in the context of Section 5.4) give quite similar estimates of β. With
outdoor data, the classic method results in fits of different quality for the different
modules. The other method can give ambiguous results, differing by up to 4.5 %,
perhaps more. Thus, the latter value is considered to be the ultimate accuracy
achievable with the new method (with the available experimental data). The uncertainty
of the classic method (as estimated in Matlab™, with 95 % confidence bounds) ranges
between ±0.6 % and ±1.5 % for the different modules. These uncertainties are
considered to be rather optimistic, as they do not account for all uncertainties in VOC
and TMOD. A very detailed uncertainty analysis as in [Dom10] is not performed in the
present work due to time constraints. For the purposes of ECT calculation, the average
of the β estimates for each module is used. For modules with two estimates, their
difference is taken as a rough approximation of the total uncertainty. For modules with
three estimates, the overall uncertainty is approximated by two standard deviations. For
153

the different modules, the so assumed uncertainties of β range between 1.4 % (for the
CIGS module) and 7.5 % (for the mc-Si modules denoted ‘ESS 2’ and ‘RS poly’).

5.6 THERMAL RESISTANCES OF THE TESTED MODULES

Figure 5.22: Determination of a module’s thermal resistance from outdoor data.

TABLE V.3: Thermal resistances of the modules.


Module cT [K(W/m2)-1]
Aged BC mono-Si (‘BP’) 0.0235
Aged mc-Si (‘GPV’) 0.0220
Back-contact mono-Si (‘Sunc.’) 0.0245
RSS mono-Si (‘RS mono’) 0.0240
RSS mc-Si (‘RS poly’) 0.0278
‘HIT’ 0.0251
mc-Si ‘ESS 1’ 0.0238
mc-Si ‘ESS 2’ 0.0251
‘CIS’ 0.0245

As in [Mid10], the in-situ thermal resistances of the studied modules are


estimated by fitting their ECT data, the measured ambient temperatures, and the
measured irradiance with Eq. (3.1), see Fig. 5.22. The obtained values (see Table V.3)
are on the average lower than the ones typical for open-rack PV configuration [Luq03a]
and those obtained in [Mid10]. This can be attributed to the elevated location of the
present test site and the typically strong winds during most of the year. On the other
hand, the modules tested in [Mid10] were partially deprived from wind-assisted cooling
154

by the cottage (see Fig. 3.5). A suspiciously low value is obtained for the CIGS module
having a glass-glass structure. However, its ECT values are obtained with a strongly-
varying ideality factor, which may have been modeled with poor accuracy. Similarly,
the estimated thermal resistance of the Chinese-made mc-Si module (denoted ‘RS
poly’) is about 16 % above the mean of the other c-Si modules, while having the same
glass-plastic structure. This module is the one with the strongest variation of the ideality
factor (see Fig. 6.24), which may have been modeled with poor accuracy as well. The
higher peak ECT values obtained for this module as compared to the rest (to be
discussed in Chapter 6) also support this explanation. Therefore, the ECT calculation of
these two modules needs refinement by improving their ideality factor models. This is
outside the scope of the present work, and may be addressed in future.

5.7 DISCUSSION
IEC 60904-5 states that the ECT procedure should only be used for irradiances
above 0.2 kW/m2 [IEC93, IEC11]. The cause (according to IEC) is the rapid drop of the
temperature coefficient of VOC, β, at lower irradiances. However, it is not clarified
whether this requirement applies to the reference irradiance G1 (if different from the
standard irradiance) or to G2, or to both. As already discussed in Section 5.4, the
procedure uses this coefficient only at G1, which defaults to 1 kW/m2 in the present
work. Defined as the partial derivative of VOC with respect to the PV device’s effective
temperature, β is integrated over temperature (at a fixed irradiance) to give the voltage
deviation from its reference value (here, VOC,STC). Prior to this, the voltage is translated
between two irradiance levels (at a fixed device temperature). Mathematically, this
translation is done by integrating the partial derivative of the voltage with respect to
irradiance (as shown in Section 5.4). It is the device’s ideality factor n (and not β) that
appears in the result of this integration, and therefore no explicit irradiance dependency
of β is involved in the calculation of ECT. On the other hand,  G is proportional to
n T according to the theorem for equality of the mixed partial derivatives (second-
order derivatives of VOC in this case). Thus, through Eq. (5.13), any irradiance
dependency of β is accounted for indirectly in the advanced ECT calculation proposed
in the thesis.
The present work reveals the actual low-irradiance limit for accurate calculation
of ECT. It is demonstrated in Section 5.4 that the latter can be well defined for
irradiances G2 as low as 0.05 kW/m2 (but not for all types of PV modules). The errors
obtained for some devices at such low illumination levels are traced to a big change of
their ideality factors (i.e. a large magnitude of n G ). In addition, the accuracy of the
standard ECT procedure is found to depend on a possible variation of n over the whole
range of irradiance values. For such devices, the associated error increases significantly
with decreasing irradiance. Using the alternative formulation proposed in the present
155

work for modules with variable ideality factors, much better ECT accuracy is achieved
for irradiances down to about 0.05 kW/m2. However, it is not correct to attribute (as in
IEC 60904-5) the low-irradiance limit to a strong irradiance dependency of β (i.e. a
large magnitude of  G ). Such limitation eventually comes from a sizable n G
(and not from a high n T to which  G is proportional).
The new ECT formulation proposed in the thesis was recently presented in
[Yor12e]. It needs a more thorough (and eventually, independent) verification, since the
back-of-module temperatures used in the analysis suffer from significant measurement
uncertainties. Future work may also investigate in greater detail the ideality factor
variations observed for most of the PV devices studied here. Eventually, the advanced
procedure (once verified) may be proposed for discussion by the PV community as a
new international standard draft (or as part of a future, 3rd edition of IEC 60904-5). Such
possibilities stress the significance of the work presented in the thesis.

Figure 5.23: Data used to estimate β of the modules denoted ‘ESS 1’ and ‘ESS 2’.

The surprising difference (10 %) between the values of β obtained for the two
modules based on ESS™ deserves special attention. Both devices come from the same
manufacturer and have identical design and size. Fig. 5.23 shows the VOC and back
temperature data used to determine the temperature coefficient of each module (for a
very narrow range of self-irradiance values around 1000 W/m2). The peak temperature
156

of the module denoted ‘ESS 2’ exceeds by 5°C that of the other module, whereas its
lowest voltage is somewhat higher. Consequently, a notably higher (that is, better) β is
obtained for the former module. The averaged estimates obtained for all modules are
given in Table V.4 in relative units for ease of comparison.
For the ‘ESS 2’ module, very similar results for β are obtained (within 0.6 %)
with the two methods used in the present work. The average value in Table V.4 includes
also the estimate provided by the independent laboratory, and thus a large uncertainty is
obtained (two standard deviations amount to 7.5 %). Similar uncertainty (in terms of
difference between the estimates from the two methods) is obtained for the module ‘RS
poly’, whereas the average is the lowest in magnitude among the screen-printed c-Si
modules. The results obtained for the two modules represented in Fig. 5.23 may have
been affected by the installation of another two modules behind them, which constitute
a different experiment, see Fig. 5.24 [Ver12]. In this way, the cooling by natural
convection and wind has been blocked partially since mid-July, 2011. However, a
detailed investigation of the effects resulting from this modification of the present test
setup is outside the scope of the thesis.

TABLE V.4: Temperature coefficients of VOC in relative units.


Module β [%/K]
Aged BC mono-Si (‘BP’) –0.337
Aged mc-Si (‘GPV’) –0.345
Back-contact mono-Si (‘Sunc.’) –0.264
RSS mono-Si (‘RS mono’) –0.311
RSS mc-Si (‘RS poly’) –0.297
‘HIT’ –0.246
mc-Si ‘ESS 1’ –0.333
mc-Si ‘ESS 2’ –0.302
‘CIS’ –0.377

The beneficial temperature coefficients obtained for the two high-efficiency


modules (denoted ‘HIT’ and ‘Sunc.’) correspond well with the theory presented in
[Gre03]. Both devices feature a relatively high per-cell VOC (729 mV and 683 mV at
STC, respectively), which is a prerequisite for its reduced temperature sensitivity. The
former module has the best β, as expected. However, a detailed quantitative analysis of
the obtained results with respect to solar cell physics is not aimed in the present work.
The ECT values obtained for the module denoted ‘RS poly’ at irradiances close
to the standard level, are notably higher (by about 6°C on the average) than those of the
other glass-plastic c-Si modules tested (see Fig. 5.25). A possible cause is poorer natural
ventilation of that module, as well as of the one denoted ‘ESS 2’, due to the couple of
modules installed behind them (see Fig. 5.24). However, the corresponding ECT values
157

Figure 5.24: North-facing modules tested in a different work [Ver12].

Figure 5.25: Temperatures of the c-Si modules on 21 June, 2011.


158

of ‘ESS 2’ are not as high. Another possible cause is an error in the value of β used in
the calculation. At such levels of illumination and for module temperature of about
45°C, an error of +6°C in ECT corresponds to a –30 % error in β according to Eq. (5.9),
which is highly unlikely. On the other hand, even a large error in the ideality factor has
very little or no effect on ECT at irradiances of about 1000 W/m2. However, it is also
possible that the correction applied to the label value of VOC,STC, which is derived in
Section 5.6 (see Table V.1), contains a significant error. Given the indirect method used
to derive the correction (involving β and a varying n), as well as the significant
uncertainty of β obtained for that module, it is difficult to identify the exact cause for its
higher ECT at this stage of the work. The characterization of this PV device is very
challenging mainly because its I-V curves cannot be accurately modeled with the classic
1-exponential I-V curve model, which assumes a fixed ideality factor.
159

CHAPTER 6 I-V CURVE ANALYSIS


6.1 CHAPTER AIMS AND OBJECTIVES
The main goal of this Chapter is to analyze and model the measured I-V curves
of the studied PV modules. In the contexts of characterizing the power performance of
the modules and calculating their equivalent cell temperatures from VOC, particular
attention must be paid to the higher-voltage region between the MPP and open circuit.
An important task is to identify a suitable model which describes accurately the I-V
curves over the range of typical operating conditions, and to test the limits of its
applicability. The derivation of a mathematical relationship between power performance
and device physics requires evaluation of the key parameters which govern the shape of
the I-V curve and analysis of their behavior at different irradiances, operating
temperatures and currents.
The Chapter begins with a literature review on I-V curve theory and modeling.
The widely used 1- and 2-exponential models are discussed, as well as the key current
conduction mechanisms in different types of c-Si cells. Later on, factors influencing the
I-V curves of modules are outlined. Several limits of the applicability of the 1-
exponential model which is usually preferred for modules are identified throughout the
Chapter, including those specific for the CIGS and CdTe technologies. The methods for
extracting model parameters from single and from multiple I-V curves which are found
in the literature are reviewed and discussed. A new method invented by the present
author is described in detail and is verified using synthetic I-V data. The method is then
used to extract the parameters of the modules tested in Grimstad and the results are
discussed. The double-light method is used to probe for a current dependency of the
series resistance and to verify the values obtain with the author’s own method. The
variability of the obtained ideality factors is discussed and modeled, and the temperature
dependency of the reverse saturation currents is modeled with a simple exponential
equation. An existing differential technique is modified by adding a correction for the
series-resistance effect and is then used to study the behavior of the ideality factor
between open circuit and the MPP. Eventually, an alternative method for parameter
extraction is discovered.

6.2 I-V CURVE ANALYSIS AND MODELING


The I-V curve of a non-shadowed PV module has three remarkable points – the
maximum-power point (MPP), short circuit and open circuit. In case of uniform
illumination, the short-circuit current ISC may be useful in estimating the normal
irradiance component G (called the ‘in-plane’ or ‘plane-of-module’ irradiance). Thanks
to its strong temperature dependency, the open-circuit voltage VOC can be used in
estimating the module’s effective operating temperature (the equivalent cell
temperature, ECT) [IEC93, IEC11]. The MPP (VM,IM) defines the module’s available
160

power PMAX(G,ECT) = IMVM and its instantaneous efficiency η(G,ECT) = PMAX/(AG)


where A is its area. The three points together define another quantity, the fill factor
FF(G,ECT) = PMAX/(ISCVOC), which characterizes the shape of the I-V curve. A high
FF means a more ‘rectangular’ curve. Besides these key points, the remainder of the I-V
curve contains information about other physical properties of the device. The resistive
losses are described by module’s internal series resistance RS, which affects the slope
between the ‘knee’ (the MPP region) and open circuit. However, other intrinsic solar
cell properties (to be discussed shortly) also influence this slope, which affects the FF.

Figure 6.1: Fill factor evolution on 29 April, 2011 (a clear-sky day) for the ten PV
modules tested in Grimstad.

Fig. 6.1 shows the FF values of the modules tested in Grimstad from a clear-sky
day. The two thin films (denoted ‘a-Si 3J’ and ‘CIS’) have notably lower FF than the
eight c-Si modules. The latter show similar behaviors over the entire irradiance range.
The Chinese-made mc-Si module (denoted ‘RS poly’) has a much lower FF at
intermediate and high irradiances, whereas the aged European-made mc-Si module
(denoted ‘GPV’) has a remarkably lower FF at the low irradiances.
Carrero et al. [Car07] suggested quantitative guidelines for selection of
commercial mono-Si modules based solely on the label values of ISC,STC, IM,STC,
161

VOC,STC and VM,STC. A set of dimensionless quantities was defined to allow module
comparisons: the STC fill factor FFSTC, as well as the ratios iM,STC = IM,STC/ISC,STC and
vM,STC = VM,STC/VOC,STC. Table VI.1 compares the values of these quantities for the c-Si
PV modules studied in the present work. The best numbers are indicated with green
color. Surprisingly, one of the two high-efficiency mono-Si modules is not recognized
by these selection criteria. On the other hand, the aged mono-Si module has the best ‘as
new’ FFSTC and the highest iM,STC. The comparison shows that the proposed selection
procedure cannot distinguish between higher and lower efficiency modules. In the case
of the module denoted “RS poly”, the three STC parameters fail to explain its much
lower FF under realistic outdoor conditions seen in Fig. 6.1. It is obvious that a more
detailed analysis of module I-V curves is needed. The ideality factor values obtained for
c-Si modules in the present work refute the analytical procedure suggested in [Car07]
which was based on the assumption that n=1. As both n and RS affect the slope of the I-
V curve between VOC and VM, the adoption of an erroneous value for n would lead to a
wrong estimate for RS. Chapter 5 of the thesis shows that even high-efficiency mono-Si
modules can have ideality factors as high as 1.35.

TABLE VI.1: Comparison of the c-Si PV modules tested in Grimstad in terms of


dimensionless figures of merit suggested in [Car07].
Denotation, Cell Technology, η0 as new FFSTC iM,STC vM,STC
‘BP’, aged buried-contact mono-Si, 13.5 % 0.771 0.944 0.817
‘GPV’, aged screen-printed mc-Si, 11.7 % 0.741 0.938 0.791
‘Sunc.’, back-contact mono-Si, 17.4 % 0.769 0.932 0.826
‘RS mono’, screen-printed mono-Si, 15.5 % 0.762 0.946 0.806
‘RS poly’, screen-printed mc-Si, 14.4 % 0.742 0.925 0.802
‘ESS 1’, screen-printed mc-Si, 12.6 % 0.744 0.909 0.819
‘ESS 2’, screen-printed mc-Si, 12.7 % 0.744 0.909 0.819
‘HIT’, heterojunction a-Si/mono-Si, 17.3 % 0.749 0.919 0.814

6.3 REVIEW OF THE EXISTING I-V CURVE MODELS


Modeling the I-V curves of PV devices is a vast and complicated topic for many
reasons. The available literature covers six decades of continuous development of
semiconductor materials and cell structures with reasonable efficiencies [Lof93, Tre98,
Goe03]. The available models differ in their mathematical formulation and in the
number of free parameters. Some authors applied a priori non-physical assumptions or
limitations for the values of some parameters. Some parameters such as the saturation
currents have a strong temperature dependency, and may also vary with device voltage
[Gre03]. Others, such as the series resistance RS, may vary with current density and
irradiance [Abe93a]. Generally, the area of a solar cell is not an equipotential surface as
162

assumed by the commonly used one-dimensional models [Fis00, Alt96, Alt05, Hei05,
Car09a]. PV modules can contain many tens of individual cells that are not identical,
plus several bypass diodes. Non-uniform temperature distribution across a PV module’s
area is typical in outdoor operation [Kin97a, Mül09]. Modern terrestrial PV modules
usually employ large-area cells, and therefore also temperature non-uniformity within
individual cells is possible. Thin films are still undergoing innovations at a rapid pace
[Hul10, Gre11], whereas their structure and the dominating current conduction
mechanisms are generally very different from those in c-Si devices [Sha04, Nik10]. The
scope of this Chapter is limited to single-junction c-Si, CIS/CIGS and CdTe devices.
Thin-film amorphous-Si (a-Si) devices are not considered in the thesis for several
reasons. Their efficiencies are much lower compared to CIS/CIGS and CdTe and make
them impractical at high latitudes. Their I-V curves and performance require much more
complex modeling [Mer98, Hul10, Mer10]. For example, the short-circuit current of
tandem and triple-junction a-Si devices has much greater spectral sensitivity than that of
c-Si, CIS/CIGS and CdTe devices [Kin00, Wil04b]. On the long term, it seems very
unlikely to the present author that the a-Si technology can win a market share similar to
that of CdTe, which is an order of magnitude smaller than that of c-Si. Micromorphous
devices (tandems of a-Si and μc-Si junctions) have a much greater market potential, but
are also excluded from the present work because of modeling complexity.

6.3.1 CLASSIC MODELS


The most widely used mathematical models of cell I-V curves are the 1- and the
2-exponential model. Their general forms for illuminated condition are given by Eqs.
(6.1) and (6.2). The lumped circuit associated with the latter is given in Fig. 6.2,
whereas a single-diode version (not shown) is the equivalent of the simpler model. I and
V denote the terminal current and voltage of the cell, respectively. The photogenerated
current (photocurrent) IL is usually considered to be voltage-independent, whereas the
irradiance G is ideally being measured with a calibrated reference sensor which is a
perfect spectral match to the PV device under consideration. In both equations, the
second term represents the current losses through an Ohmic (i.e., linear) shunt with
resistance RP, which is typically very large in good-quality c-Si cells. This term would
generally dominate the slope of the I-V curve at low forward voltages, as well as at non-
breakdown reverse-bias conditions. Too low an RP would divert significant part of IL at
higher forward voltages, thus causing a somewhat reduced open-circuit voltage VOC.
The series resistance RS is often assumed constant along the I-V curve. Under special
conditions, the associated voltage drop IRS can reduce the short-circuit current ISC. As
long as IRS<0.25 V, ISC of a c-Si cell may be considered identical with IL and linearly
dependent on irradiance, if the spectrum is not varied [Wol63]. I0, I01 and I02 are termed
(reverse) saturation currents and increase strongly with device’s temperature. However,
there is a lack of agreement in the literature on the exact mathematical form of their
163

temperature dependencies [Coo97, Gow99, Gre03, Luq03b]. The dimensionless


quantities n, n1 and n2 are the so called ideality factors of the diodes (also, diode quality
factors), which are considered constant, and usually n1<n2 is assumed.
V  IR S   V  IR S  
I  V   I L  G ,T    I 0  T  exp    1 (6.1)
RP   nVT  

V  IR S
I  V   I L  G ,T  
RP
(6.2)
  V  IR S     V  IR S  
I 01  T  exp    1  I 02  T  exp    1
  n1VT     n 2 VT  

Figure 6.2: The lumped-circuit PV cell model corresponding to Eq. (6.2).

Dark I-V curves are obtained using a power supply with the cell acting as a load.
In this case, the current source is removed from the lumped-circuit model and the term
IL is omitted in the corresponding equation. Some model parameters may have different
values in dark and under illumination [Sit89, Abe93a]. It is sometimes assumed that the
illuminated curve can be obtained by shifting the dark curve by ISC along the current
axis, if the series-resistance effect is negligibly small. This assumption is called ‘the
shifting approximation’ (also, ‘the superposition principle’) and is not always fulfilled
[Rob94a, McI01]. Also reverse dark and illuminated curves can reveal important cell
properties. Eqs. (6.1) and (6.2) imply that the illuminated curve of the forward-biased
cell resides in the 1st quadrant and the dark forward I-V curve resides in the 4th quadrant.
It is also possible to extend the illuminated curve measurement beyond VOC in the 4th
quadrant. There are variants of the 1- and 2-exponential model, which use the opposite
convention and have opposite signs of the major terms, whereas the term V–IRS appears
in the exponentials. It is sometimes practical to use the current density J (J-V curves)
instead of the total cell current I [Hei05, Fon11a], in which case the photogenerated
164

current and the saturation currents are replaced by the corresponding current densities,
whereas RS and RP have area-normalized units of Ωcm2.
A special type of characteristic, the ISC-VOC curve, can also be useful to study a
PV device [Wol63, Abe93a, Sch01]. If the illumination intensity is measured instead of
the ISC, one obtains the suns-VOC curve [Mac00]. Such curves benefit from eliminating
the series-resistance effect, but their measurement requires controlled laboratory
conditions and specialized hardware. Due to limitations of the present experimental
setups, only single-quadrant illuminated forward I-V curves are studied in the thesis.
The 1-exponential model stems from Shockley’s ideal diode equation derived
for p-n junctions in diodes and transistors [Sho49]. No ideality factor appeared therein
(i.e., n≡1), whereas the voltage drop IRS was considered negligibly small. In the early
1960s, n of most practical c-Si cells ranged between 2 and 3 at room temperature, but
values as high as 5 were considered possible. The origin of those high values was not
clearly understood [Sho61, Wol63]. In an attempt to model the effects of the distributed
nature of RS, a 2-diode circuit with 2 current sources and 2 series resistances was
considered in the latter work. Experimental ISC-VOC curves of a 2-cm2 c-Si cell were
modeled with a 2-exponential equation with n2≡1 and a temperature-dependent n1 of
about 3. This seems to have been the first application of a 2-exponential model in the
literature [Wol77]. However, the value n2≡1 was derived in [Wol63] from theoretical
considerations (following Shockley’s diffusion theory) rather than from the presented
experimental data. Exact fitting of an illuminated I-V curve was eventually achieved by
using a 4-exponential equation. The corresponding lumped circuit included 4 diodes, 2
series resistances and 2 current sources [Wol63]. That work showed clearly the possible
complexity of a PV cell’s I-V curves. A later work studied cells exposed to high doses
of proton radiation, for which the I-V curves were dominated by an exponential term
with n1 ranging between 1.08 and 1.1 [Wol77].
In later works on I-V curve modeling of c-Si and thin-film cells, the 1-
exponential model was used more rarely [Sit89, Cra95, Oue99, Che01, Hei05, Che11].
A few authors even assumed n≡1. Textbooks can give unrealistic I-V curves of Si solar
cells [Bre10b]. For example, a recently published textbook recommended n=1 for
modeling of PV cells, modules and arrays, regardless of the technology type, which is
simply misleading [Kal09]. In addition, the example I-V curve given therein was non-
physical, with a slope dI / dV   at open circuit. In general, the 1-exponential model
is a rough approximation, because the relative importance of different recombination
losses within the cell may change along the I-V curve. A multi-exponential model is
therefore preferable, in which the saturation currents and the ideality factors depend
only slightly on voltage [Alt96]. The more general 7-parameter 2-exponential model (in
which n1 and n2 are free parameters) was found superior to three alternative techniques
for translating I-V curves of solar cells to desired irradiance and cell temperature
165

[Coo97]. Green proposed a more abstract generalized model, which was not limited to a
particular PV cell type [Gre03].
Many authors who adopted a 2-exponential model made a priori assumptions
about the ideality factors n1 and n2. The 1st exponential term is commonly attributed to
low-injection Shockley-Read-Hall (SRH) recombination in the emitter and base regions
of c-Si cells, and n1≡1 is often assumed [Rob94c, McI01, Dyk04, Mey04, Hei05,
Ale11]. Some high-efficiency mono-Si cells may indeed follow the 2-diode model quite
precisely [Bre10b]. However, some cases of SRH recombination involve unequal
capture cross-sections for electrons and holes, resulting in injection-level dependent
recombination rate characterized with ideality factors exceeding 1. These cases include
but are not limited to interstitial iron contamination typical for mc-Si cells, boron-
oxygen complexes in p-doped c-Si, as well as surface recombination [Abe93b, Mac00,
Mac01, Sch01]. Furthermore, many authors assumed also n2≡2 [Cha85, Gow99, Mül04,
Hao05, Pys07, Sze07b, Wer08, Gre09]. The same assumption was made in [Mid07], but
n1 was allowed to differ from unity in that work. Nevertheless, n1=1 was used in PSpice
simulations of a mono-Si module, which were focused on bypass diodes and partial
shadowing. The work of Gow and Manning [Gow99] referenced above has become
very popular in the power electronics community, and has been cited in at least three
hundred later works [Sch12].
The 2nd exponential term is usually attributed to recombination in the space-
charge region of the p-n junction [Vei95, Bre10c]. Several studies showed that the
ideality factor associated with this mechanism is generally smaller than 2 in c-Si cells
[McI00, McI01, Gre03]. In addition, the magnitude of the corresponding recombination
current was found to be negligible in such cells. According to diode physics, that
mechanism becomes significant only if the width of the space-charge region is
comparable to the carrier diffusion lengths [Str00]. However, in c-Si cells, the width of
the depletion region is typically well below a micron (for an approximated formula, see
[Nel03b, Sze07a]), whereas the minority carrier diffusion length can exceed the wafer
thickness, which is usually a few hundred microns [Goe03, Neu07, Q-C11]. On the
other hand, recombination at unpassivated edges and trap-assisted tunneling in the
depletion region can result in exponential currents with n2≥2 [Rob94b, McI01]. Non-
linear shunts in c-Si cells can have exponential characteristics with ideality factors
exceeding 3 [Bre04, Bre11b]. The semi-logarithmic dark I-V curve of a modern
15.6×15.6 cm2 mc-Si cell is said to have a linear region below 0.3 V with ideality factor
of about 3 [Bre10b]. An exponential shunt with ideality factor of about 6 has been
identified in a mc-Si cell [Zho11].
For many c-Si cell types, the measured I-V curves could not be modeled with
the common 2-exponential model. Various effects, such as recombination-rate
saturation, were shown to introduce shoulders (‘humps’) in the semi-logarithmic curves
[Rob94c]. Similar behavior was observed for high-efficiency mc-Si cells due to
166

injection-level dependent minority carrier lifetime [Mac00]. However, the latter study
considered illumination levels far exceeding the range typical for non-concentrating PV.
McIntosh [McI01] used a more complex 2-diode lumped circuit to model I-V curves of
cells with floating-junction surface passivation. A 3-diode circuit was needed to analyze
the effects of resistance-limited enhanced recombination (e.g., enhanced edge
recombination) on the I-V curve [McI01, Hus04]. A 3-exponential equation with a
current-dependent RS was used in a recent work to fit the experimental curve of an
elongate c-Si cell [Fon11a, Fon11b].

6.3.2 SAETRE-MIDTGÅRD-DAS MODEL


A mathematically very different model was introduced in [Mid08, Sae11,
Das11]. Only two dimensionless parameters were found necessary to describe the shape
of an individual illuminated forward I-V curve with a reasonable accuracy. The reverse
characteristics (including the breakdown region) could also be modeled by adding one
more parameter [Mid08, Sae11]. However, this model does not eliminate the need for
modeling the irradiance and temperature dependencies of device’s ISC and VOC. In
addition, the relationship of this model’s parameters to e.g. the device’s RS is not clear.
An attempt to link this model to the 1-exponential I-V curve model was made in
[Das12]. Due to its intrinsic complexity, the Midtgård-Saetre-Das model is not
investigated in the thesis. However, the present work provides an enormous amount of
I-V curves from several very different types of PV modules, which may serve as a basis
for further investigation of the model.

6.3.3 SOLAR CELLS WITH CURRENT- AND IRRADIANCE-DEPENDENT RS


The models presented in Eqs. (6.1) and (6.2), as well as in Fig. 6.2, are one-
dimensional and assume that the entire terminal current flows through the series
resistance, whereas the whole p-n junction is biased by V+IRS. Also, invariant RS is
assumed between illuminated and dark conditions, between different illumination
levels, as well as along the I-V curve. However, the series resistance is distributed in
nature, similar to the photogenerated current, shunt resistance and reverse saturation
current(-s). In a mainstream n+p screen-printed c-Si PV cell operated at the MPP, parts
which are closer to the front terminals operate at a lower forward bias compared to the
remote parts. Such cells have completely metallized rear (aluminum back-surface field,
Al-BSF), and therefore the emitter sheet resistivity determines the 2D current flow
patterns [Abe93a, Neu07, Q-C11]. A lumped RS is a good approximation only if the
potential differences over the cell area are small compared to VT [Hei05]. Some current
(and power) is lost due to these differences. Because of the generally exponential nature
of the local J-V curves (to be discussed further in this Section), a small increase of the
operating voltage around the MPP can result in a significant drop of J [Alt96]. This
167

results in J-V mismatch between different parts of the cell. The associated power loss is
termed as “nongeneration loss” [Alt05].
For some cell designs, such as buried-contact (BC) and passivated emitter, rear
locally-diffused (PERL) c-Si cells [Wan90], different current flow patterns in dark and
illuminated conditions can result in different RS, which can also depend strongly on the
current density [Abe93a, Alt96, Pys07]. An earlier generation PESC c-Si cell, due to the
full Al-BSF, had a comparatively low, almost invariant illuminated RS, practically equal
to that in dark [Gre86, Abe93a]. Modern industrial screen-printed c-Si cells have very
similar structure, but with SiNX instead of SiO2 as emitter passivation and anti-
reflection coating (ARC). A p+ Al-BSF is applied over the whole back side of the cells.
Therefore, they are similarly expected to have RS that is nearly invariant with current
density and irradiance level. However, some interdigitated back-contact (IBC) cells and
elongate cells have a current-dependent RS [Fon11a, Fon11b]. For the studied IBC cells,
30 to 40 % increase was seen between open circuit and the MPP. The variation was
larger for the cell with the higher open-circuit RS. Otherwise, the series resistance of
IBC cells can be very low compared to that of cells with front metallization. This is
because the current in the former is not conducted along the emitter (which adds a
significant contribution to the total RS in the latter), whereas nearly half of the back
surface can be covered with metallization [Ker06, Neu07]. This is confirmed for the c-
Si modules tested in the present work, as described later in Section 6.4.
Araujo et al. [Ara86] analyzed analytically the effect of a distributed cell
resistance for a wide range of photocurrent densities, including ones typical for
concentrator solar cells (e.g. JL = 10 A/cm2). The analysis focused on the emitter and
base regions, while the front and the (full) rear metallization were considered
equipotential. With this assumption, the illuminated series resistance was found to
decrease between short circuit and open circuit. However, at lower irradiances (resulting
in lower photocurrent densities), the short-circuit value practically equaled the open-
circuit value. According to numerical simulations, the difference was less than 2 % for
JL < 0.2 A/cm2, whereas the typical 1-sun photocurrent density of most c-Si cells is
within 0.04 A/cm2 [Neu07]. The results were obtained for emitter and base resistivities
typical for a high-concentration cell, but a practically constant RS follows from the
derived equations also for non-concentrating PESC cells. However, the fingers of the
front metallization grid have a non-zero resistance and the voltage drop along them
cannot be neglected [Zha94, Alt96, Alt05], given their relatively large lengths in
modern large-area c-Si cells. It was demonstrated that relatively high finger resistances
(well above 1–2 Ω/cm) would result in significant non-linear increase of the lumped
grid resistance between open circuit and short circuit [Alt96]. In modern industrial
screen-printed c-Si cells, the finger resistance is in the order of 0.15-0.30 Ω/cm
[Met07a], but on the other hand their typical lengths (see e.g. [Met07b]) are much larger
168

than in the 4-cm2 PERL cells studied in [Alt96]. Therefore, the possibility for some
variation of RS along the I-V curve of a large-area screen-printed cell should be taken
into account. This is mandatory in the case of passivated emitter and rear cells (PERC)
[Bla89, Gat11, Ste11] with a structure similar to that of PERL cells, in which lateral
current flow in the base increases RS and makes it dependent on current density. Screen-
printed c-Si cells can be plagued by finger discontinuities and high contact resistance
between the front metallization and the emitter [Hei05]. Both defects result in excessive
potential differences over the cell area leading to I-V curves with drastically reduced fill
factors. Similar curves were successfully simulated using 1-exponential local J-V curves
with n = 1.3, while varying the series resistance in different parts of the cell. Adding
redundant lines to the front grid is often used in order to mitigate the effects of locally
increased series resistance by providing alternative paths for the current (see Fig. 6.3).

Figure 6.3: Redundant line in the front metallization along the cell perimeter.

Fischer et al. [Fis00] proposed that an additional resistive parameter (one that
accounts for the current crowding phenomenon) is used when fitting dark and
illuminated J-V curves of non-optimum screen-printed cells in development stage. With
this approach, the authors were able to obtain consistent 2-diode model parameters for
all curve types, including the JSC-VOC curve. For a pseudo-square screen-printed mono-
Si cell of size 12.5×12.5 cm2 with an H-pattern front metallization grid, RS was found
to increase from 5.5 mΩ at open circuit to 11.6 mΩ at the MPP [Tru07]. However,
details on the estimation algorithm were not given, whereas an ideal p-n junction (with
n=1) was assumed for the presented resistance mapping method. In another work, the I-
169

V curve of a similar cell was fitted with a 7-parameter 2-exponential cell model with a
variable RS [Car09a]. The global RS was represented as a sum of a constant pure Ohmic
resistor and a voltage-dependent distributed resistance. The series resistance was found
to rise linearly with current at a rate of 0.11 mΩ/A, starting from about 3.5 mΩ at open
circuit. The authors concluded that neglecting the distributed nature of RS would result
in unrealistic values of the other fitted parameters, especially n2.
Resistive and recombination losses in screen-printed mono-Si cells with B-
doped base were studied in [Ste11] using experimental and device simulation results.
Before the initial degradation, the reference cell with emitter sheet resistivity of 80 Ω/□
had an almost constant series resistance (around 0.5–0.6 Ωcm2) between open circuit
and MPP. However, RS at MPP (estimated with the double-light level method) rose to
about 0.8 Ωcm2 in the degraded state, whereas the open-circuit value was not affected.
On the other hand, the MPP value obtained with the JSC-VOC method was practically
unchanged. The authors recommended the former method as the more accurate one for
cells with strong injection-level effects. However, it is questionable whether injection-
dependent bulk SRH lifetime can cause strong variation of RS along the I-V curve in
cells with fully metallized rear. As already discussed above, the 2D current flow pattern
in such cells is determined by the resistances of the emitter and the fingers. The
emitter’s sheet resistance in n+p cells is defined by the phosphorus doping profile and
cannot be affected by formation of boron-oxygen complexes, because the boron doping
concentration is some orders of magnitude lower than that of phosphorus. Finger
resistance is also unaffected during degradation of this particular type. Therefore, the
current flow pattern in the p-doped base is not supposed to change in the degraded state.
The present author believes that it was the ideality factor of the cell, and not the series
resistance, which had changed due to degradation. This hypothesis is supported by the
work of Schmidt et al. [Sch01].
The main conclusion drawn from the above literature review is that I-V curve
modeling with a constant RS should not be done regardless of the type of the studied PV
technology. Screen-printed c-Si cells with optimal front grid configuration and full Al-
BSF are expected to have irradiance-independent RS with very small variation between
open circuit and MPP, mainly due to non-zero finger and contact resistances.

6.3.4 SHUNTS IN SOLAR CELLS; LOCAL J-V CURVES; MAPPING OF n AND J0


In general, the shunts in PV cells are localized formations crossing the p-n
junction and can be found over the whole cell area [Bre07a]. In particular, grain
boundaries in mc-Si cells are very probable locations for different types of shunts. Some
of them have linear I-V characteristics, but many others behave like diodes with ideality
factors sometimes exceeding 4 [Bre04, Bre07a]. A non-isolated, non-passivated edge
crossing the p-n junction is a type of recombination-induced shunt, which can behave as
170

a diode with n≈2 [McI01, Bre07a]. Good edge isolation or passivation (by e.g. a thermal
oxide) results in improved (i.e., lower) ideality factors at intermediate cell voltages
[Gre86, Cou06b]. The significance of edge shunting decreases with increasing the ratio
of the cell’s area to its perimeter [McI01]. Nine different types of shunts in mono- and
mc-Si cells were distinguished in [Bre04]. Some of these were grown-in defects of the
material, whereas the others were process-induced. Precision infrared lock-in
thermography (described in [Bre03]) may be used to locate the shunts, which can then
be analyzed using microscopic and other techniques. Contrary to the expectations, most
shunts in mc-Si cells do not originate at crystal defects like grain boundaries,
precipitates and dislocations, but are rather process-induced. The dominant grown-in
defects in mc-Si cells leading to reduced (i.e., poor) RP are needle-like, highly n-doped
SiC precipitates, which cross the whole cell along grain boundaries [Bre04, Bre07b,
Car09b, Bre10a]. These filaments act as linear (Ohmic) shunts.
Due to non-homogeneous material quality in mc-Si cells, the photogenerated
current density JL varies laterally across the cell area (usually by up to 10 %), whereas
the saturation current density J0 can vary by more than one order of magnitude [Hei05,
Tru07]. Lock-in thermography has been applied for mapping of n and J0 over the whole
cell area, and is particularly useful for localization and study of shunts with diode-like I-
V characteristics [Bre03]. This technique has revealed that in typical mc-Si cells, the
variations of injection current density related to crystal defects are due to variation of J0
rather than of n. Ideality factor imaging of a 10×10 cm2 EFG-Si cell revealed that the
area outside of the shunts had n≈1.4, whereas all local shunts had n>5. The I-V curve of
the whole cell (affected by the shunts) had n≈1.8. For mc-Si cells, ideality factor closer
to unity (e.g., 1.27) had been obtained for most of the area, thus proving domination of
“diffusion current” (low-injection SRH recombination) [Bre07a, Bre10b, Bre10c]. The
modeling approach used with this mapping technique has evolved over the past decade.
A global 1-exponential I-V model with a homogeneous ideality factor n1 (exceeding 1
in mc-Si cells) was used in [Bre10c] for the “baseline” shunt-free cell regions. For the
regions with shunts, a local 2-exponential model was used, in which I02 and n2
described the characteristic of a particular shunt. In later works, a 2-exponential J-V
model with n1≡1 and n2≡n was used and the quantities J01, J02 and n were mapped over
the whole cell area [Bre11a, Bre11b]. The 2nd exponential term was explicitly attributed
to depletion-region recombination, whereas n1≡1 was assumed to hold even in mc-Si
cells, thus neglecting the injection-level dependency of minority carrier lifetime present
in this particular material, as well as in mono-Si PV devices [Mac00, Cou06a].
Thermoreflectance imaging is another technique used for locating and characterization
of shunts in different types of PV cells [Zho11]. It is important that local ideality factor
(in terms of cell area mapping) is distinguished from the local ideality factor (in terms
of the cell’s global I-V curve) introduced later in this Chapter.
171

To summarize the last two parts of this Section, the I-V curves of a PV cell can
be affected to a different degree by local inhomogeneities such as grain boundaries and
various types of shunts, as well as by 2- and 3-dimensional effects arising from the
design of front (if any) and rear metallization.

6.3.5 FROM CELLS TO MODULES


Researchers and professionals analyze empirical current-voltage characteristics
of individual PV modules and arrays of modules in order to estimate parameters, such
as RS, and to eventually detect signs of degradation or faulty connections. Designers of
inverters and MPPT algorithms use I-V curve models of PV modules in simulation
studies [Nge09]. The power electronics community is often tempted to use complex
approaches involving adaptations of the 2-exponential cell model for PV modules and
arrays [Gow99]. The latter work also assumed n1≡1 and n2≡2 for “polycrystalline”
(supposedly mc-Si) PV cells, besides the sophisticated equations relating the other
parameters to irradiance and device temperature. The 2-exponential model with either
the first or both assumptions is still popular among researchers [San10, Dol11, Zeg11].
In [Ale11], I-V curves of a c-Si PV module were simulated assuming n1≡1 and n2=3.7.
The conductivity, cross-section and soldering points of ribbons used for c-Si cell
interconnection are major contributors to a module’s RS [Fri08]. In industrial c-Si PV
modules employing cells of size 15.6×15.6 cm2 with two busbars (H-pattern front
metallization), almost 50 % of the total series resistance was found to come from the
ribbons. The other half came from the cell (excluding the busbars, which accounted for
only 3 %). For an optimal ribbon cross-section of about 0.3 mm2, the use of three
interconnecting ribbons reduced the per-cell resistance by almost one half. Ribbons,
sub-string interconnects, the connections in the junction box(-es), as well as module’s
own cables and connectors, add extra Ohmic series resistance to the intrinsic RS of the
cells, which is supposed to mitigate any current dependency of the latter significantly.
As discussed earlier in this Section, such dependency is expected to be low for the
mainstream screen-printed c-Si cells with 3 busbars due to the shorter front-grid fingers
and the full rear-side metallization. Consequently, the series resistance of modules
employing such cells can be assumed to be practically constant along the I-V curve.
The more complex model given in Eq. (6.2) is easily adapted from a cell to a
module by multiplying the ideality factors with the number of serially-connected cells
NS in the module (assuming that all cells are practically identical). With some
exceptions (e.g., the HIT module tested in the present work), the modern PV modules
rarely have paralleled cell strings [Woy03]. Nevertheless, the adapted model is equally
applicable to modules with paralleled strings, as paralleling can easily be accounted for
by scaling up the parameters IL, I01, I02 and 1/RP. One can expect that with increasing
voltage the 1st exponential term will eventually dominate over the 2nd exponential term
and the shunt-current term. An earlier work by the present author revealed that the
172

ideality factor of the studied mc-Si modules was practically constant in the higher-
voltage part of the I-V curves, and definitely larger than unity [Yor10b]. Therefore,
fitting with a 2-exponential model with n1≡1 would result in erroneous values for the
remaining module parameters, especially RS. As already discussed in Chapter 1, the 1-
exponential I-V curve model was preferred for PV modules in e.g. the Sandia
performance model [Kin04, Pra10] and some IEC standards [IEC93, Pol11b]. The
simpler model showed a very good agreement with experimental data from most PV
module types [Mer10], with the notable exception of a-Si, which required the more
complex model given in [Mer98]. Del Cueto [Cue98, Cue99] adopted the 1-exponential
model when analyzing field data of a-Si and CdTe modules, but preferred a 2-
exponential model with n1=1 and 1<n2≤2 in the case of c-Si devices.

Figure 6.4: RS-corrected I-V curve fitted with Eq. (6.1) except the shunt term. The
higher-current part of the plot is emphasized in order to show a greater detail.

The Ohmic shunt resistance RP of a c-Si cell can be determined from the slope
of the I-V curve at large reverse bias or near short circuit where the shunt current
dominates over the exponential currents [Oue99, McI01, Mül04, Met07b]. However, the
latter region in module’s characteristics can be dominated by the bypass diodes even in
the absence of direct shadowing, see Figs. 6.4 and 3.3 [Yor10a, Yor11a]. Therefore,
interpreting this slope as a shunt conductance is incorrect. The apparent shunt resistance
173

obtained by linear fitting at the lower voltages can be much lower than the true Ohmic
RP, and may vary in an unpredictable manner depending on the operating conditions
(e.g. non-uniform illumination caused by albedo effects), see Fig. 6.5 [Yor11a]. Proper
determination of RP should therefore be done using I-V curves recorded at very low
levels of illumination, at which small mismatches between ISC of the cells in the module
become unimportant. In addition, data recorded at overcast conditions should be used in
order to avoid mismatch due to ground-reflected light or due to angular effects at AOI
close to 90°. Note that at clear-sky conditions a stable value was not reached in Fig. 6.5,
even at very low in-plane irradiances. Mermoud and Lejeune [Mer10] observed a quasi-
exponential dependency of the shunt resistance (defined therein as the inverse I-V curve
slope near short circuit) for different PV technologies, but failed to account for, or even
to mention, the bypass diodes present in the modules. Instead, they modeled the entire I-
V curves with a quasi-exponential shunt resistance, and thus perhaps overestimated the
shunt current at the higher voltages, including at the MPP. Exact determination of the
Ohmic shunt resistance of the studied PV modules is outside the scope of the present
work. For the module represented in Fig. 6.5, the value is at least a few kΩ and the
resulting shunt current is negligibly small at all practical voltages.

Figure 6.5: Apparent shunt resistance of a mc-Si module on a sunny day [Yor11a].
174

In a 15 % efficient mc-Si cell, the shunt current at the MPP is assessed to be


below 0.1 mA/cm2 at STC, which is negligibly small as compared to IM [Neu07]. The
present author showed in an earlier study that the Ohmic shunt-current term in the 1-
exponential model does not describe accurately the experimental I-V curves of mc-Si
modules [Yor11a]. Fig. 6.6 shows the average apparent shunt current for the mc-Si
module represented in the previous two figures, which was defined as the difference
between the fitted and the measured I-V curve as in Fig. 6.4. The plot was created from
20 consecutive I-V data sets recorded at irradiance of about 600 W/m2. The standard I-
V curve models given in Eqs. (6.1) and (6.2) predict that an Ohmic shunt current should
increase linearly with the RS-corrected voltage, whereas the apparent shunt current
behaves so only up to a certain voltage, well below that at the MPP. An Ohmic shunt
current would equal roughly 0.1 A at the MPP, whereas the apparent one is an order of
magnitude lower and more than two orders of magnitude below IM. For simulation
purposes, the corresponding part of the I-V curve can be modeled with a voltage-
dependent shunt-current term. The high resolution of the I-V curves recorded in the
present work, as well as in the experimental setup used in [Yor10a, Yor10b, Yor11a],
provides a solid basis for similar studies in future. However, the I-V curve analysis in
this Chapter is focused on the higher-voltage region, between the knee and open circuit.

Figure 6.6: Apparent shunt current of the module represented in Fig. 6.4 [Yor11a].
175

Individual cells within a module can have extremely low shunt resistances
[McM96, Dyk04, Ish12]. Such drastic differences between the cells in a module are not
accounted for by the 1- and 2-exponential I-V curve models adapted from Eqs. (6.1) and
(6.2) as in [Yor10b], because they assume that all cells are practically identical.
However, the analysis presented later in this Chapter is performed on the higher-voltage
parts of I-V curves recorded at reasonably high irradiances, at which shunted cells
perform well, and the shunt current is much smaller than that injected through the p-n
junction [McI01, Met07b]. Consequently, the effect of individual cells with low shunt
resistances on a module’s characteristics is expected to be very small, especially near
open circuit, from which region RS, n and I0 are determined in the present work, as well
as in [Yor10b, Yor12a].

6.4 METHODS FOR DETERMINATION OF RS, n AND I0


This Section reviews the methods used for estimation of cell and module
parameters from empirical I-V curves. The cited works are analyzed critically with
respect to the accuracy of the adopted models (if any), the assumptions made and the
physical consistency of the parameter sets obtained with the particular method. Because
this Chapter is focused on module characterization, attention is paid on the applicability
of models developed for cells in the case of module I-V curves. Module characteristics
are usually modeled with a 1-exponential equation, and therefore the parameters of
interest in the context of the present work are RS, n and I0, whereas RP is left outside of
scope for reasons explained in the previous Section.

6.4.1 CURVILINEAR FITTING OF INDIVIDUAL I-V CHARACTERISTICS


Non-linear fitting of cell I-V curves has been widely used in the literature. It
involves the assumption of a particular functional dependency relating current to
voltage and requires initial values (guesses) for the model parameters, thus being
mathematically and computationally complex. Either 1- or 2-exponential equations are
fitted in most cases. The more complex model has 5-, 6- and 7-parameter versions,
depending on the assumptions made for n1 and n2, as already discussed in Section 6.3.
An elegant way to fit empirical curves with the Saetre-Midtgård-Das model by simple
coordinate-system transformation was presented in [Sae11]. However, its parameters
are variable by definition, and are difficult to relate to the more physical ones of the 1-
and 2-exponential models focused on here. Appelbaum et al. [App93] showed that
different fitting methods with the same error tolerance could lead to widely different
parameter sets (including non-physical values such as a negative RS), while all of them
achieved satisfactory curve fitting. This was attributed to the fact that the parameters
had very different orders of magnitude, and the solution had a very flat optimum in most
of the parameters. The authors concluded that the 7-parameter 2-exponential model
performed superior to the simpler ones. The parameter consistency problem was
176

recognized also in later works [Got97, Hei05]. In the former work, the adequacy of
curvilinear fitting was studied using synthetic instead of empirical I-V curves. Fitting of
experimental curves often involves noise in the data, which may impose additional
robustness requirements and an improved algorithm [Bur96, Hei05, Che11]. In the latter
work, 4 different regions of the I-V curve were subjected separately to polynomial curve
fitting with the 1-exponential model, and the parameters were obtained by solving a set
of algebraic equations. Choosing the simplest model resulted in unrealistically high
ideality factor (above 2) and reverse saturation current for a 15.6×15.6-cm2 mc-Si cell
with a low shunt resistance and a FF of only 59 %. After laser isolation of the localized
shunt, realistic values were obtained for both parameters. Very similar values of RS
were obtained in both cases (about 2 mΩ), which are in the range typical for modern
screen-printed cells [Fri08]. Neither the 1-exponential nor a 2-exponential model
achieved satisfactory fitting when large potential differences were developed over the
cell area [Hei05]. However, n1≡1 was erroneously assumed (along with n2≡2), which
seems to have resulted in effectively fitting the 1-exponential model with n=2. The
more complete 7-parameter model was not given a try. On the other hand, the authors
adopted a more realistic ideality factor of 1.3 for the local RS-free J-V curves used to
model the global I-V characteristic of a mc-Si cell suffering from non-uniform contact
resistance. From the cell’s ISC-VOC curve, an ideality factor of 1.14 was obtained by
fitting with the 1-exponential model. The total saturation current of a cell can be
determined from fitting a reverse dark I-V curve [Wol63].
The need for a curve-fitting method suitable for PV modules operated at the
MPP, without scanning a complete I-V curve, was recognized and addressed in [Tol12].
MPPT algorithms in PV inverters operate in small vicinity around the MPP most of the
time because of the inherent current ripple of the first converter stage [Moh03b, Nge09].
The corresponding portion of the characteristic could be measured with an enabled
inverter, and then fitted. However, too simple a 1-exponential I-V curve model with
n≡1 was adopted in [Tol12], which automatically invalidates any results obtained in the
case of PV technologies with ideality factors well above unity. In addition, such a fitting
approach neglects the possibility for mismatch between the cells in the module, which is
more probable at higher currents (including the MPP region).

6.4.2 PARTIAL FITTING OF SEMI-LOG I-V CURVES


Analysis of semi-logarithmic I-V curves has been a classic way to reveal the
presence of different mechanisms in junction devices such as diodes and PV cells
[Wol63, Wol77, Kam98, Str00]. Curvilinear fitting at large currents of the semi-
logarithmic dark I-V curve can be used for determination of a constant lumped RS, but
is inapplicable when the parameter depends on current or illumination level [Abe93a].
Such dependencies were revealed for the series resistance and the ideality factor of thin-
film CdTe and CIS cells [Sit89]. Consequently, their determination from semi-
177

logarithmic dark or shifted illuminated I-V curves was preferred to techniques using
different irradiance levels (to be discussed shortly). A complementary differential
technique used together with the semi-logarithmic I-V curve analysis is the study of the
local ideality factor m (using m-V plots) introduced later in this Chapter. This artificial
parameter mixes together shunting, resistive losses and recombination currents with
different ideality factors. It is often not possible to distinguish a portion of the I-V curve
that is dominated by a single current conduction mechanism. By analysis of a synthetic
classic m-V curve, the present author demonstrated that the ideality factor of low-
injection SRH-recombination current cannot always be accurately determined from the
maximum slope of the semi-logarithmic I-V curve [Yor12a]. The I-V data was
simulated with a 2-exponential equation with n1≡1 and n2≡2, whereas the local ideality
factor reached a minimum of 1.193 at about 0.55 V. On the other hand, ideality factor
determination was found possible for mc-Si PV devices by analyzing RS-corrected I-V
data.

6.4.3 COMPARISON OF CURVES OF DIFFERENT TYPE


The assumption that both dark and illuminated I-V curves can be described by
the same mathematical model with constant parameters has led to the idea of RS
determination by comparing a dark and an illuminated characteristic (shifted by ISC)
[Mia84, Abe93a]. The measured device must be kept at the same operating temperature,
which may be challenging in the illuminated case if a high-intensity steady-state light
source is used. This method is not applicable to PV technologies with irradiance-
dependent RS [Sit89, Abe93a]. In such cases, the voltage difference between the dark
and the shifted illuminated curve at open circuit is unaffected by the illuminated series
resistance RS,LIGHT, but is entirely due to the dark series resistance RS,DARK. On the
other hand, the voltage difference corresponding to the MPP of the illuminated curve
(shifted by ISC) is almost entirely due to RS,LIGHT. This is because the corresponding
point in the dark curve is in the low-current region (IDARK=ISC–IM<0.1ISC for c-Si
devices [Ser10a, Yor10c]) and is practically unaffected by RS,DARK [Abe93a]. These
considerations were made for c-Si cells that dominated the PV market at that time,
assuming good quality devices with negligible shunting and good fill factors (implying
a rather low RS). However, in the case of thin-film devices with rather high RS,DARK, the
determination of RS,LIGHT with the above method may not be very accurate. For the
CIGS module tested in the present work, RS,LIGHT is found to range between about 1.5
and 3 Ω (see Figs. 6.7 and 6.17), whereas RS,DARK>RS,LIGHT usually holds for this
technology [Sit89]. At 1 sun, ISC–IM≈0.14ISC≈0.45 A, and so (ISC–IM)RS,DARK ≥ 0.9 V,
which is not negligible and would introduce an error in RS,LIGHT at the MPP of more
than 0.3 Ω (15 %), if the method from [Abe93a] was applied.
178

When RS,DARK and RS,LIGHT depend on current density, one may be able to
evaluate each of them by comparing (at a fixed device temperature) the corresponding
shifted I-V curve to the ISC-VOC curve, which is usually unaffected by series-resistance
effects [Abe93a, McI01]. Voltage-independent ideality factor and reverse saturation
current (in terms of the 1-exponential model) are assumed, which is not satisfied in all
types of PV. In the case of such voltage dependency, RS,DARK of a non-concentrator n+p
cell is weakly affected, whereas RS,LIGHT can be significantly underestimated at voltages
well below VOC, especially at the MPP [Abe93a, Ste11]. Deviation from linearity of the
dark semi-logarithmic I-V curve in the higher-voltage region (before the onset of RS-
related effects) is an indicator of a voltage-dependent J0 [Abe93a]. The ISC-VOC method
was found inapplicable in the case of double-sided BC cells, in which the reverse
saturation current was irradiance-dependent [McI01]. Consequently, RS,LIGHT of those
cells was determined after [Alt96] using I-V curves recorded at different irradiance
levels. The existing variants of the latter approach are reviewed in the following part.

6.4.4 COMPARING I-V CURVES TAKEN AT DIFFERENT IRRADIANCES


The original method for RS determination used two curves recorded at different
illumination levels, while maintaining the same device temperature [Wol63]. It was first
suggested by L.D. Swanson in 1960. A point was chosen in each curve that was located
at certain distance ΔI from ISC along the current axis. The voltage difference ΔV
between the corresponding points in the two curves and the difference ΔISC between the
two short-circuit currents were then used to evaluate RS=ΔV/(ΔISC). For a particular cell
with relatively high series resistance, the authors obtained different values for different
ΔI. In the given example, it is not correct to speak about a current-dependent RS(I),
because the actual dependency was on the diode current ΔI in terms of a 1-diode
lumped-circuit model. In [Alt96], the original method was discussed in the context of a
more general multi-exponential J-V curve model, where each recombination mechanism
present in a PV device was characterized by its ideality factor nk and reverse saturation
current density J0k. The authors concluded that the method would be applicable only as
long as RS, J0k and nk were the same in both J-V curves for a given ΔJ. That condition
imposed the requirement for a small ΔJSC, which was not satisfied in [Wol63].
Eventually, the method was improved in [Alt96] by limiting the light intensity
difference within 10 %, so that the effect of non-linear recombination processes could
be greatly minimized. This made it suitable for cell structures in which RS depended not
only on current density, but also on light intensity. In addition, the accuracy of the
obtained values was improved significantly by recording many samples of current and
voltage for the chosen point instead of sweeping the complete J-V curve. For the high-
efficiency c-Si cells studied therein, the uncertainty increased drastically between the
179

MPP and short circuit for illumination levels below about 2 suns. An earlier application
of the original method demonstrated that RS was strongly overestimated in the flat
region of the I-V curve [Mia84]. More generally, the current density in the dependency
RS(J) obtained with the double-light method is specified with some uncertainty that is
present at all current levels. In [Alt96], J was defined as the mean of the current
densities at the corresponding points of the two J-V curves. The associated uncertainty
is therefore equal to about a half of the difference ΔJSC between the short-circuit current
densities of the two I-V curves, i.e. up to 5 % of JSC. For example, when evaluating
RS(J) corresponding to open circuit of the J-V curve recorded at the lower illumination
level, J is set equal to ΔJSC/2, whereas ΔJSC is limited within 10 % of JSC.
The “multi light method” for determination of RS(J) presented in [Fon11a,
Fon11b] is an extension of the double light method, in which at least three J-V curves
recorded at different illumination levels are needed. It was therein applied to
interdigitated back-contact (IBC) and elongate cells, and three J-V curves were found
sufficient in most cases. In addition to the 1-sun J-V curve, two more curves were
recorded at illumination levels differing by ±10 %, but at the same cell temperature. The
use of 5 and 7 J-V curves significantly improved the accuracy at the lower cell voltages.
With the more sophisticated approaches, the characteristic voltage dependency of RS
derived in [Ara86] was observed, in which a constant value was reached at the lower
voltages, as well as at the very high voltages where the cell was operated as a load. The
ideality factor was studied locally along the J-V curve after correcting the shifted
illuminated semi-logarithmic J-V curve for the series-resistance effect. Furthermore, by
comparing the dark J-V curve to the RS-corrected shifted illuminated curve, RS,DARK(J)
was also obtained. These works clearly demonstrated the general variability of the
ideality factor and the series resistance along the J-V curve of a particular c-Si cell type,
and that the device temperature must be kept constant to within ±0.1°C in order to
achieve satisfactory accuracy with the multi light method.
The procedure for evaluation of RS of c-Si devices prescribed by the standard
IEC 60891 is also based on comparing I-V curves recorded at different illumination
levels [Mül08, Fon11b, Pol11a]. Müllejans et al. [Mül08] adapted the standard
procedure for curves recorded indoors at varying illumination typical for some solar
simulators. A recent validation study found that the results obtained with the 1st edition
of the procedure were affected by decisions left at the operator’s discretion, and thus a
rather wide range of probable values of RS could be obtained [Pol11a]. This ambiguity
originated from the procedure itself, which involved translation of the measured curves
to a certain irradiance level using a trial value of the series resistance. A trial value of
RS is considered acceptable if the maximum difference between PMAX of two translated
curves is within 0.5 %. For a “poly-Si” module, the acceptable values ranged between
180

0.7 and 1.064 Ω, which is ±20 % around the optimum RS (for which ΔPMAX=0)
specified in the more recent edition of IEC 60891 [Pol11a]. However, a wide range of
optimum values (between 0.88 and 1.3 Ω) was obtained when varying the illumination
level between 150 and 900 W/m2 at a fixed module temperature of 50°C. Example I-V
curves were not shown, nor was the number of data points in each curve specified. In
addition, details on the interpolation applied prior to PMAX estimation were not given,
and also the effects of measurement errors on the accuracy of the procedure were not
investigated. The 1st edition of the standard procedure was found to give higher
optimum values of RS [Pol11b]. In the latter study, the validation of IEC 60891 was
extended to include also various thin-film technologies. It was concluded that for the
tested “poly-Si”, CIS and CdTe modules, RS could be considered as independent on
irradiance and temperature. A double-illumination method, in which also trial values of
RS are investigated, was presented in [Lee10]. However, this method is characterized
with much greater mathematical complexity and was not extensively validated for the
PV technologies tested in the present work.

6.4.5 ALTERNATIVE METHODS USING MULTIPLE I-V CURVES


Various other methods for deriving cell or module parameters, which require
more than one I-V curve, are found in the literature. These usually require controlled
experimental conditions, such as a constant device temperature and sometimes also a
constant illumination level, which are more difficult to achieve in a reasonable time
frame with an outdoor testing setup. For this reason, such methods have not been tried
in the present work. For example, resistors with different values R can be put in series
with the PV device, thus obtaining several I-V curves of different shapes for a given
illumination level [Mia83, Mia84]. (This approach was used also in [Hyv03] for
validation of the method used therein to estimate the series resistance.) RS can then be
determined from the upper limit IL of ISC when the total series resistance reaches zero,
whereas I0 and n can be obtained from linear fitting of ln(IL–I) plotted over I(RS+R).
Because of the temperature dependency of ISC, this method also requires a fixed device
temperature. In addition, evaluation of IL is done under reverse bias, which requires I-V
measurement in two quadrants. The method assumes a 1-exponential cell I-V curve
model without a shunt-current term [Mia83]. It is therefore inapplicable to cells with
relatively small Ohmic shunt resistance, and perhaps also to PV modules, in which the
bypass diodes can affect the I-V curve’s shape at currents close to ISC.
Assuming a 1-exponential I-V curve model, Del Cueto [Cue98] introduced a
method to determine RS and n from the reciprocal open-circuit slope of multiple J-V
curves recorded at different illumination levels and a fixed device temperature. The
slope was expected to vary linearly with 1/JSC, which was true at higher irradiances for
the single-junction thin-film modules tested outdoors. In the case of c-Si devices, a
181

generally invalid 2-exponential model with n1≡1 was adopted later in [Cue99]. The
measured I-V curves were first translated to module temperature of 25°C. No details
were given about the used procedure, moreover this operation is associated with some
degree of uncertainty [Coo97]. If the same I-V curve model was used, this further
invalidates the parameter estimation method. For modules with current- and irradiance-
independent RS, this parameter could be estimated rather accurately even for a constant
n1>1 in the case when correction for temperature was not needed. This is so because RS
is the vertical-axis intercept of the fitted line, whereas the ideality factor only affects the
line slope. If RS varied along the I-V curve, the method would yield the value at zero
current. In more complex cases where both parameters varied with voltage, the method
would be totally compromised. Its application in the determination of RS and n should
be possible for some of the modules tested in the present work because of their ‘benign’
parameters, and due to the enormous number of available I-V curves. This has not been
attempted due to time constraints, but may be tried in future.

6.4.6 ALTERNATIVE METHODS USING A SINGLE I-V CURVE


Analytical solutions for the 1- and 2-exponential model parameters may offer
input values (initial guesses) for curve-fitting techniques with reasonable uncertainties
(within 10 % [Cha87]). However, the involved mathematical complexity is significant,
even in the case of the most simple 2-parameter model assuming n1≡1 and n2≡2, which
is inappropriate to the case of e.g. mc-Si PV. Therefore, using the 1-exponential model
is recommended in the case of module I-V curves, which can be affected by a greater
cell mismatch and activation of bypass diodes at the higher currents. Examples of
analytical approaches resulting in approximate values for RS (after some assumptions
are made) are given in [Mia84]. For a cell tested at high illumination level or with a
rather high series resistance, so that RSIL/(nVT)≫1 is satisfied, RS = –dV/dI at open
circuit (when the I-V curve resides in the 1st quadrant; the sign changes when it is in the
4th quadrant). For a module with NS serially connected cells, the condition becomes
RSIL/(NSnVT)≫1. The method would be applicable to a modern industrial screen-
printed mc-Si cell (with n≈1.2, delivering ISC≈IL≈8.5 A at STC) if RS≫3.6 mΩ,
whereas the typical value for such cells is just a few mΩ [Fri08]. Another method
assuming a 1-exponential I-V curve model requires RSISC≪VOC, so that the plot of –
dV/dI over 1/(ISC–I) is a straight line with a vertical-axis intercept equal to RS.
Werner [Wer88] used three differential techniques based on the dark incremental
conductance dI/dV of Schottky diodes and solar cells. Accuracy better than 1 % was
claimed for the obtained values of RS. One of these techniques was adapted to
illuminated I-V curves in [Oue99]. Integration of the total (or partial) area enclosed by
the I-V curve has been used for evaluation of RS [Ara82, Pys07]. The expression for RS
182

based on the total area A is given (after some manipulation) in Eq. (6.3) and involves
also the ideality factor n, which was assumed equal to unity in the latter work. In the
original work, the corresponding term was considered negligible at high short-circuit
currents in the context of concentrator cells. However, this may not apply to non-
concentrating devices tested under natural sunlight. The ratio VA≡A/ISC has units of
voltage and a magnitude comparable to VM, and therefore the difference VOC–VA can be
comparable to nkT/q. For a hypothetical cell with RS=0, VOC–VA=nkT/q. For a
modern industrial screen-printed c-Si cell with RS≈2 mΩ and n≈1.2 tested at STC,
VOC–VA≈1.3nkT/q (assuming ISC≈8.5 A). Note that the application of this method to
modules would require adaptation of the original formula by multiplying n with the
number NS of serially connected cells. For the Chinese-made mono-Si module studied
in the present work, the parameters given in [Yor11a] lead to VOC–VA≈1.6NSnkT/q.
Thus, it becomes clear that the area integration method is not applicable to the
mainstream screen-printed c-Si technology, since usually n is not known a priori. In
addition, minor mismatch between the cells in a module can make the I-V curve deviate
from the 1-exponential model at high currents [Yor11a].
2  A nkT 
RS   VOC    (6.3)
I SC  I SC q 
The double-light method for determination of RS was found inapplicable to PV
arrays rated several kW [Kun04]. Therefore, a simulated I-V curve was used together
with an experimental one. However, the used simulation approach seems to contradict
solar cell physics. The simulated curve had a lower ISC compared to the experimental
one, while both had the same VOC and VM (assuming equal operating temperatures).
Both curves were also assumed to have equal fill factors. The assumption of equal VOC
in both curves contradicted the 1-exponential model, according to which VOC should
vary with irradiance at a fixed device temperature [IEC93]. The assumption of
irradiance-independent FF was non-physical as well, because the authors also assumed
an invariant ratio IM/ISC. For a BP 585F module, RS = 0.4 Ω was obtained [Kun04]. An
aged module of the same type is tested in the present work (the one denoted ‘BP’), for
which RS = 0.27 Ω was obtained with the method introduced by the present author
[Yor10b, Yor11a, Yor12a]. This is much lower than what was obtained in [Kun04],
even if zero increase of RS is assumed due to the long-term outdoor degradation.
Sera et al. [Ser07] determined analytically the parameters of the 1-exponential
model using data sheet values, which included the temperature coefficients of power
and voltage. For a BP Solar BP-MSX120 module, ideality factor of about 1.4 and series
resistance of 0.47 Ω were obtained. The ideality factor value may seem plausible for a
mc-Si module [Dir12]. However, for the same module type, one of the authors obtained
183

RS=1.15 Ω at the MPP in a later work [Ser10a]. He kindly provided the present author
with several I-V curves from that module. With the method presented in [Yor10b], the
values 0.805, 0.746, 0.774 and 0.862 Ω were obtained from curves recorded at
irradiance of 610, 665, 836 and 984 W/m2, respectively. Despite the presence of some
noise near open circuit (see Fig. 3.2), the obtained values were quite similar, with a
mean of 0.797 Ω and standard deviation of 0.050 Ω (6 % relative). Being informed of
these results, the data owner double-checked his calculations and discovered that the
originally published value had been erroneous [Ser10b]. Eventually, he obtained a
corrected value of about 0.9 Ω, which was much closer to the present author’s estimate,
but still much higher than what obtained earlier for the same BP Solar model [Ser07].
The accuracy of the parameter estimation procedure used in the earlier work is thus
questionable. It is possible that the numerical algorithm used therein to solve a system
of three very complex algebraic equations had overestimated n on behalf of RS, because
both parameters together shape the I-V curve between open circuit and the MPP.
The novel method proposed by the present author in [Yor10a, Yor10b, Yor12a]
is based on the 1-exponential I-V curve model which assumes that the series resistance
is independent on the current density. At high enough levels of illumination, the higher-
voltage part of the I-V curve is practically unaffected by the shunt current in good
quality mc-Si devices [Yor11a]. Consequently, the function
 I I   I 0  V  IR S
ln  L   ln    (6.4)
 IL   I L  NSnVT
would be linear with respect to (V+IRS)/NS at the higher voltages. Thus, the vertical-
axis intercept of the fitted line would give I0. If the device’s operating temperature is
known, n can be determined from the slope of the linear region. The method does not
require knowledge of the operating temperature and can also be applied to arrays of
modules. The I-V curve must be recorded at a relatively high in-plane irradiance (at
least 0.4 - 0.6 kW/m2, depending on the particular PV technology). It is often practical
to approximate the photocurrent IL with the short-circuit current ISC. However, the cells
in a PV module are not absolutely identical in terms of their photocurrents, and this may
add some error to the estimated parameters. It has been found that 2 % error in IL results
in about 5 % error in RS. This strong sensitivity is not surprising, given that the
photocurrent appears twice in the left-hand side of Eq. (6.4). Semi-logarithmic plots of
experimental I-V data can be created with an arbitrary trial value of RS (see Fig. 6.7),
which would deviate from a straight line close to open circuit. This property of the plots
can be used to determine the true value of the parameter by minimizing the least-squares
residual of the partial linear fit (see Fig. 6.8). For a given I-V curve, a proper algorithm
can extract the value of the series resistance that results in the best linearity of the
corresponding semi-logarithmic plot.
184

The standard procedure for evaluation of RS given in IEC 60891 similarly uses
multiple trial (hypothetic, mock) values. However, it results in a rather wide range of
acceptable values (sometimes ±20 % around the mean), whereas the new method gives
a single optimum value. A similar approach was suggested recently for the next revision
of the IEC standard [Pol11a, Pol11b].

Figure 6.7: Semi-logarithmic plots created from one I-V data set of the CIGS
module, but corrected with different trial series resistance values. The curves meet
at open circuit where IRS = 0. NS is the number of serially-connected cells.

The method proposed in [Yor10b] can be easily adapted for dark I-V curves. In
a study on thermoreflectance imaging of shunts in mc-Si cells, Zhou et al. [Zho11] fitted
a cell I-V curve with “the Shockley diode equation”. The exact form of the equation was
not given, and it is not clear whether the authors had accounted for the combined series
resistance of the cell and the current probes. An ideality factor of 4.5 was obtained,
which is too high even for an EFG-Si cell [Mey04]. However, the I-V curve given in
[Zho11] looks as if it was dominated by a rather high RS. With the new method, the
present author has estimated RS ≈ 0.3 Ω (see Fig. 6.9). The corrected semi-log plot
yields I0 ≈ 12 μA and n ≈ 2.4 (presuming a cell temperature of 25°C). The so obtained
185

ideality factor is still too high for an EFG-Si cell and may result from the presence of
other shunts, which have not been accounted for. This is beyond the scope of the
example intended to demonstrate the applicability of the method also to dark I-V curves.

Figure 6.8: Least-squares residual of the linear fit as a function of the trial RS.

For the particular I-V data set represented in Fig. 6.8, the least-squares residual
has a non-zero minimum. This indicates that perfect linearity of the semi-logarithmic
plot cannot be achieved even for the optimum value of RS. Analysis of the RS-corrected
local ideality factor (using m-V plots) revealed two possible causes [Yor12a]. First,
some noise was found to be present in the I-V data; and second, an unexpected increase
of the ideality factor by about 5 % was discovered near open circuit (see Fig. 6.33
further in this Chapter). Both phenomena are possible causes of degraded accuracy of
the algorithm for determination of RS. If the partial linear fitting of semi-logarithmic
plots includes the data close to open circuit (which is an intuitive choice), perfect
linearity cannot be achieved even with completely noise-free I-V data. Therefore, the
corresponding region should be excluded from the portion of the plot that is fitted. Plots
of the local ideality factor over device current revealed that the current range to be
skipped was between 0 A to about 0.6 A (see Fig. 6.34). The latter value defined the
cut-off at lower currents. The cut-off at higher currents should be chosen so that the
curving of the semi-logarithmic plot at intermediate voltages is avoided (see Fig. 6.7). A
186

value equal to 0.5ISC worked well for the mc-Si modules tested in [Yor12a] for I-V
curves recorded at irradiances above 400 W/m2. At lower irradiances, the fitted region
of the I-V curve becomes too short for reliable fitting, until eventually the upper cut-off
becomes equal to the lower cut-off (i.e. 0.5ISC = 0.6 A). Optimization of the algorithm
for an arbitrary PV device is possible by analyzing its RS-corrected semi-logarithmic I-
V curves and the corresponding m-I plots.

Figure 6.9: Application of the author’s method for RS determination [Yor10b] to


dark I-V curve of a mc-Si cell [Zho11]. IP(V) is the exponential current through a
non-linear shunt present in the cell. The lines are guides for the eye.

The implementation of the algorithm in Matlab™ is given in Appendix D. It


starts with an initial range of trial RS values between zero and a certain upper limit. For
a particular PV device, the latter can be estimated from semi-logarithmic plots like the
ones in Fig. 6.7. The upper limit should result in over-compensation for the series-
resistance effect expressed in curving of the semi-logarithmic plot to the left in the
open-circuit region. It is very resource-demanding to evaluate the least-squares residual
for too many trial values of RS like in the case shown in Fig. 6.8. Instead, an iterative
procedure has been programmed, which does the evaluation only for eight trial values in
187

a given range. The two trial values for which the two least residuals are obtained are
then used as a new range for the next iteration. A very rapid convergence is achieved –
resolution of 0.01 mΩ requires less than ten iterations. However, the software may
occasionally enter in an infinite loop with imperfect I-V data sets, which can result in
discontinuous plots of the least-squares residuals (see Fig. 6.10). Infinite loop execution
is avoided by adding proper code that breaks the loop after cycling has been detected.
Another possible stumbling point in the algorithm’s implementation is the choice of too
high a upper limit for the initial range of trial RS values, which makes the code
terminate with an error message. Namely, the starting upper limit is device-specific.

Figure 6.10: The optimum region of a plot similar to that given in Fig. 6.8.

6.4.7 VERIFICATION OF THE NEW METHOD


The new method for determination of RS has been applied to synthetic I-V data
modeled with the 1-exponential model (see Fig. 6.11). The following parameter values
have been used for the simulation: irradiance between 200 and 1000 W/m2; module
temperature of 25°C; ISC,STC = 8.5 A; n = 1.3; RS = 0.5 Ω; I0(25°C) = 70 nA; NS = 72
cells connected in series; RP = ∞ (i.e., no shunting). Linear scaling of the short-circuit
current with irradiance has been assumed. Each curve consists of 4096 data points.
Initially, noise-free curves have been simulated with a constant current step. Then,
random noise has been added to the current and the voltage with somewhat pessimistic
188

(i.e., too high) amplitudes of 0.05 A and 0.05 V, respectively. In this way, the obtained
synthetic I-V curves resemble closely the experimentally recorded data in the higher-
voltage part, to which the method is applied. The estimated values of RS for the
different levels of illumination are given in Fig. 6.12.

Figure 6.11: Synthetic I-V curves of a c-Si module corresponding to ECT=25°C.

The method is very accurate for irradiances down to about 400 W/m 2, and yields
underestimated values of RS below that level. This is because the algorithm has been
applied for currents between 0.6 A and 0.5ISC, and this range becomes too narrow at
low irradiances. For I-V curves modeled without any noise, the method yields RS =
0.500 Ω regardless of the illumination level. In this way, noise has been identified as a
major cause for degraded method accuracy at low irradiances.
The method has also been tested with synthetic I-V data modeled with RS(I) =
0.5 + I/100 Ω, but without any noise, in order to emphasize the method’s performance
with a varying series resistance. For these curves, the algorithm has been applied with
two different current ranges (see Fig. 6.13). In both cases, the estimates obtained for the
higher irradiances strongly underestimate the true value of RS at open circuit. The
results are practically the same when the I-V data are modeled with varying module
temperature ECT = 25 + 0.035G (in Centigrade). Therefore, it has been concluded that
189

the method is generally not applicable at high irradiances, where RS varies moderately
with current. On the other hand, rather accurate estimates of the open-circuit value can
be obtained from I-V curves recorded at low irradiances. Adding noise to the I-V data
has been found to result in estimates deviating significantly from the ‘noiseless’ ones
shown in Fig. 6.13, thus making the method inapplicable also at low irradiances.

Figure 6.12: Values of RS extracted from the I-V curves shown in Fig. 6.11.

Figure 6.13: Estimation of RS from synthetic I-V data simulated with a current-
dependent series resistance. The algorithm is applied for two current ranges.
190

6.5 SERIES RESISTANCES OF THE TESTED MODULES


In this Section, results are presented and discussed for the series resistances of
the tested c-Si modules and the CIGS module estimated with the method described in
[Yor10b, Yor12a]. The limits of the method’s applicability are also considered.

6.5.1 CORRECTION FOR EXTRA CABLING AND CONNECTORS


The originally measured I-V curves are affected by additional series resistance
due to the long cables linking each module to its electronic load and also due to the
extra pair of connectors. The method used for determination of RS is applied on these
curves. Therefore, the additional resistance must be subtracted from the obtained values
in order to know the intrinsic series resistance of each module. The measurement of the
added resistances is described in Section 3.8. The obtained resistances are given in
Table VI.1 together with the corresponding uncertainties from the linear fitting (at 95 %
confidence level). Faulty connections resulting in abnormally high contact resistance
have been considered as a possible explanation of the fill-factor degradation at higher
irradiances observed for the module denoted ‘RS poly’ (see Fig. 6.1). However, the
measured value is not very different from those obtained for the other modules.

TABLE VI.1: Added series resistance due to extra cabling and connectors.
Module Added ΔRS [mΩ] U95 [%]
Aged BC mono-Si (‘BP’) 87.4 0.6
Aged mc-Si (‘GPV’) 65.7 0.75
Back-contact mono-Si (‘Sunc.’) 85.3 0.75
RSS mono-Si (‘RS mono’) 96.7 0.75
RSS mc-Si (‘RS poly’) 85.7 0.75
‘HIT’ 83.0 0.0
mc-Si ‘ESS 1’ 97.7 0.75
mc-Si ‘ESS 2’ 109.6 0.6
‘CIS’ 98.3 0.75
‘a-Si 3J’ 95.4 0.5

6.5.2 INTRINSIC RS AT LOWER CURRENTS


In Fig. 6.14, the estimates of the intrinsic RS of the tested c-Si modules are
plotted over each module’s ECT. For most of the devices, there seems to be a slight
positive temperature dependency. However, it cannot be attributed solely to module
properties because of the limited accuracy of the method, and also because the
temperature coefficient of the subtracted resistance of interconnecting leads has not
been accounted for. The estimates should be regarded as RS at lower current levels,
because the method derives them from the I-V curve part between 0.6 A and 0.5ISC.
Because of the relatively low value of ISC of the aged mc-Si module denoted ‘GPV’
191

(about 3 A at standard irradiance), only I-V curves recorded at irradiances above 600
W/m2 have been used to obtain the estimates given in Fig. 6.14. For that particular
module, a relatively small part of the I-V curve has been used to determine RS, which
may explain the large spread of the obtained values. When a fixed current range is used
that includes also the part near open circuit (where the local ideality factor is slightly
larger), the spread is significantly reduced (see Fig. 6.15), whereas the mean at a given
ECT is practically the same in both cases. The residual spread, as well as those seen in
the values obtained for the other modules, is attributed to noise in the I-V data similar to
that seen in [Yor12a], which is discussed later in Section 6.7.

Figure 6.14: Series resistances of the eight c-Si modules tested in Grimstad. The
data are from 29/04/2011.

As already discussed in Section 6.3, RS of some c-Si cell designs may increase
with current density. Therefore, this should be investigated for all the modules tested by
applying an appropriate method, e.g. the double-light method. A varying series
resistance is likely in the case of the mc-Si module denoted ‘RS poly’, because its cells
have only two front busbars (see Fig. 6.16) and therefore larger finger lengths compared
to equally-sized cells with three busbars. Furthermore, their front metallization lacks
192

redundant lines along the perimeter, which would serve as alternative current paths in
regions with interrupted fingers or with too high a contact resistance between the fingers
and the emitter layer. In this context, the relatively high RS of this module as compared
to that of the one denoted ‘RS mono’ is not surprising, although both have equal
number of cells connected in series and almost equal cell area. However, the cells in the
latter module have a front metallization with three busbars and redundant lines (see Fig.
6.3). On the other hand, the mean RS estimate of the former module is almost
independent on irradiance. In case of a significant variation of the module’s RS with
current, the obtained estimates would decrease with irradiance as shown in the previous
Section (see Fig. 6.13). However, the synthetic I-V curves used therein were modeled
assuming a uniform module temperature, which is usually not the case in outdoor
operation. In particular, the cell in the module that is in front of the junction box will
have regions operating at a higher temperature compared to that of the remaining cells.
Consequently, its I-V curve will have somewhat lower VOC and VM. However, for a
large number of cells connected in series (NS=72 for ‘RS poly’), this will have a little
effect on the module’s I-V curve.

Figure 6.15: RS values obtained for the module denoted ‘GPV’ from two different
parts of the I-V curve.

Synthetic I-V curves have also been modeled with one cell being hotter than the
others (by 16°C at standard irradiance, which is too pessimistic a worst-case scenario).
Nevertheless, the method yields practically the same RS estimates as the ones given in
Fig. 6.13. More complex temperature nonuniformities over the module’s area are
possible, but their simulation is beyond the scope of the present work. It is concluded
193

that the estimates obtained with the author’s method for the module denoted ‘RS poly’
do not suggest a significant variation of RS with current. This is because their plot is
qualitatively very similar to the plots obtained for most of the remaining c-Si modules
(see Fig. 6.14), while differing from the plot with closed symbols in Fig. 6.13.

Figure 6.16: A cell of the module denoted ‘RS poly’.

Figure 6.17: RS estimates for the CIGS module based on data from 29/04/2011. No
correction is made for the resistance of the interconnecting cables.
194

The relatively low values obtained for the module with IBC cells correspond
well with the descriptions of this technology found in the literature [Ker06, Neu07]. The
estimates obtained for the modules denoted ‘ESS 1’ and ‘ESS 2’ (with identical design)
are practically equal, which could be expected.

Figure 6.18: RS estimates of the CIGS module plotted over the back-of-module
temperature.

It has been shown already in Section 6.4 that the present author’s method for RS
determination from individual I-V curves seems to be applicable also for the CIGS
module (see Fig. 6.7). The corresponding RS-corrected semi-logarithmic plots are linear
at the higher voltages. Due to its relatively low ISC (about 3.1 A at STC), the algorithm
was applied for currents between 0 A and 0.5ISC. The estimates obtained for irradiances
between 200 and 1000 W/m2 are given in Fig. 6.17. The significant spread at irradiances
below 300 W/m2 suggests that the method becomes very inaccurate at low illumination
levels, at which the used current range becomes too narrow, and thus the noise present
in the I-V data becomes significant. In general, the obtained values decrease with
irradiance, but the hysteresis suggests the importance of the module’s temperature. In
Fig. 6.18, the estimates are plotted over the back-of-module temperature. In this plot, an
even better negative correlation between the two variables is seen. The variation with
irradiance of the estimates obtained for this module is qualitatively similar to the
variation seen in Fig. 6.13 for synthetic I-V curves modeled with a current-dependent
RS. Therefore, it is very likely that the physical RS of the tested CIGS module increases
significantly with current. This hypothesis is supported by the results obtained with the
double-light method, which are presented later in this Chapter.
195

6.5.3 PROBING FOR A CURRENT DEPENDENCY OF RS

Figure 6.19: Application of the double-light method in the open-circuit region.

The double-light method for determination of RS(I) has been applied to I-V data
of all the ten modules tested in Grimstad. The method requires two I-V curves taken at
the same module temperature but at two not too different levels of illumination. This is
done in order to avoid the effects of injection-dependent recombination and varying RS
[Alt96]. In the present work, the following approach is taken. The two curves are
chosen so that the ECT is the same within ±0.1°C, while the corresponding irradiances
differ by no more than 10 % and no less than ΔGMIN. The latter quantity is defined
empirically for a given illumination level, so that the influence of any residual noise in
the I-V data on the method’s accuracy is small. For irradiances between 900 and 1000
W/m2, an optimal ΔGMIN of about 50 W/m2 has been used, and 30 W/m2 between 450
and 500 W/m2. First, the I-V curves are smoothed using a procedure described in
Section 3.11. The diode current ID=ISC–I is then plotted over the voltage V (see Fig.
6.19). Thus, the voltage difference ΔV between the two plots at a given ID gives RS(I)=
ΔV/ΔISC where ΔISC=ISC2–ISC1 and I=(I1 + I2)/2. The implementation of the algorithm
in Matlab™ is given in Appendix E. For a given level of illumination, the method is
applied to multiple pairs of I-V curves. The obtained estimates of RS(I) are plotted in
the same axes, and are then averaged by fitting (see Fig. 6.20). No correction is made
for the added resistance of the extra cabling given in Table VI.1. When this correction is
accounted for, the average value at lower currents can be compared to the results for the
same module obtained from individual I-V curves with the author’s new method (except
196

for the a-Si module). Such a comparison is given in Table VI.2 for the c-Si devices and
the CIGS module for illumination levels around 950 W/m2. The values correspond to
module temperatures between about 40°C and 50°C. For all the modules, there is a very
good agreement between the results obtained with the two methods.

TABLE VI.2: Module RS at low currents (in Ω) obtained with two methods for
irradiances around 950 W/m2. The values exclude the resistance of additional cabling.
Module Individual I-V curves Double-light method
Aged BC mono-Si (‘BP’) 0.29 0.28
Aged mc-Si (‘GPV’) 0.32 0.32
Back-contact mono-Si (‘Sunc.’) 0.097 0.091
RSS mono-Si (‘RS mono’) 0.32 0.31
RSS mc-Si (‘RS poly’) 0.425 0.41
‘HIT’ 0.275 0.28
mc-Si ‘ESS 1’ 0.245 0.24
mc-Si ‘ESS 2’ 0.25 0.23
‘CIS’ 1.75 1.71

Figure 6.20: RS(I) of the module denoted ‘BP’ determined with the double-light
method from 61 I-V curve pairs for illumination level around 950 W/m2.
197

The presence of residual noise is evident in each individual plot given in Fig.
6.20 in black color. A qualitatively similar behavior was seen in the local ideality factor
plots presented in [Yor12a], see Fig. 6.33 further in this Chapter. The red plot in Fig.
6.20 is obtained from curvilinear fitting in Matlab™ using the ‘Smoothing Spline’
algorithm with a smoothing parameter of 0.99. Because of the definition used for I, its
minimum value equals ΔISC/2, and therefore the obtained curves for RS(I) do not reach
the open circuit. This is expected to affect the accuracy of the method, because it
actually matches data points with RS(I1) in the 1st I-V curve to points with RS(I1+ΔISC)
in the 2nd curve. For example, the 2nd curve is unaffected by RS at open circuit, whereas
the corresponding point in the 1st curve is affected by RS(I1=ΔISC). Thus, the voltage
difference between the two points is entirely due to RS(ΔISC), but the method denotes
the obtained value RS(ΔISC/2). The errors originating from this ambiguity are
considered to be small, because ΔISC is intentionally chosen much smaller than
ISC1<ISC2. The noise in the I-V data obtained with the present experimental setup seems
to cause much larger errors, which is evident both from the ripple and the spread seen in
the plots in Fig. 6.20. In addition, the method’s accuracy becomes much poorer when I
approaches ISC1, which was observed also in [Mia84, Alt96, Fon11b]. In the first work,
a sharp increase was seen in the RS values obtained for the corresponding region of the
cell’s I-V curve. A possible explanation is the usually much higher value of local
ideality factor as compared to that at higher voltages [Rob94b, Rob94c]. Similar
behavior is seen also in the results of the present work (see Figs. 6.21 and 6.22). In the
corresponding region, the module’s I-V curves may be affected by current mismatch
between the cells resulting in bypass diodes activation, and thus an increasing local
ideality factor can be expected in each cell. The result is abnormally high or abnormally
low (even negative) values of RS near short circuit. Theory predicts a more gradual
variation in the corresponding region, which was confirmed experimentally [Ara86,
Fon11b]. It is also possible that the positive trend seen in the averaged curve in Fig.
6.20, as well as the negative trend seen in Fig. 6.22, is related to the abnormal values
obtained close to short circuit. Therefore, it can be concluded that, for the present I-V
data set, the double-light method cannot reveal small variations of RS with current.
Ideally, measurement of RS(I) should be done indoors, with a very precise experimental
setup and under strictly controlled conditions.
For the CIGS module, the double-light method reveals a significant increase of
RS with current for irradiances about 950 W/m2 and ECT around 40°C (see Fig. 6.23).
The actual behavior around the MPP is obscured by artifacts typical for the short-circuit
region, as well as by the poor quality of the I-V data. For irradiances about 450 W/m2
and cell temperatures around 13°C, a much higher RS of about 3 Ω at low currents is
obtained for this module. Such significant changes are seen also for the triple-junction
198

a-Si module (1.2 Ω vs. 3.5 Ω), but not for any of the c-Si modules. Future work may
include separation of the irradiance and temperature effects on RS for the two thin-film
devices. The obtained results correspond well to the strong decrease of RS with
irradiance in CIS and CdTe cells revealed in an earlier study and hypothetically linked
to photoconductivity in the CdS window layer [Sit89].

Figure 6.21: Same as in Fig. 6.20 but on a larger scale.

Figure 6.22: RS(I) of the module denoted ‘ESS 2’ determined with the double-light
method from 40 I-V curve pairs for illumination level around 950 W/m2.
199

Figure 6.23: RS(I) of the CIGS module determined with the double-light method
from 50 I-V curve pairs for illumination level around 950 W/m2.

The local ideality factor behavior near open circuit revealed in [Yor12a] and
discussed further in this Chapter is expected to affect the double-light method’s
accuracy at the lowest current levels. Its effect would be rather small, because the
corresponding voltage increase (as compared to an ideal 1-exponential I-V curve) is in
the order of 0.01 V (to be discussed later in Section 6.7). For a mainstream c-Si module
at about 1-sun illumination and ΔISC ≈ 0.8 A, the associated error in RS(I) at open
circuit would equal barely 12.5 mΩ, or about 5 % of the intrinsic RS. Due to the very
noisy I-V data obtained in the present work, such a small variation cannot be detected
for the tested modules. The rise of the averaged RS of the module denoted ‘BP’ seen
near I = 0 A in Fig. 6.20 may have resulted from this peculiar property of the ideality
factor. However, the latter must first be studied in greater depth and with a very precise
experimental setup, which is outside the scope of the thesis. Another challenge in the
application of the double-light method to discrete I-V data is that the data points in the
two I-V curves are usually not perfectly matched in terms of ID. Therefore, one must
interpolate one of the curves when programming the algorithm. The large number of
data points per I-V curve used in the present work is beneficial also in this context,
because linear interpolation is sufficient. However, in cases of much poorer resolution
200

along either a part of or the entire curve (as in Figs. 2.4 and 3.1), curvilinear fitting may
be needed that can be more difficult to program.
The per-cell series resistances obtained in the present work for c-Si modules
made with large-area screen-printed cells range between 3.8 and 5.7 mΩ, in good
agreement with the values found in the literature [Fri08, Car09a]. Unrealistically low
value of 0.077 mΩ was obtained for a similar cell in [Bra10].

6.6 IDEALITY FACTORS DETERMINATION


As already discussed in Chapter 5, knowledge of n is needed for accurate
calculation of ECT. The standard procedure requires estimation of the ideality factor
from the suns-VOC curve of the PV device [IEC93]. This is not straight-forward in the
case of outdoor characterization, because the difference between the cell temperature
and back-of-module temperature is irradiance-dependent, even at steady conditions. An
alternative approach is to determine n from the RS-corrected I-V curve, assuming that it
is identical to that of the suns-VOC curve. This assumption is valid as long as the 1-
dimensional 1-exponential I-V curve model is applicable to the particular PV device
under consideration.

Figure 6.24: Evolution of the ideality factors of eight c-Si PV modules on a sunny
day, 29.04.2011 [Yor12d]. Data are limited to irradiances from 0.6 to 1.0 kW/m2.
201

In the present work, n is determined from the linear higher-voltage region of a


RS-corrected semi-logarithmic plot (see Fig. 6.7). Due to imperfect quality of the I-V
data, the value of RS obtained from each curve contains some error, which affects the
accuracy of n as well. The scatter plots of RS-estimates of the c-Si modules are linear
with respect to device temperature (see Figs. 6.14 and 6.15). The random error (the
scattering) is eliminated by linear fitting, and the fitted values RS(ECT) are then used to
correct the corresponding semi-logarithmic plots. Consequently, the so obtained values
of n contain very little scattering, and thus even a small variability in this parameter is
clearly revealed (see Fig. 6.24). For the CIGS module, the obtained RS varies with
irradiance and operating temperature in a complex way, and therefore a different
approach is taken. The estimates are plotted over time and are fitted with ‘Smoothing
Spline’ in Matlab™ (see Fig. 6.25).

Figure 6.25: Random error elimination from RS obtained for the CIGS module.

The irradiance profile corresponding to Fig. 6.24 is given in Fig. 6.26. However,
the ideality factor plot of the module denoted ‘RS poly’ (as well as for some other
modules) is not as smooth. The ripple seen therein has been traced to some I-V curves
affected by bypass-diode activation (see Fig. 6.27), and not to a temperature
dependency of n as initially suspected. The cause of the current mismatch between the
202

cells cannot be identified from images recorded with the TV camera, but nonuniform
reflection from the building’s roof is one possible explanation.

Figure 6.26: Irradiance profile corresponding to the ideality factors in Fig. 6.24.

The Sandia PV module database used in the NREL’s System Advisor Model
software [NRE12b, NRE12c] lists three modules denoted ‘BP Solar 585’ made between
year 2000 and 2002, with ideality factor of either 1.24 or 1.38. It is not clear whether
any of those is identical to the aged BC module tested in the present work (a BP Solar
585F model), for which notably lower values of n are obtained. This may be checked in
future together with researchers at Sandia, and a comparison of the methods used to
derive the parameter can also be done. Distributions of the ideality factors in the
database were given recently in [Pra10]. For most of the mono- and mc-Si modules, n
ranged between 1.216 and 1.49. Just one mono-Si module had an ideality factor very
close to 1. The values obtained in the present work fall in that range. Looking at the
ideality factors alone, one cannot distinguish between the different materials and cell
technologies. The results show that high-efficiency technologies do not necessarily
imply ideality factors approaching unity – the module with HIT cells and the one with
back-contact cells have ideality factors of about 1.3 or more. Values between 1.00 and
1.46 were obtained recently in [Kim10] for different commercial crystalline-Si PV
modules. Lo Brano et al. [Bra10] used a 1-exponential model where the ideality factor
incorporated the elementary charge and Boltzmann’s constant, thus having units of V/K.
By stripping q and k from their ideality-factor estimates for PV modules, classic
(dimensionless) ideality factors of 0.87 and 0.73 are obtained. These values are well
outside the range for c-Si modules in the Sandia database. A more realistic (classic)
value n = 1.287 was obtained for a c-Si cell made by Q-Cells.
203

Figure 6.27: I-V curves obtained at equal irradiances on 29 April 2011, but giving
different estimates of n.

For most of the modules, the obtained n increases with irradiance, but for the
one denoted “RS poly” the variability is particularly strong. There are several possible
causes for this variation. This behavior may be due to actual solar cell physics – caused
either by injection-dependent recombination or distributed series resistance effects, or a
combination of both. In the first case, the obtained values would equal the local ideality
factors of the ISC-VOC curve (see Fig. 5.16). In the second case, the obtained n would
depend on the current range near open circuit used in the method for estimation of RS.
Some of the tested specimens have cell structures for which a current-dependent RS is
more likely, such as the aged BC module and the IBC module. In order to mitigate the
possible artifacts from such variability, the method has been applied to a limited part of
the I-V curve close to open circuit, as discussed in the previous Section. The effect of a
possible current dependency of RS is further reduced by the rather large length of the
cables connecting each module to its electronic load, which add about 0.1 Ω to the
intrinsic series resistance. The IBC module has a comparable value of RS at low
currents, at which this parameter’s variation is relatively small [Fon11a, Fon11b].
Therefore, its change between 0 and 2 A is believed to have been almost cancelled. On
the other hand, the module denoted ‘RS poly’ has a much larger intrinsic RS, and
204

therefore the added cable resistance is too small to mitigate effectively eventual current
dependency. Such a condition seems to be present, because a notably higher n (as well
as a higher RS) is obtained at high irradiances when using a larger part of the I-V curve
as in [Yor11a], see Fig. 6.28. The two current spans become equal at irradiances of
about 600 W/m2, at which level also very similar ideality factors are obtained.

Figure 6.28: Ambiguous ideality factor results obtained for the module ‘RS poly’.

Figure 6.29: Values of n obtained for the aged BC module.


205

Figure 6.30: Ideality factor values obtained for the CIGS module.

On the other hand, rising of RS with current would cause its underestimation at
the higher irradiances (see Fig. 6.13), which is not seen in the values obtained for any of
the studied modules. Therefore, the variation of n of the module denoted ‘RS poly’ is
not entirely due to significant distributed-RS effects. For the other screen-printed
modules and for the BC module, the estimates of n obtained with different current
ranges are either practically the same or differ by a few per cent at all irradiances. The
positive correlation with irradiance is otherwise preserved (see Fig. 6.29).
The open-circuit ideality factor formulation given by Green [Gre03] allows the
general variability of this parameter. In Green’s Eq. (18), the numerator equals the
“raw” value (here denoted nRAW) of the ideality factor of the dominant recombination
mechanism in the PV cells at open circuit. For SRH recombination in the bulk and at
surfaces (which dominates non-concentrating c-Si cells under practical operating
conditions), nRAW = 1. This allows one to simplify Eq. (18) in [Gre03] to the form given
in Eq. (6.5), where the quantity A(V) accounts for the voltage dependency of the reverse
saturation current. The detailed derivation is given in Appendix F.
n RAW 1
n  (6.5)
kT  1 dA   1 dA 
1   n RAW 1  VT  
q  A dV   A dV 
The ideality factor exceeds unity virtually for all commercial non-concentrating
c-Si PV modules [Pra10]. Such values suggest that for practical c-Si cells, the sign of
206

the derivative dA/dV is always negative. For example, if n≈1.45, from Eq. (6.5) one
obtains dA/dV ≈ –0.31A/VT corresponding to a 31 % increase in the reverse saturation
current for every 25.7 mV decrease in the cell voltage (at standard cell temperature). A
constant n along the I-V curve would mean a perfectly exponential voltage dependency
of the reverse saturation current, which is not necessarily the case for all cell structures
[Mac00, McI01]. However, the present work is focused on the possibility for significant
variation of the open-circuit ideality factor with irradiance, like the one obtained for the
mc-Si module “RS poly”. In mc-Si cells, recombination depends on carrier injection
level, which is affected not only by the operating voltage, but also by the level of
illumination [Abe93a]. Irradiance dependency of the reverse saturation current is not
covered by the general formalism used in [Gre03] to model an arbitrary device’s I-V
curves. Nevertheless, such a dependency is expected to affect the magnitude of A(V) in
Eq. (6.5) at different levels of illumination, which would result in an irradiance-
dependent n. In thin-film PV cells, the depletion region thickness is comparable to that
of the cell, and so junction recombination cannot be neglected. Therefore, Eq. (6.5) is
not valid for these technologies and one should use the more complex Eq. (18) in
[Gre03]. Larger ideality factors can be expected in thin-film devices as compared to c-Si
ones. The values of n estimated for the CIGS module for irradiances down to 200 W/m2
are given in Fig. 6.30. The red line therein is obtained from linear fitting of the values
above 800 W/m2, and the corresponding coefficients are used in the ECT calculation
with the variable n model given in Eq. (5.13). The correction needed at lower
irradiances is applied directly on ECT instead of on n, as described in Section 5.4. The
values are similar to the ones obtained in [Sit89] for a similar technology (between 1.6
and 2 at irradiances between 0 and 1 sun and device temperature of 300 K). In that
work, the parameter showed signs of saturation at higher irradiances and decreased at
higher cell temperatures, reaching nearly constant values above 350 K.
1 n n  1
 (6.6)
n T T
The anticipated effect of the device’s operating temperature on n can also be
evaluated from Eq. (6.5). The associated relative temperature coefficient is given in Eq.
(6.6) and is always positive for n>1. It is expected that the higher an ideality factor
value is from unity, the greater its temperature dependency. However, the magnitude of
the temperature coefficient is very small. For n=1.45 and a realistic T=320 K, Eq. (6.6)
gives only about 0.14 %/K. Such variations are much lower than the method’s accuracy,
and cannot be detected with the present experimental setup in the case of c-Si modules.
For thin-film devices, a relationship similar to Eq. (6.6) can be obtained, in which n is
scaled down by the factor <nRAW> (that equals the numerator of Eq. (18) in [Gre03],
namely the weighted raw ideality factors of all recombination mechanisms). Therefore,
the resulting temperature coefficient of n is similar in magnitude to that predicted for c-
Si devices, i.e. very small.
207

It is also possible that a systematic, irradiance-dependent error of n is causing


the observed variation. The parameter’s evaluation from RS-corrected semi-logarithmic
plots involves ECT, while the calculation of the latter involves n. At standard
irradiance, an erroneous n would cause a zero error in ECT, which follows from Eq.
(5.8a). The accuracy of the corresponding estimate of n is affected by the uncertainties
of ECT, RS and ISC, but not from an error in n affecting ECT. At other levels of
illumination, an error in n would act on itself through ECT, thus increasing its total
uncertainty. The feedback is positive and at 600 W/m2 it produces an additional error in
n of only ±0.2-0.3 % per ±10 % base (intrinsic) error. The feedback is so weak because
the effect on ECT is very small at the specified irradiance – within ±1°C for the tested
c-Si modules, which was obtained using Eq. (5.8b). The error in n due to an erroneous
ISC (acting via RS) is small compared to the observed variation, see Fig. 6.27 and the
lower plot in Fig. 6.28. Furthermore, this error is not systematic, and is therefore ruled
out. (Perhaps it could be avoided by linear extrapolation of the I-V curve at higher
voltages, well beyond the kink due to bypass diode activation, see the red plot in Fig.
6.27.). The noise present in the I-V data has a random and not systematic nature; it is
believed to have been almost eliminated by the linear fitting of the obtained RS, and is
thus ruled out as well. Other possible sources of error in n include its local behavior
near and including open circuit (see Figs. 6.33 and 6.34 in the next Section), which is
systematic, but does not seem to vary with irradiance. This will be discussed in the
following Section of the thesis. The error in ECT due to an erroneous β given by Eq.
(5.9) is systematic, as well as irradiance-dependent (see Fig. 5.12), and the uncertainty
of the latter for the module denoted ‘RS poly’ is among the largest seen in Table V.2
(the two estimates differ by about 8 %). The maximum ΔECT on the day represented in
Figs. 6.24 and 6.28 is about –1.2° when assuming Δβ/β ≈ +4 %, resulting in Δn/n ≈
+0.37 % at irradiance of about 1000 W/m2 and an even lower value at 600 W/m2. This
is too small and cannot explain the variation of n for that module, which exceeds 6 %.
Temperature nonuniformity over a module’s area can be both systematic and
irradiance dependent. At high irradiances, a cell located in front of the junction box will
normally operate at higher temperatures than the other cells, whereas cells adjacent to
the frame will be cooler (especially the ones at the corners). Consequently, the module’s
I-V curve will be a superposition (in terms of voltage for cells connected in series) of
the individual I-V curves of all cells (assuming no bypass-diode activation) affected by
the strong temperature dependency of their voltages. Such differences are not expected
at low irradiances. However, thermal imaging of the studied modules has not revealed
notable differences in the temperature distribution between the module denoted ‘RS
poly’ and the other modules with similar design (e.g. the one denoted ‘ESS 1’, see Fig.
6.31). Thus, nonuniform temperature is ruled out as a main cause for the larger variation
of n of the former module. The varying n seems to be an intrinsic property of the
208

particular device, and is therefore assumed to be present also in its ISC-VOC plot, which
is the basis for calculation of ECT.

Figure 6.31: Thermal images of the module denoted ‘RS poly’ (left) and the one
denoted ‘ESS 1’ (right, the module in the upper right part of the image).

TABLE VI.3: Coefficients for the linear ideality factor model given in Eq. (5.13)
obtained for the tested PV devices.
Module n0 [ - ] a  103 [W-1m2]
Aged BC mono-Si (‘BP’) 1.138(3) 0.091(3)
Aged mc-Si (‘GPV’) 1.325(4) 0.071(3)
Back-contact mono-Si (‘Sunc.’) 1.307(8) 0.045(5)
RSS mono-Si (‘RS mono’) 1.193(4) 0.065(3)
RSS mc-Si (‘RS poly’) 1.43(1) 0.264(7)
‘HIT’ 1.29(2) –0.11(1)
mc-Si ‘ESS 1’ 1.182(4) 0.045(3)
mc-Si ‘ESS 2’ 1.165(4) 0.070(4)
‘CIS’ 1.94(1) 0.44(1)

The ideality factors obtained in the present work are used to obtain the
coefficients n0 and a in Eq. (5.13) from linear fitting in Matlab™. Their values are
given in Table VI.3. The uncertainties in the brackets are for 95 % confidence level. The
present accuracy of n is too low to also allow estimation of the temperature coefficient
b. According to the analysis of Eq. (6.6) done earlier in this Section, minimal
temperature dependency of n can be expected for c-Si PV, as well as for thin-film
devices. Therefore, b appearing in Eq. (5.13) is set to zero. The corresponding effect on
ECT is expected to be very small for most of the irradiances considered in the present
209

work. For most of the c-Si modules, the observed variation of n with irradiance is very
small, and ECT can be calculated quite accurately using the standard procedure with a
constant ideality factor. This is not the case for the module denoted ‘RS poly’, for which
the variation is notably stronger.

6.7 LOCAL IDEALITY FACTOR ANALYSIS


Werner [Wer88] used a 1-exponential model for parameter determination from
individual I-V curves of Schottky diodes and solar cells. He showed that a numerical
agreement between fitted and measured I-V curves was insufficient to prove the
physical validity of the assumed current transport models. In addition, accurate
prediction of the incremental conductance dI/dV had to be achieved. The conductance
plot of a surface-damaged diode given in that work departed significantly from the
theory which predicted voltage-independent n and RS. On the other hand, the
corresponding semi-log I-V plot showed a very good numerical agreement with the
theoretical curve. Thus, the conductance plot revealed the failure of the assumed model,
which was not obvious in the standard I-V plot. In the present work, another type of
differential technique is applied in the analysis of experimental I-V data.
The analysis of the local ideality factor m (using m-V plots, to be defined
shortly) is a differential technique which facilitates the characterization of various
physical mechanisms in solar cells [Rob94b, McI99, Mac00, McI01, Alt02, Bre06,
Cou06a, Bre10b, Fon11a, Fon11b]. The ideality factor of a junction is proportional to
the reciprocal slope dV/d(lnI) of the dark semi-log I-V curve in its linear region
[Sho61]. However, the curve may lack a clearly defined linear region and can contain
kinks or humps (shoulders). The classic local ideality factor m(V) is a generalization of
the constant ideality factor n as defined by the standard 1-exponential model. It can be
defined for dark I-V curves, ISC-VOC curves and ISC-shifted illuminated I-V curves of an
individual p-n junction device as follows [Rob94b]:
 V 
I  V   I0 exp   (6.7)
 m  V  VT 
where VT = kT/q is the thermal voltage; k is Boltzmann’s constant; T is the absolute
cell operating temperature; and q is the elementary charge. For dark I-V curves, m is
proportional to the reciprocal slope of the tangent to the semi-log curve [McI01]:
1  dV  I  dV 
mV      (6.8)
VT  d  ln I   VT  dI 
In the case of illuminated I-V curves of solar cells, the above equation must be
modified to the following form [Yor10b]:

mV 
 ISC  I   dV   I  I SC   dV 
     (6.9)
VT  d  ISC  I   VT  dI 
210

For I-V curves of PV modules, Eqs. (6.7) through (6.9) must be adapted by
accounting for the number NS of cells connected in series. In Eq. (6.7), the denominator
in the exponent must be multiplied by NS, whereas Eq. (6.9) is modified to the form

mV 
 I  ISC   dV 
  (6.10)
NS VT  dI 

Figure 6.32: Classic versus RS-corrected local ideality factor (m-V plots).

The classic local ideality factor (i.e., the m-V plot) of a mc-Si module for the
intermediate and higher voltages is shown in Fig. 6.32 with closed red symbols.
Shunting, edge recombination, surface-passivation charges and other mechanisms in
solar cells lead to increase of m at low and moderate cell voltages [Rob94b, McI01,
Alt02]. In addition, I-V curves of PV modules are often affected by bypass-diode
activation at low and intermediate voltages, because minor current mismatch between
the cells is usually present even in the absence of shadowing [Yor10a] (see also Fig. 3.3
in Chapter 3). Consequently, high local ideality factor values are calculated, which
appear as humps in the corresponding part of the m-V plot (see Fig. 6.32). At the higher
voltages, the classic local ideality factor is increased due to the series resistance effect,
which makes more difficult the analysis of features appearing as bumps in this part of
the m-V plot (e.g. Fig. 7b in [McI99]). Although RS-corrected m-V plots have existed
211

and continue to appear in the literature [Bre06, Fon11a, Fon11b, Yor12a], the
mathematical formulation of a RS-corrected local ideality factor was introduced for the
first time by the present author in [Yor10b]. It reads:

mSR  V  
 ISC  I   d  V  IRS     I  ISC   d  V  IRS  
    (6.11)
NS VT  d  I SC  I   NS VT  dI 
Eq. (6.11) applies to illuminated module I-V curves. For a single cell, one needs
to set NS=1. The formulation applies also to cell structures with a current-dependent
series resistance RS(I) like the ones studied in [Abe93a, Alt96, Fon11b]. The open
symbols in Fig. 6.32 represent the RS-corrected m-V plot created from the same I-V
curve used to obtain the classic m-V plot. With the series-resistance effect removed, it
becomes clear that the corrected local ideality factor is essentially constant over a range
of the higher voltages. In [Yor12a], the present author derived a proof that the
corresponding value should equal the ideality factor n1 in the case if a 2-exponential I-V
curve model had been adopted.

Figure 6.33: Behavior of RS-corrected m-V plots of a mc-Si module at the higher
voltages. 40 I-V curves recorded consecutively at almost the same irradiance are
represented in various colors. The thicker black line is obtained by averaging of
the corresponding m-V plots [Yor12a].

Unexpected feature of mSR near open circuit is the slight increase by 5-9 %
compared to the base value (see Fig. 6.33). By comparison of multiple RS-corrected m-
V plots viewed on a smaller scale, one is also able to see whether they are affected by
noise, which otherwise may not be distinguishable in the corresponding I-V curves.
212

Averaging of many plots eventually cancels the random variability (see the thicker
black line in Fig. 6.33) and reveals that mSR is practically constant at the higher voltages
(as expected), but in the open-circuit region. If the observed rise is not an artifact either
from the equipment or from the data processing, then it must have originated from PV
device physics. Several potential causes of artifacts have been ruled out in [Yor12a],
such as the turn-on characteristic of the electronic load and the inductance of the long
interconnecting cables. Rise in mSR near open circuit was also seen in the case of a
single cell (of the same technology and make), the I-V curves of which were recorded
with a completely different experimental setup. Later in this Section, this phenomenon
is discussed in detail also for the modules tested in Grimstad.

Figure 6.34: The RS-corrected m-I plot corresponding to the thick line in Fig. 6.33.

The RS-corrected local ideality factor can be plotted also over the device current.
In the RS-corrected m-I plots (introduced for the first time by the present author in
[Yor12a]), the region of almost constant mSR can stretch over a range of several Amps,
depending on the value of ISC (see Fig. 6.34). The novel plots reveal also the portion of
the I-V data set that should be used for determination of the series resistance by the
method introduced in [Yor10b], namely the region with a relatively constant mSR. In
addition, they allow one to see how the ideality factor behaves near open circuit. To the
author’s knowledge, m-I plots have not been used previously in the literature.
213

Analysis of m-V plots at the low and intermediate voltages may be useful for
detection and rejection of ‘imperfect’ I-V curves (with ‘kinks’ caused by bypass diodes,
see Figs. 3.3 and 6.27) during automated processing of hundreds of curves. However,
this is beyond the scope of the thesis and is left for future work. The focus of this
Section is on the higher-voltage part of the RS-corrected m-V and m-I plots, in
particular the behavior of the local ideality factor mSR at the MPP, as well as near open
circuit. For the c-Si modules and the CIGS module tested in Grimstad, RS-corrected m-I
plots have been created from 26 I-V data sets recorded consecutively on 29 April, 2011
at irradiance of about 1000 W/m2 (at the peak of the profile given in Fig. 6.26), see Figs.
6.35 through 6.43. The higher-current part of the plots is intentionally skipped in order
to emphasize the flat region, the MPP region, as well as the open-circuit region.

6.7.1 IDEALITY FACTOR BETWEEN OPEN CIRCUIT AND THE MPP

Figure 6.35: RS-corrected m-I plots of the aged mc-Si module denoted ‘GPV’.

The ideality factor increase near open circuit revealed in [Yor12a] is seen for all
c-Si modules tested in the present work, and to a much smaller degree for the CIGS
module. The strongest change at open circuit with respect to the basic value (by about 9
%) is obtained for the aged mc-Si module, the aged BC mono-Si module, and the IBC
mono-Si module. For the remaining c-Si modules, the change is within 3.5 to 5 %. The
former three modules have much smaller numbers of series-connected cells (36 in the
two aged devices and 32 in the IBC module), whereas for the latter ones NS equals
either 60 or 72. On the other hand, there are as many as 104 cells in the CIGS module,
for which the ideality factor changes by as little as about 1 %. This behavior is
214

obviously not a part of the noise present in the I-V curves, because the m-I plots of
some modules are almost noise-free (see Figs. 6.35, 6.40 and 6.42).

Figure 6.36: RS-corrected m-I plots of the module denoted ‘RS mono’.

Figure 6.37: RS-corrected m-I plots of the module denoted ‘ESS 1’.

For one of the mc-Si modules studied in [Yor12a], an alternative method for
determination of RS and n has been tested, namely plot “type C” in [Wer88] modified
for illuminated PV devices (see Fig. 6.44). For an I-V curve described by the 1-
215

exponential model, a straight line would be expected, with a slope equal to –RS and a
vertical-axis intercept proportional to n. For practical I-V data, only a partial linearity
has been obtained, whereas notable curving has been observed near the open-circuit
region (for the highest values of the diode current ID in Fig. 6.44). On the other hand,
the alternative method has yielded practically the same results for the two parameters as
the ones obtained with the method developed by the present author.

Figure 6.38: RS-corrected m-I plots of the module denoted ‘ESS 2’.

Figure 6.39: RS-corrected m-I plots of the module denoted ‘RS poly’.
216

Figure 6.40: RS-corrected m-I plots of the aged BC mono-Si module denoted ‘BP’.

Figure 6.41: RS-corrected m-I plots of the module denoted ‘HIT’.

Asymmetric SRH recombination takes place when the majority-carrier lifetime


is significantly larger than the minority-carrier lifetime. Since the SRH lifetime is
dependent on excess carrier concentration, the SRH recombination in p-type bulk
changes from τn dominated to τn+ τp dominated as the applied voltage is increased. If τn
is much smaller than τp, there will be a “bump” in the m-V plot [Cou06a]. The ratio of
the two lifetimes is defect-specific [Mac00, Cou06a]. In the latter work, m-V plots were
217

simulated only for one set of cell parameters, at which the bump peaked at m ≈ 2 for V
≈ 0.65 V. However, the effects of varying e.g. the cell temperature on the bump’s
location and magnitude were not explored. Analysis of the effects of varying a cell’s
parameter on the resulting bump in the m-V plot was performed in [McI99]. In floating-
junction-passivated c-Si cells, bumps at a wide range of cell voltages were found
possible. The increase of the local ideality factor observed in the present work always
takes place close to open circuit. However, the current range of the region of interest
differs between the modules. Extending the I-V measurements beyond VOC (i.e.
operating the PV modules in two quadrants) is a possible scope of future studies focused
on characterizing this phenomenon over a broader voltage range. The rise may have
resulted from a current-dependent series resistance (larger by about 3 % at open circuit
for the modules studied in [Yor12a]). Another possible cause for this phenomenon is the
low-current characteristic of the p+p high-low junction in the rear part of a n+p cell (the
aluminum back-surface field, Al BSF [Abe00]). This junction is in series with the cell’s
p-n junction, and the voltages across both junctions are superimposed. For cells made
with n-doped wafers, front- and back-surface fields can be implemented by creating n+n
high-low junctions. Detailed investigation of the high-low junction hypothesis is beyond
the scope of the thesis.

Figure 6.42: RS-corrected m-I plots of the IBC mono-Si module denoted ‘Sunc.’.

The rise of the ideality factor with respect to the basic value corresponds to a
very small increase of VOC compared to an I-V curve following precisely the 1-
exponential model (only 0.2 % in the case of the module denoted ‘RS poly’, see Fig.
6.45). This deviation is insignificant for the maximum power output, but is relevant to
the ECT calculation (it translates to a temperature difference of about –0.6°C in the
218

case shown in Fig. 6.45). For the three modules for which the ideality factor rises by
about 9 % at open circuit, the deviation of VOC from the value expected from the 1-
exponential model may be even higher in relative terms (as well as the associated
difference in ECT). The investigation of these details is left for future work. The
comparison of empirical I-V curves to their best fit with the 1-exponential model as in
Fig. 5.38 shows that the rise of the local ideality factor near open circuit is not an
artifact of the data processing implemented in Matlab™.

Figure 6.43: RS-corrected m-I plots of the module denoted ‘CIS’.

Figure 6.44: Estimation of RS and n after [Wer88] (adapted for illuminated PV).
219

Figure 6.45: I-V curve deviation from the 1-exponential model near open circuit.

Figure 6.46: Indoor I-V curve measurement of a one-cell mc-Si mini-module.

The observed ideality factor behavior near open circuit may be an artifact from
the measurement equipment. The I-V curves are swept starting from open circuit, where
the MOSFET transistors of the electronic loads are in the off-state. The transition
between off- and on-state often results in some discontinuity at open circuit (see Figs.
3.18, 3.33 and 6.45). In order to test this possibility, I-V curves of a one-cell mc-Si
mini-module have been measured. A very simplified experimental setup with a
220

capacitive load and manual switching between short-circuit and open-circuit condition
was used (see Fig. 6.46). The voltage is sensed at the module’s terminals, whereas the
current is estimated from the voltage drop across a current probe connected in series
with the module and the load. The current probe consists of five identical high-power (5
W) resistors rated 47 mΩ, which are soldered together in parallel configuration, thus
totaling 9.4 mΩ with a joint power dissipation capability of 25 W. In addition, the
paralleling reduces significantly the uncertainty of the probe’s rated resistance. The
mini-module is illuminated with natural sunlight coming through a window. I-V curves
have been swept using a NI PCI6221 data acquisition board and a simple LabVIEW-
based code, starting close to short circuit and ending at open circuit (see Fig. 6.47). The
obtained RS-corrected m-I plots show the same behavior near open circuit (see Fig.
6.48) as those of the modules studied in [Yor12a] and the ones represented in Figs. 6.35
through 6.42. Therefore, it is concluded that the observed phenomenon is not caused by
the particular type of electronic load. Future work may include analysis of I-V curves
and m-I plots with much broader current range. This will require a more sophisticated
circuitry with a much lower combined on-state resistance, including a current probe in
the order of 1 mΩ and a semiconductor switch with a gate/base-driver subcircuit. The
switch should preferably be a minority-carrier device (such as an IGBT), which would
have a much lower on-state resistance due to conductivity modulation compared to
majority-carrier device of a similar voltage rating (such as a MOSFET) [Moh03c].

Figure 6.47: I-V curve of the mini-module shown in Fig. 6.46.

Inductive effects of long interconnecting cables have also been ruled out as a
possible cause for the peculiar ideality factor behavior near open circuit. This is because
in the present outdoor test setup, as well as in the similar one used in [Yor12a], the
temporal derivative of the current dI/dt is practically constant along a module’s I-V
221

curve. Nevertheless, such possibility has been investigated by increasing the I-V curve
sweep time up to 3 seconds. This ten-fold decrease in dI/dt has not affected the
magnitude of the change of mSR (see Fig. 6.49). In the case of the mini-module, dI/dt
has reached almost zero near open circuit due to the capacitive nature of the load,
whereas the interconnecting cables have been almost two orders of magnitude shorter.
Nevertheless, the rise of mSR has a similar magnitude to that obtained in [Yor12a].

Figure 6.48: RS-corrected m-I plot obtained from the I-V curve given in Fig. 6.47.

Figure 6.49: RS-corrected m-V plots from I-V curves taken at various sweep times.
222

Figure 6.50: I-V curve of a mc-Si cell measured indoors with a pulsed solar
simulator, before and after correction for varying illumination level. Data are used
by courtesy of Muhammad Tayyib (with the University of Agder).

Figure 6.51: The m-I plots corresponding to the m-V plots given in Fig. 6.32.

The rise of the local ideality factor near open circuit observed in the present
work and in [Yor12a] seems to have remained undetected so far even for mainstream c-
Si PV cells with screen-printed front-side metallization and full Al BSF at the rear side.
One possible explanation is that the PV industry does high-throughput in-line cell
223

testing using pulsed solar simulators before sorting the cells in bins according to their
key parameters. Correction procedures are applied to the raw I-V curves measured
under varying illumination level, e.g. using the standard procedure prescribed in IEC
60981 [Mül04, Mül08]. An example of such curves is given in Fig. 6.50. The number of
data points (in the order of 100) is too small for a detailed analysis by means of
differential techniques. For accurate estimation of PMAX, the resolution is best near the
MPP and is much lower near open circuit. Furthermore, the correction introduces some
error to the final I-V curve, which is thus not suitable for accurate testing of theoretical
models. On the other hand, the I-V curves recorded in the present work contain a few
thousands of data points (2000 in the one given in Fig. 6.47) measured using natural
sunlight under clear-sky conditions, which offer very stable illumination.
From the classic m-V plot in Fig. 6.32 one can deduce that the effect of RS
diminishes at intermediate and low voltages, i.e. at higher currents. This means that
incomplete compensation for RS in RS-corrected m-V and m-I plots would have a very
small effect in the MPP region (see Fig. 6.51). Therefore, the rise of mSR near the MPP
in the m-I plots given in Figs. 6.35 through 6.42 is mainly due to the ideality factor of
the individual cells in each module or due to current mismatch between them, and not
due to eventual current dependency of RS. Note that the discrepancy between the
corrected and the uncorrected m-I plot in Fig. 6.51 is due to the entire RS. The RS-
corrected m-I plots given in Figs. 6.35 through 6.42 are created using the fixed values of
RS (including the contribution of additional cabling and connectors) obtained with the
method introduced in [Yor10b]. In case a module’s RS rises with current, the resulting
overestimation of mSR at the MPP would be very small, being caused by the difference
RS(IM)–RS(I≈0), and not by the entire RS(IM). This is especially true for the studied c-Si
modules, for which no significant current dependency of RS has been deduced. For the
CIGS module, RS has been found to rise with current, but the corrected m-I plots show a
practically constant ideality factor between open circuit and the MPP (see Fig. 6.43).

6.7.2 ESTIMATION OF n AND RS FROM CLASSIC LOCAL IDEALITY FACTOR


The classic local ideality factor in Fig. 6.51 varies almost linearly with current.
One may want to fit the linear regions in both plots and see at what current level the two
lines cross each other. Fig. 6.52 confirms the intuitive guess that the crossing point is
exactly at ISC. This property of the classic local ideality factor is in fact a new method
for determination of a PV device’s ideality factor (namely, the value of the partial linear
fit at short circuit, 1.1883 in the given example). Furthermore, the negated and
multiplied by NSVT slope of the line fitted to the classic local ideality factor equals
exactly the estimate of RS obtained with the author’s earlier method (0.440 Ω in the
224

given example). Thus, the non-compensated m-I plot is shown to be very beneficial (in
the case of a current-independent RS), which is another contribution of the thesis.

Figure 6.52: Partial linear fitting of the m-I plots given in Fig. 6.51.

6.8 REVERSE SATURATION CURRENTS

Figure 6.53: Reverse saturation currents of the tested modules.

The estimates of I0 for the eight c-Si modules and the CIGS module tested in
Grimstad are given in Fig. 6.53. The results are based on I-V data from 29 April, 2011
225

and correspond to the ideality factors given in Figs. 6.24 and 6.30. For the CIGS device,
the obtained values are several orders of magnitude higher compared to those of most c-
Si modules. This could be expected due to the much poorer crystallographic quality of
the absorber layer in the former. Most of the semi-logarithmic plots vary practically
linearly with the operating temperature, and the slopes are very similar for most of the
c-Si devices. For the modules denoted ‘RS poly’, ‘Sunc.’ and ‘BP’, a hysteresis is seen,
which can be attributed either to curvilinear semi-logarithmic suns-VOC plots (resulting
in irradiance-dependent n and I0) or to a poor accuracy of the method.
In the available literature, rather complex approaches have been taken when
modeling the temperature dependency of I0, involving expressions of the form given in
Eq. (6.12a), where B (not to be confused with the quantity used in Eq. (5.15)), γ and λ
are constant for a given PV device [Fan86, Coo97, Gow99, Str00, Gre03, Luq03b,
Mer10, Yor12a]. Theory predicts (under some constraints) that γ = 3/n [Fan86, Gre03],
whereas λ = n when the bandgap energy EG is defined at T = 0 K [Gre03]. However, a
much simpler empirical approach is proposed in the present work in order to make I-V
curve modeling easier. For example, it is not clear what value should be chosen for EG
in the case of heterojunction PV such as the HIT module, as well as in the case of CIGS
thin films, in which the bandgap may be tuned over a wide range [Sie10]. Therefore, the
obtained temperature dependencies are modeled with Eq. (6.12b), where α is assumed
constant in the given temperature range. The values of I0(25°C) and α are determined
for each module from linear fitting of the corresponding semi-logarithmic plot with
respect to ECT. The results are listed in Table VI.4.

 E 
I0  T   BT exp   G  (6.12a)
  kT 

I0  ECT   I0  25°C exp   ECT  25°C (6.12b)

TABLE VI.4: Fitted coefficients for I0(ECT).


Module I0(25°C) [nA] α [1/°C]
Aged BC mono-Si (‘BP’) 2.95 0.146
Aged mc-Si (‘GPV’) 47 0.127
Back-contact mono-Si (‘Sunc.’) 7.9 0.115
RSS mono-Si (‘RS mono’) 9.0 0.141
RSS mc-Si (‘RS poly’) 97 0.142
‘HIT’ 6.8 0.067
mc-Si ‘ESS 1’ 8.5 0.139
mc-Si ‘ESS 2’ 4.6 0.142
‘CIS’ 2.1×104 0.109
226

For most of the c-Si modules, I0(25°C) is between about 3 and 9 nA, except for
the mc-Si ones denoted ‘GPV’ and ‘RS poly’, for which it is an order of magnitude
higher. It is important to note here that the area of the cells differs between the modules
[Yor11a], and ideally I0(25°C) should be given in units A/cm2. However, this is not
always straight-forward, because the mono-Si devices denoted ‘BP’, ‘RS mono’ and
‘Sunc.’ have pseudo-square cells (see Figs. 3.7 and 6.3), whereas the cells in the HIT
module have a more complex geometry. The cells in the CIGS module are just a few
millimeters wide (but are otherwise longer than one meter). Therefore, the estimation of
a cell’s area would involve a large uncertainty. The values of I0(25°C) obtained for the
mc-Si modules based on Elkem Solar Si (the ones denoted ‘ESS 1’ and ‘ESS 2’) are
better (i.e. lower) than that of ‘RS mono’. Surprisingly, the estimate for ‘ESS 2’ is about
half that obtained for ‘ESS 1’. The magnitude of I0(25°C) is proportional to B in Eq.
(6.12a) which differs between the modules, whereas α characterizes the temperature
dependency of I0 in the considered temperature interval. For five of the c-Si modules,
very similar values of α are obtained. However, one should not generalize that modules
made with screen-printed c-Si cells have practically the same temperature dependency
of I0, because the module denoted ‘BP’ is made with buried-contact cells, and also
because α is notably lower for the aged mc-Si module. In addition, the value obtained
for ‘RS poly’ describes reasonably well only the upper branch of its plot given in Fig.
6.53.
With the present algorithm for determination of RS, n and I0, the processing of I-
V curves recorded at high irradiances on a single day is very time-consuming (takes
several hours). The new method developed in the previous Section is expected to speed
up significantly the parameter-estimation procedure. This will facilitate the analysis of
data from multiple days, allowing one to study the variation of I0 over a much wider
range of operating temperatures.

5.7 CHAPTER SUMMARY


This Chapter reviews the assumed models and the methods used in the literature
for analysis of I-V curves of PV devices and estimation of their electrical and electronic
parameters. A new algorithm is presented for determination of RS from individual
illuminated I-V curves. The ideality factor and the reverse saturation current are then
extracted in the classic way. The method is applied to in-situ measured data of eight c-Si
modules representing four different cell technologies and one thin-film CIGS module.
The obtained results agree well with the estimates of RS from the double-light method,
which indicates that the new algorithm yields physically meaningful parameters.
Notable variation of RS with current (as well as with illumination level) is observed
only for the thin-film module.
227

The ideality factor estimates suggest a general variability of this parameter in c-


Si PV modules. For the majority of the tested devices, only a slight increase with
illumination level is seen, which may be due to an error of the method. Ideally, the
estimates should have been presented together with error bars indicating the total
uncertainty. The detailed quantification of the latter is complicated by the specific
method used to linearize the semi-logarithmic plots when determining RS. Therefore,
the evaluation of the total uncertainty is left for future work. Stronger variation of n is
obtained for the Chinese-made mc-Si module denoted ‘RS poly’. However, the results
need verification by measuring a suns-VOC or an ISC-VOC curve recorded at a fixed
operating temperature, which is not possible to obtain with the present outdoor test
setup. If the variation is independently confirmed, it may require a revision of PV
performance models assuming a fixed ideality factor such as the Sandia array
performance model [Kin04]. The latter was recently shown to be less sensitive to errors
in the ideality factor as compared to other parameters [Pra10]. Despite this, a new
analysis of the model accuracy may be required, one aiming to investigate the possible
implications of a variable n, particularly when modeling the performance of thin-film
devices, in which such variation seems to be the typical case. Extension may also be
needed of the IEC 60904-5 standard for ECT calculation from VOC in order to cover
also PV devices with notable irradiance dependency of the ideality factor. Such an
update is especially needed in the case of CIS/CIGS devices, for which the ideality
factor variation reported in the literature is confirmed by the present work.
An improved definition of local ideality factor is suggested resulting in m-V
plots unaffected by the series resistance. For the first time, m-I plots are introduced and
used in I-V data analysis. For all the studied c-Si modules, the novel differential
techniques reveal rising of the ideality factor at open circuit by up to 9 %. This
deviation from the classic 1-exponential I-V curve model resembles the “bumps” in
simulated and experimental m-V plots of different types of c-Si cells reported in the
literature. Detailed study of this behavior would require extension of the I-V curve
measurements well beyond the VOC. The phenomenon questions the accuracy of ECT
calculated using the standard procedure, which assumes a constant n.
A second new method is developed for determination of RS and n from a single
I-V curve, which is based on m-I plot of the classic local ideality factor. The new
approach has the potential to reduce significantly the time needed for automated
parameter extraction from hundreds or even thousands of I-V curves. It may be used in
future work for analysis of parameters from outdoor data recorded over many days.
Analysis of the local ideality factor at the intermediate and low voltages can be used for
filtering out I-V curves affected by bypass-diode activation due to small current
mismatch between the cells in a PV module, thus leading to erroneous estimation of
parameters.
228

A simple empirical model of the temperature dependency of I0 is suggested,


which achieves satisfactory fitting of experimental data over a narrow range of 30°C.
The faster algorithm for determination of RS and n invented in the present work allows
also the quick evaluation of I0 from Eq. (5.10a), and should facilitate the study of its
temperature dependency over a much wider range of ECT values.
229

CHAPTER 7 POWER RATING AND MODELING


7.1 CHAPTER AIMS AND OBJECTIVES
The main goals of this Chapter are to characterize the power and energy
performance of the studied PV modules and to explain with device physics the
significant module-to-module differences for generic (screen-printed) c-Si PV seen in
[Hul11], as well as in the present work. The objectives of this Chapter are: a review of
the PV performance models found in the available literature; power rating of the
modules at STC; evaluation of the coefficients k1 through k6 of the power performance
model of the European Commission’s Joint Research Centre (JRC) for each of the
studied modules; a comparison between the averaged coefficients for the screen-printed
c-Si devices and the representative coefficient sets obtained in [Hul10, Hul11];
derivation of theoretical equations linking the coefficients k1 and k2 to I-V curve model
parameters of a PV module; a comparison between the predicted and empirical values
of k1 and k2 for the studied c-Si modules; and a comparison between the energy yields
of the modules measured in the year 2011 and those predicted by the JRC performance
model.

7.2 REVIEW OF PROMINENT PV PERFORMANCE MODELS


At present, PV modules are still rated and marketed in terms of their maximum
power at STC, whereas information for a device’s performance at different illumination
levels and operating temperatures is rarely provided by the manufacturer. However, two
modules with the same STC efficiency may have very different energy production at the
same site. In terms of annual efficiency, c-Si modules may outperform the thin-film
modules by a factor of almost two; but in terms of kWh/kWP, the thin-film modules can
have a significant advantage [Wil04b, Pho11d]. From an investor’s perspective, another
important parameter is the annual energy yield of a PV system based on a given module
type under a particular location’s climatic conditions. Its evaluation requires, among the
other considerations involved, a performance model of the DC power output of an
individual PV module (assuming a 100 % efficient MPPT in the system’s power
conversion devices). An example of such a model (used in the European Photovoltaic
Geographical Information System, PVGIS) is given in [Hul10, Hul11]. A standard,
widely accepted model is currently not available. Different research groups used various
models, which in addition evolved over time, e.g. [Hul08, Hul10]. Alternative models
are still being developed and different PV technologies require different approaches,
especially for devices with strong spectral sensitivity, such as a-Si [Bey11]. The
continuing theoretical development seems inevitable, given the multidecadal evolution
of the dominant c-Si PV class and the short innovation cycles of emerging thin-film
technologies [Hul10].
230

7.2.1 THE AMERICAN SANDIA MODEL


The Sandia Array Performance Model was developed at the American Sandia
National Laboratories [Kin96, Kin98b, Kin04, Pra10]. The irradiance and temperature
dependencies of the voltage VM and the current IM at the MPP are modeled separately.
This involves determination of ten parameters: the STC values of the two quantities
mentioned above, their temperature coefficients, n, NS, and four other coefficients
denoted C0 through C3. In the more recent version of the model, IM is assumed to be a
quadratic function of the effective irradiance and a linear function of module
temperature, whereas VM is modeled as a quadratic function of the irradiance logarithm
and a quadratic function of the temperature [Kin04, Pra10]. Consequently, their product
PMAX is a cubic function of module temperature and a complex function of irradiance.
The determination of all the required parameters is a multi-step procedure and is not
straight-forward. The model’s complexity can be explained with its broader
applicability (it covers the less predictable a-Si technology, as well as concentrating
PV).

7.2.2 EUROPEAN MODELS


Eight prominent European performance models as of 2007 were compared in
[Fri07]. The then version of the JRC model (used in the PVGIS until 2010 [Hul08,
Hul10]) was an adaptation of the Sandia model with simplified irradiance and
temperature dependencies of IM and VM. Determination of coefficients was also
simplified by directly fitting PMAX(G,T). Another model fitted IM and VM separately,
while using almost the same equations as the JRC, and in addition accounted for the
difference between the operating temperature of the cells and the one measured at the
module’s back. Half of the models separated the efficiency’s dependencies on irradiance
and temperature by defining either η(G,25°C) or PMAX(G,25°C), and the corresponding
temperature coefficient. The latter was assumed constant in three models and irradiance-
dependent in the fourth. The remaining four models mixed both dependencies in a
single power or efficiency surface. Besides the mathematical differences, another major
distinction between the eight models was the approach used to determine their
parameters. Some research groups did this by fitting, while the others used either
statistical approaches or curve averaging. Eventually, for the same c-Si module, the
groups obtained temperature coefficients of efficiency ranging between 0.40 %/°C and
0.52 %/°C, and also quite different efficiency-overirradiance curves at module
temperature of 25°C. (An example of such curves is given in Fig. 7.1) A later round
robin of the eight approaches led to the conclusion that module characterization at
different irradiances and temperatures was a major source of uncertainty of the energy
yield prediction [Dit10]. The characterization accuracy was determined by the
measurement uncertainty and the mathematical model used to describe both
dependencies. A standard procedure was defined in IEC 61853-1 for indoor
231

determination of the performance surface PMAX(G,T) [Roy08, Dit10, Lad10, Kok11,


Sau12]. A matrix of 23 combinations of seven irradiance values (ranging between 100
and 1100 W/m2) and four temperature values (between 15 and 75°C), was specified.
This approach usually involves translation of I-V curves to the desired conditions,
which introduces additional errors. On the other hand, the fitting of experimentally
measured points from the performance surface offers flexibility, as well as potential for
very accurate characterization by including much more data.
It can be argued that the three mathematical models assuming an irradiance-
independent temperature coefficient of either PMAX or η are inferior to the other five,
because otherwise the extensive characterization specified in IEC 61853-1 would not be
necessary. However, it is possible that, for the dominating c-Si PV technology based on
screen-printed PESC cells, the temperature dependency is linear and practically constant
with irradiance. This may not be the case for c-Si modules made with HIT, IBC, PERC
and other cell types. Therefore, the more general recent version of the more
comprehensive JRC model (presented in [Hul10, Hul11] is tested with outdoor data
from the c-Si modules studied in the present work and the two characterized in [Yor10b,
Yor12a]. The current JRC model is described in detail further in this Chapter.

7.2.3 MODELS USED IN SOFTWARE FOR ENERGY YIELD PREDICTION


The commercial software PVsyst was based on a slightly modified 1-exponential
I-V curve model assuming irradiance dependency of the shunt current [Mer10, Sau12].
The mathematical model used in another commercial software, PV*SOL, has not been
made publically available. The default relative efficiency curves of the two programs
disagreed significantly with experimental data from “standard” mc-Si module types and
led to energy underestimation of up to 6 % [Sau12]. The largest discrepancies took
place for low irradiance levels (ΔηREL = – 11.6 % at 200 W/m2 for PV*SOL). Indeed,
one can expect a larger disagreement between the different efficiency models as the
operating conditions move further away from those defined at STC. Similar observation
was made in another study, which compared the aforementioned two programs and
PVSim (developed by SunPower Corporation and based on the Sandia model) [Don10,
Sun12]. Changing the efficiency model from the default one was straight-forward in
PV*SOL, but required numerical methods in the case of PVsyst [Sau12]. In addition,
the ideality factor n of the PV module was not included among the modifiable
parameters of the latter software. Consequently, an optimal efficiency model could be
obtained only by supplying RS and three parameters of the complex (and unnecessary in
the case of c-Si PV) equation describing the irradiance dependency of RP. This
approach seems to be non-physical due to keeping n fixed at an arbitrary constant value
and compensating its effect on the I-V curve’s shape by adjusting RS. In addition, it is
not clear how PVsyst addresses technologies with varying n or with current- and
irradiance-dependent RS.
232

The Sandia model was incorporated in the System Advisor Model (SAM)
software developed by the American National Renewable Energy Laboratory (NREL),
which uses the Sandia PV module database [NRE12d]. The latter contains the values of
all parameters required by the model for prediction of a PV array’s output, including
those describing the angular and spectral effects. The latest version dated November,
2011 lists over 500 PV modules of different technologies [Han12]. In the context of
exploding diversity of commercially-available PV devices discussed in Chapter 3, this
individual approach to each model offered by each manufacturer seems very resource-
demanding, and therefore unpractical. Hence, a more general approach is desirable in
which performance, angular and spectral effects are modeled equally for each subgroup
of similar PV devices.
The current version of the PVGIS (employing the JRC model) fulfills this only
partially by differentiating between the c-Si, CdTe and CIS/CIGS technologies
[PVGIS]. A validation study of the JRC model for mainstream c-Si modules concluded
that mono- and mc-Si devices could be treated as one subgroup, which did not affect
significantly the prediction accuracy [Hul11]. The early version of the PVGIS
considered only mono- and mc-Si modules (which still greatly dominate the market)
and used 10 years of meteorological data from ground-based instrumentation [Sur07,
Hul08]. Over the time, support was also added for thin-film PV, but initially a model of
the efficiency dependencies on irradiance and device temperature was only
implemented for crystalline silicon (c-Si) [Sur08a]. The tool has been under continuous
improvement and sophistication. More meteorological stations have been represented,
and a solar irradiation database of much better spatial resolution (based on satellite
images and covering a different period of time) has supplemented the ground data
[Hul10]. In addition, the long-term averaging of irradiance and temperature (found to
result in biased PV energy estimates) was replaced by probability density functions of
the climatic data. Furthermore, a more representative performance model based on data
from 16 generic c-Si module types (as of 2010; 18 as of 2011) replaced the earlier
version based on parameters derived from a single PV device [Hul08, Hul10, Hul11].
In the more recent versions of the PVGIS tool, support was added also for thin-
film CdTe, CuInSe2 (CIS) and Cu(In,-Ga)(Se,S)2 (CIGS) modules. Averaged model
coefficients were determined from outdoor data of just three CIS/CIGS modules
representing two manufacturers, and from three CdTe modules representing a single
producer. It was pointed out that the variety of represented modules should be expanded
and that subgroups based on cell design should be defined within each major PV class
[Hul10]. The possible differences in performance under varying irradiance and device
temperature in the case of “generic” c-Si modules (presumably with screen-printed
cells) were investigated in a more recent study [Hul11]. Individual model coefficients
were determined from indoor-measured data of 18 module types. It was found that the
inter-module variation at low irradiances far exceeded the predicted geographical
variation of c-Si PV performance in Central Europe, but was otherwise lower than the
233

typical uncertainty of a site’s annual solar irradiation. It was subsequently concluded


that using the model with an averaged coefficient set added only a small uncertainty to
the energy yield estimates.

Figure 7.1: Relative efficiencies of the studied c-Si PV modules for T = T0 = 25°C.

The PVGIS models the dependencies of a PV module’s maximum power PMAX


on the plane-of-module irradiance G and its operating temperature T in terms of the
dimensionless quantity relative efficiency [Hul10]:

  G ,T  PMAX  G ,T  G 0
REL  G ,T    (7.1)
0 G P0

where P0 and η0 are the rated power and efficiency at standard testing conditions (STC),
respectively; G0 = 1000 W/m2 is the standard irradiance; PMAX and η are respectively
the instantaneous maximum power and the corresponding efficiency at G and T. The
assumed functional form of the ηREL model is given in Eq. (7.2) where G’ = G/G0 and
T’ = T – T0, whereas T0 = 25°C is the standard operating temperature.

REL  G ', T '  1  k 1lnG ' k 2  lnG '  T ' k 3  k 4lnG ' k 5  lnG '   k 6T '2
2 2
(7.2)
 

The empirical coefficients k1 through k6 are obtained by fitting of experimental


data with Eq. (7.2). Fig. 7.1 shows plots of the relative efficiencies versus irradiance of
the ten c-Si modules studied in the present work for module temperatures fixed at 25°C
(i.e., for T’ = 0°C). The slope of each plot at G0 (i.e., for lnG’ = 0) equals k1/G0,
234

whereas the low-light behavior of ηREL is governed by k2. The coefficient k3 is easily
identified as the relative temperature coefficient of PMAX at STC. The three remaining
coefficients modify the shape of the plot at device temperatures different from T0. For
eight of the modules represented in Fig. 7.1, the plots look quite similar, despite the
different cell designs. On the other hand, the efficiency curves of those denoted ‘mc-Si
2’ and ‘mc-Si aged’ differ significantly at the intermediate and low irradiances. This
would result in different annual energy yields (and revenues) per kilowatt-peak of rated
(grid-connected) PV power, whereas the investment cost is proportional to the latter.
Thus, PV system owners would benefit from using modules performing much better at
the intermediate irradiances than at 1000 W/m2. A wider acceptance of this investment
strategy may make the PV manufacturers optimizing their products for energy yield
more competitive on the PV market. However, no relationship was revealed in [Hul10,
Hul11] between the model coefficients and I-V curve parameters such as the series
resistance RS and the ideality factor n (assuming a 1-exponential I-V curve model).
This Chapter contributes to the earlier research presented in [Hul10, Hul11] by
providing model coefficients for ten c-Si modules and one CIGS (CuInGaSe2) module.
The represented PV technologies include mainstream mono-Si and mc-Si screen-
printed, passivated emitter cells (PESC) with either 2 or 3 front busbars, heterojunction
a-Si/mono-Si cells with intrinsic thin a-Si layers (HIT), buried-contact (BC), as well as
interdigitated back-contact (IBC) mono-Si cells. The studied mc-Si modules represent
different types of SoG-Si feedstock: polysilicon purified in the classic Siemens
chemical route [Goe03]; compensated SoG-Si from an advanced metallurgical route
(Elkem Solar Silicon, ESS™ [Fri04]); and a Chinese-made mc-Si of unknown type
(perhaps an upgraded metallurgical grade Si, UMG-Si [Sin09]). In addition, theoretical
relationships relating k1 and k2 to I-V curve parameters are derived and verified.
Modeling of angular and spectral effects on PV performance is beyond the scope of the
present work. Instead, an approximate assessment of their magnitudes (as a percentage
of the annual energy yield) in Southern Norway is made throughout the thesis, as well
as in the present Chapter.

7.3 POWER RATING OF THE TESTED MODULES


The dominant factor in the energy prediction is the module’s power rating
PMAX,STC [Wil06]. According to leading European PV research groups, the best that one
can expect from outdoor data is matching the module’s parameters to standard
irradiance and temperature (SIT) [Moh10]. In this approach, normal AOI and matching
to a standard solar spectrum is not required. AOI<20° is usually considered matching
STC conditions (AOI=0°) [Wil04b]. However, in the present case of outdoor module
power rating, the AOI is in fact unimportant. This is because the used methods (to be
described shortly) apply linear fitting to values typically within ±15 % of module’s STC
power. As already discussed in Section 4.7, for modules with a front glass, the
235

reflectance losses for AOI within about 50° are practically the same as in the case of
normal incidence of light. In-plane irradiances close to 1 kW/m2 cannot be obtained at
higher AOIs, because the cosine loss becomes too high (50 % at 60°, higher for flatter
angles) and also the reflectance losses are increased. Even with the strongest
overirradiances observed in Southern Norway (of about 1.5 kW/m2), the in-plane
irradiance at an AOI of 60° would only reach about 0.75 kW/m2. However, the
strongest overirradiance events are not likely at such high AOI (which also mean high
solar azimuthal angles, and therefore lower solar elevations leading to lower clear-sky
irradiances). For these reasons, it was concluded that AOI filtering of outdoor data was
not necessary. Nevertheless, power rating was also implemented with I-V data recorded
at AOI≤50° which, as expected, did not lead to much different results (to within ±0.15
%).

Figure 7.2: “Southern” method for derivation of PMAX at SIT conditions. The self-
referenced irradiance GS is limited within 1000±50 W/m2.

Depending on the location of the outdoor test setup, different power rating
procedures may be applicable [Moh10]. In general, all the parameter estimation
procedures can be divided into two geographically-defined groups: “southern” and
236

“northern”. The first approach plots the parameter of interest versus module temperature
at a constant irradiance of 1000 W/m2 (within ±0.5 %). A by-product of this method is
the temperature coefficient of that parameter. In the case of power rating, the plotted
quantity is the module’s PMAX scaled by the factor G0/G. One can expect that this
correction should reduce the spread in the data subjected to linear fitting, thus
improving the overall accuracy of the method. This expectation is confirmed by the
present work, since the obtained “southern” plots contain minimal spread (see Fig. 7.2).
In the present work, ECT (derived as described in Section 5.4) is used as the
module temperature. Self-referenced irradiance (based on module’s ISC) is preferred to
data from the available irradiance sensors. The self-referencing approach eliminates the
need for matching the standard solar spectrum, provided that calibration has been
performed [Moh10]. As described in Section 4.7, a global self-irradiance based on data
from nine modules is defined and used in the present work. The simple, intuitive
requirement for equality of the individual self-irradiances eventually has led to the
development of a method for their simultaneous calibration without the need of a
reference device.
The “northern” method for rating of power and other parameters is applicable in
locations (typically at high latitudes) where SIT conditions occur naturally [Moh10].
The parameter (usually PMAX, ISC, VOC, IM or VM) is plotted over irradiance at a
constant (within ±1°C) module temperature of 25°C (see Fig. 7.3). In some locations,
both the “northern” and the “southern” methods are applicable. This is also the case of
the present test site, which is located on the southern coast of Scandinavia. The values
derived with the two methods for the nine modules studied in the present work are very
similar (within ±1.2 % or less), in line with the findings in [Moh10]. It can be argued
that none of the methods is preferable to the other in locations where both of them are
applicable. Correction for the loss due to the added series resistance ΔRS of extra
cabling and connectors (determined in Section 6.5) is applied prior to each method’s
2
application by adding I M RS to the value of PMAX extracted from each I-V curve.
Results for PMAX,SIT and its temperature coefficient obtained with the “southern”
method are listed in Table VII.1, together with the label values of PMAX,STC provided by
each module’s manufacturer.
The standard “northern” plots contain fewer data points and have larger spreads
as compared to the “southern” ones, as seen in Figs. 7.3 and 7.2. One possible reason is
the lack of correction for the temperature dependency of PMAX. However, its
temperature coefficient γ (not to be confused with the quantity used in Eq. (6.12a),
which is available from the “southern” method, provides the means for such a
correction. The standard plots given in Fig. 7.3 have been redone by scaling PMAX with
the factor 1/[1+ γ(ECT–25°C)]. The result is seen in Fig. 7.4, where the spread in the
plots is notably reduced. Nevertheless, the newly derived values of the modules’ powers
237

at SIT conditions did not differ (to within 0.2 %) from those derived with the original
method. Therefore, the improvement of the “northern” plots achieved in the present
work has little practical significance. Moreover, it requires prior knowledge of γ, which
is usually not available.

Figure 7.3: “Northern” method for derivation of PMAX at SIT conditions. ECTs are
limited within 25±1°C.

Figure 7.4: The “northern” method improved by the present author.


238

The ratings obtained for the two aged modules are lower than the label values,
which can be expected. The performance reduction over 14 years for the module made
with BC cells is 6.6 %, or about 0.5 %/year. This equals the median degradation rate for
c-Si PV reported in the literature, but is below the average value of 0.7 %/year [Jor10].
This excludes the initial light-induced degradation (LID) typical for the n+p c-Si
technology. The mean initial performance degradation of 66 c-Si modules after
exposure to 60 kWh solar irradiation amounted to 2.6 % [Dun03]. However, one can
expect that the manufacturer provides the stabilized PMAX,STC in the label. For the aged
mc-Si module, the degradation over 5 years of exposure to sunlight (4.5 of which in the
field) is 3.3 %, or 0.67 %/year, which is about the average rate reported in [Jor10].
However, there is a significant uncertainty in the order of 2-3 % associated with these
rates, which comes from the irradiance measurement performed by the manufacturer
during the factory rating. The self-referenced irradiance uncertainty is much lower in
the present work thanks to the applied correction factors described in Section 4.7. The
major mode of performance degradation for both aged modules is the reduction of
ISC,STC discussed in Section 4.7 in the context of self-referenced irradiance calculation.
On the other hand, a VOC,STC degradation (estimated in Section 5.4) is seen only in the
case of the BC module denoted ‘BP’, whereas this parameter seems to have increased
with time for the aged mc-Si module denoted ‘GPV’.

TABLE VII.1: PMAX rated at SIT conditions and its temperature coefficient.
Module PMAX,SIT [W] Label PMAX,STC [W] γ [%/K]
Aged BC mono-Si (‘BP’) 79.4 85 –0.474
Aged mc-Si (‘GPV’) 49.3 51 –0.503
Back-contact mono-Si (‘Sunc.’) 101.8 95 –0.384
RSS mono-Si (‘RS mono’) 292.3 300 –0.464
RSS mc-Si (‘RS poly’) 274.7 280 –0.481
Sanyo HIT (‘HIT’) 248.5 240 –0.333
SoG-Si 1 (‘ESS 1’) 226.5 216.30 –0.471
SoG-Si 2 (‘ESS 2’) 229.4 217.47 –0.431
AvanCIS (‘CIS’) 110.5 110 –0.505

For the new modules denoted ‘RS mono’ and ‘RS poly’, the obtained ratings are
lower than the label ones, which is partially attributed to irradiance uncertainty and
corresponds well to the correction factors applied to ISC,STC in Section 4.7. However, the
“southern” plot of ‘RS poly’ contains significant spread, which is another major source
of uncertainty (see Fig. 7.5). This may have been caused by albedo effects resulting in
imperfect I-V curves like the one shown in Fig. 6.27. For the HIT module and the thin-
film CGIS module, the obtained ratings are practically identical with the label ones.
239

However, much higher values are obtained for the module with back-contact cells (by
about 7 %), as well as for the modules denoted ‘ESS 1’ and ‘ESS 2’ (by about 5 %). For
the former device, the discrepancy can be explained only partially with the needed
correction of ISC,STC of +2.4 %, as well as that of VOC,STC of +3.2 % (the latter
determined in Section 5.4). The large performance ‘bounty’ can be explained with the
fact that this module was optimized for 12-Volt applications (lead-acid battery
charging), and not for grid-connected PV systems. It becomes obvious from its data
sheet that there is no binning of modules with different power ratings as in the case of
e.g. the AvanCIS PowerMax™ CIGS modules. Instead, a single model name “SPR-95”
is used, thus guaranteeing 95 W at STC. On the other hand, the relative temperature
coefficient of 0.38 % stated therein is in excellent agreement with the value obtained in
the present work. For ‘ESS 1’ and ‘ESS 2’, the corrections of ISC,STC and VOC,STC total
0.4 % and 2.3 %, respectively, which cannot explain the obtained excess performance
(especially for the former module). This issue can be investigated in detail in future
work, which may include analysis also of IM,STC and VM,STC, as well as of entire I-V
curves recorded at conditions very close to SIT.

Figure 7.5: Performance rating of the modules denoted ‘RS mono’ and ‘RS poly’.

Another possible approach to the power rating of PV modules is the translation


to STC of I-V curves recorded at different irradiance levels and device temperatures.
This is suitable in cases when only a limited number of I-V curves are available due to
e.g. time constraints. Such techniques are associated with an inherent error of PMAX,STC
due to the translation itself [Coo97]. More recently, Tsuno et al. [Tsu09] proposed a
translation procedure based on linear interpolation or extrapolation with greatly relaxed
240

requirements for the starting I-V curves. In particular, the four reference I-V curves can
be measured at completely different irradiances and module temperatures. The authors
reported an average accuracy of –0.45 % (i.e. there was an overall underestimation of
PMAX,STC) with a standard deviation of 0.68 %. Figs. 7.6 and 7.7 show an application of
this approach to experimental data from two c-Si modules tested in the present work.
Using reference I-V curves measured outdoors has offered the possibility to reduce their
number to just two, thus applying a single interpolation or extrapolation instead of three
as in the general case. This is achieved by selecting a couple of reference curves
measured either at TMOD=25°C or ECT=25°C, but at different irradiance levels.
However, TMOD differs significantly from ECT at the higher irradiances in the present
work (shown earlier in Chapter 5), and therefore ECT is the natural choice.

Figure 7.6: I-V curve extrapolation to SIT conditions after [Tsu09] for the module
denoted ‘ESS 1’. The starting I-V curves are recorded at TMOD = 25°C.

An advantage of this I-V curve translation method is its capability to


occasionally produce curves reaching short circuit, even when both reference curves are
incomplete due to the non-zero on-state resistance of the electronic load (see Fig. 7.6).
On the other hand, the use of illuminated reference curves results in a translated curve
which does not reach open circuit. This is due to the method’s requirement for matching
the data points of the reference curves two by two, so that I2=I1+(ISC2–ISC1) is always
fulfilled. Consequently, the translated curve is limited to currents I3=I1+a(I2–I1) ≥
a(ISC2–ISC1), where a positive extrapolation factor a > 1 is used and ISC2–ISC1 > 0.
Consequently, the translated curve lacks any data points between open circuit and the
specified minimum current level. In the original paper, two dark I-V curves were used
241

as references (i.e. I1 ≤ 0) [Tsu09]. This had supposedly allowed the authors to obtain
complete translated curves extending to open circuit. Dark I-V curves are not available
in the present work. However, obtaining only the part around the MPP of the SIT curve
is sufficient for the purpose of power rating. On the other hand, the estimation of
VOC,STC with this method using illuminated reference curves would require complicated
curvilinear extrapolation resulting in some additional error.

Figure 7.7: I-V curve extrapolation to SIT conditions after [Tsu09] for the module
denoted ‘ESS 2’. The starting I-V curves are recorded at ECT = 25°C.

When using discrete I-V data, a point with I2=I1+(ISC2–ISC1) does not always
exist in the 2nd reference curve. Eventually, interpolation must be applied along the
curve in order to create the needed point artificially, thus introducing an additional
error. In this context, the use of 4000+ data points per I-V curve in the present work is
highly beneficial, since thus the interpolation is done over a very narrow current range.
Consequently, linear rather than curvilinear interpolation is sufficient. Another
drawback of the method due to the discrete reference data is the poorer accuracy of the
translation in the high-current region. This is attributed to the scarcity of data points and
the possible activation of bypass diodes in the ‘flat’ region of the I-V curve. An extreme
example is given in Fig. 7.7, where the translated curve does not reach short circuit, has
large gaps, and contains many erroneous data points in the MPP region. However, the
artifacts can be excluded manually from the plot, and then PMAX,STC is easily determined
242

after applying a correction +IM,STCΔRS to the I-V curve in order to compensate for the
voltage drop in the extra cabling and connectors. At this stage of the present work, I-V
curve translation to STC has not been used as an alternative method for performance
rating. Future work may include comparison between the results in Table VI.1 and
values obtained with the method just described.

7.4 MODELING PERFORMANCE IN TERMS OF RELATIVE EFFICIENCY


In the following, several important questions about the performance model used
in the PVGIS online tool are answered. Why modules of the same type can have a
vastly different relative efficiency at intermediate and low irradiances? In Fig. 7.1, the
Chinese-made mc-Si module (denoted ‘RS poly’) and the aged European-made mc-Si
module (denoted ‘GPV’) have notably different relative efficiency plots. Similarly,
some of the modules analyzed in [Hul11] performed quite differently from the bulk.
What is the physical explanation of this behavior, and what are the effects of the various
I-V curve parameters such as RS and n on a module’s performance under different
operating conditions? What is the relationship between these parameters and the
coefficients k1 through k6 used in the relative efficiency model given in Eq. (7.2)?
These questions are addressed analytically starting from the shunt-free 1-exponential I-
V curve model, from which an expression for ηREL is derived. In addition, the yearly
averaged ηREL of the studied c-Si modules and their expected energy yields are
predicted using in-plane irradiance and module temperatures measured in 2011. These
values are then compared to the actually measured yields obtained with the MPPT
algorithm of the data acquisition system. The key results were presented in [Yor12f].

7.4.1 ESTIMATION OF k1 THROUGH k6 FOR THE STUDIED MODULES


Outdoor data of ten c-Si modules and one CIGS module from the months April
through September, 2011 are used to estimate the coefficients k1 through k6 of the
relative efficiency model introduced in [Hul10]. Eight of the c-Si devices belong to the
experimental setup in Grimstad described in Chapter 3, whereas the remaining two are
those studied in [Yor10b, Yor12a]. The relative efficiency of each module is calculated
using Eq. (7.1) from filtered experimental data and plotted over lnG’ and T’. Then, the
obtained surface is fitted in Matlab™ with Eq. (7.2). For each module, its ECT derived
from VOC and its self-irradiance GS derived from ISC are used to calculate the two
independent variables of the model. Any obvious outliers are manually excluded from
the data set prior to the fitting. The data recorded before 1 April, 2011 are disregarded in
order to avoid partially shadowed I-V curves (due to snow, ice or the nearby telecom
tower in the case of the modules tested in Grimstad). Further filtering of the data
includes the requirement for a good agreement between the self-irradiances of the nine
modules (to within ±5 % from their median value). In this way, erroneous values of ISC
due to e.g., incomplete I-V curves, are generally avoided. In addition, the requirement
243

Module k1 [ - ] k2 [ - ] k3 [K-1] k4 [K-1] k5 [K-1] k6×106 [K-2]

*A slightly different notation of the coefficients was used in that paper.


Aged BC mono-Si (‘BP’) –0.0090 –0.0309 –0.00478 –0.00034 –0.00002
TABLE VII.2: Fitted coefficients of the relative efficiency model.

+1.0
Aged mc-Si (‘GPV’) +0.0290 –0.0308 –0.00506 –0.00016 +0.00033 +2.2
Back-contact mono-Si (‘Sunc.’) +0.0290 –0.0040 –0.00389 +0.00018 +0.000025 +1.6
RSS mono-Si (‘RS mono’) +0.0052 –0.0225 –0.00464 –0.00045 –0.00024 –0.9
RSS mc-Si (‘RS poly’) –0.0243 –0.0302 –0.00490 –0.00036 –0.00038 +2.7
HIT +0.0070 –0.0154 –0.00308 +0.00025 –0.00002 –7.0
SoG-Si ‘ESS 1’ +0.0064 –0.0178 –0.00467 –0.00002 +0.00007 –0.3
SoG-Si ‘ESS 2’ +0.0031 –0.0185 –0.00437 –0.00024 –0.00018 +1.1
CIGS (‘CIS’) –0.0042 –0.0344 –0.00496 +0.00045 –0.00004 –1.9
CIS [Hul10] –0.005521 –0.038492 –0.003701 –0.000899 –0.001248 +1
Screen-printed c-Si, averaged +0.0072 –0.0210 –0.00456 –0.000286 –0.000109 +1
Screen-printed c-Si [Hul10] –0.017162 –0.040289 –0.004681 +0.000148 +0.000169 +5
Screen-printed c-Si [Hul11]* –0.01724 –0.04047 –0.0047 +0.000149 +0.000147 +5
244

AOI<50° is imposed on the data. For such AOI, the reflectance losses in modules with
a front glass are practically equal to those at normal incidence of light [Mar01]. Finally,
only irradiances above 100 W/m2 are included in the modeling in order to avoid the ill-
defined ECTs of some modules below this level.
The robust bisquare fitting algorithm is preferred in order to suppress any
remaining outliers in the data. In fact, replacing the unity in the fitted equation with a
seventh coefficient (degree of freedom) allows one to test the accuracy of PMAX,STC
estimation. For all the modules, fitting with the so modified Eq. (7.2) results in
practically the same set of coefficients, while unity is obtained for the seventh (as
expected). The obtained values for the modules installed in Grimstad are listed in Table
VII.2 together with an averaged coefficient set for the seven c-Si modules with screen-
printed cells from both test locations, and averaged sets for the same technology given
in [Hul10, Hul11]. In addition, Table VII.2 lists the coefficient set obtained in [Hul10]
for CIS PV. The temperature dependency of ISC is not accounted for when calculating
each module’s self-irradiance. It is shown in Appendix G that the worst-case error in
ηREL resulting from this would be about 0.5 % relative, which is negligibly small. The
results obtained for the two mc-Si modules installed in Kristiansand are skipped due to
space limitation. In addition, there is some evidence (to be discussed shortly) that their
coefficients (at least k1) have not been determined very accurately.
Good agreement between the averaged sets obtained in [Hul10, Hul11] for
screen-printed c-Si and the one obtained in the present work is seen only for k3 (the
relative temperature coefficient of PMAX at STC). The obtained values are practically
identical with the ones listed in Table VII.1. The big differences in magnitude and sign
for the remaining coefficients may be partially explained with their large variation
among such ”generic” c-Si modules. The standard deviations are as follows: σk1=0.0165
(about 2.3 times the mean value); σk2=0.0072 (34 %); σk3=0.00037 K–1 (8 %);
σk4=0.00019 K–1 (67 %); σk5=0.00024 K–1 (2.2 times the mean); σk6=0.0000016 K–2
(1.3 times the mean). In general, the largest differences between the ηREL curves come
from k1 which determines the slope at G = 1000 W/m2 (see Fig. 7.1). This coefficient
can take both positive and negative values, whereas k2 seems to be always negative due
to device physics. These observations are analyzed analytically further in this Chapter.
A somewhat better agreement between the present work and [Hul10] is obtained for the
CIS/CIGS technology. However, the results obtained by the present author come from a
single module, and therefore they are not very representative for this PV type.

7.4.2 EXPRESSING k1 AND k2 THROUGH I-V CURVE PARAMETERS


The large variation in most of the coefficients (except for k3) among screen-
printed c-Si modules is strange. The resulting irradiance dependencies of ηREL can be
245

very different as well (see Fig. 7.1), which was observed also for the modules tested in
[Hul11]. For example, the Chinese-made module denoted ‘mc-Si 2’ in Figs. 7.1 and 7.8
(‘RS poly’ throughout the thesis) performs somewhat better than the rest at low
irradiances, but much worse at high irradiances, at which its fill factor degrades
significantly due to an increasing n (see Figs. 6.1 and 6.24). Significant differences
were observed also among the c-Si modules tested outdoors in 2011 by Photon Lab,
which used relative efficiency to analyze each device’s weak-light performance
[Pho12a, Pho12b, Pho12c]. That study argued that, in general, a module should perform
better under lower irradiances due to smaller resistive losses. However, no clear link
has been found so far between the performance model coefficients k1 through k6 and the
module’s I-V curve parameters such as the ideality factor n and the series resistance RS.
Identification of such relationships could be beneficial for PV manufacturers and may
lead to design of modules producing superior energy yields. The present work suggests
analytical expressions derived from the standard 1-exponential I-V curve model with
constant n and RS, which link the coefficients k1 and k2 to measurable physical
parameters. In the MPP region, the I-V curve is assumed to follow exactly the
exponential dependency as the one represented in Fig. 6.32, whereas shunting and cell
mismatch effects are assumed negligible according to [Yor11a]. In addition, the ratio
μ=IM/ISC is assumed constant for most c-Si modules [Ser10a, Yor10c]. The detailed
derivation is given in Appendix H. For the first performance model coefficient, it gives:

nNS v 0   R S I SC ,STC nNS v 0  R S I M ,STC


k1   (7.3)
VOC ,STC  nNS v 0 ln 1      RS ISC ,STC VM ,STC

where NS is the number of serially-connected cells; v0 is the thermal voltage at STC;


ISC,STC and VOC,STC are the short-circuit current and the open-circuit voltage at STC,
respectively; IM,STC and VM,STC are the MPP current and voltage at STC, respectively.
In the hypothetical case of an ideal PV module with RS=0, Eq. (7.3) simplifies to:

1  I 
k1   ln1 1    SC ,STC  (7.4)
VOC ,STC
 ln 1     I 0 ,STC 
nNS v 0

where I0,STC is the reverse saturation current at STC (denoted I00 in Appendix H). The
last expression in Eq. (7.4) is independent on n and always positive. This means that, at
standard device temperature of 25°C, such an idealized module would perform worse at
lower irradiances compared to STC just like most of the real modules represented in
Fig. 7.1. According to Eq. (7.4), one way of improving the performance at intermediate
irradiances is by minimizing the rate of recombination in the cells, while increasing the
246

photogenerated current. Experience shows that the second coefficient, k2, is always
negative and results in even poorer performance at intermediate and low illumination
levels. Therefore, according to Eq. (7.3), only a large enough RS can make a module
perform better at intermediate irradiances than at STC (at constant T = 25°C), by
making k1 negative and of large enough magnitude. This can be achieved by optimizing
the module design in such a way that RS ≫ nNSv0/IM,STC. This conclusion is very
important, because normally, module makers would try to minimize the resistive losses
in their products in order to achieve better performance at STC (and thus a larger profit).
On the other hand, n cannot be made arbitrary low. Using large-area cells would
increase IM,STC, but this is also expected to reduce RS by about the same factor.
Eventually, a change of NS can be expected to result in a proportional change of RS. It
seems after all that k1 is determined mainly by the cell technology used, and cannot be
varied over too wide a range of optional values. However, the results obtained in
[Hul11] and in the present work for the mainstream screen-printed c-Si technology
indicate that such fine-tuning is possible and is seen in commercial PV devices. For
example, the module denoted ‘mc-Si 2’ in Fig. 7.1 (‘RS poly’ throughout the thesis) has
a negative k1. This can be attributed to its relatively high RS (according to [Yor11a] and
the results presented in Chapter 6) due to using only two busbars in cells of size
15.6×15.6 cm2. The best-performing new module in Photon Lab’s comparison,
REC230AE, was similarly made with large-area cells having only two busbars [Pho12a,
Pho12c, REC12]. On the other hand, the IBC module (denoted ‘Sunc.’ throughout the
thesis) has one of the largest positive values of k1 among the modules represented in
Fig. 7.1, which is well explained with the relatively low RS typical for this cell
technology [Ker06, Neu07, Yor11a].
A comparison between fitted and theoretical values of k1 for the studied modules
is given in Table VII.3. For most of these, the two estimates agree reasonably well
except for ‘mc-Si 1’ and ‘mc-Si 3 ESS’, which are part of the experimental setup in
Kristiansand. The latter issue needs further investigation and must be addressed in
future work. Note that errors in n and RS can increase the relative uncertainty of k1
significantly, because the corresponding terms in the numerator in Eq. (7.3) have
opposite signs. In addition, from the analysis of RS-corrected m-I plots presented in
Section 6.7 it becomes clear that the region with an almost constant ideality factor quite
often excludes the MPP; moreover, the I-V curves of some PV module types cannot be
accurately described by the 1-exponential model due to an irradiance-dependent ideality
factor. Furthermore, there are PV technologies for which RS depends on the current
density and may also depend on the illumination level, as discussed earlier in Chapter 6.
Even in such cases, Eq. (7.3) can be used as a quantitative guideline if maximizing the
annual energy yield is among the priorities of a module’s design process.
247

Figure 7.8: Relative efficiencies averaged over the modules’ temperatures in 2011.

TABLE VII.3: Fitted vs. predicted values of k1 for ten c-Si PV modules.
Module Fitted k1 [ - ] Theoretical k1 [ - ]
‘RS mono’ 0.0052 –0.0029
IBC (‘Sunc.’) 0.0290 0.0248
HIT 0.0070 0.0028
BC aged (‘BP’) –0.0090 –0.0066
mc-Si 1 Polysilicon [Yor12a] 0.0178 0.0050
mc-Si 2 (‘RS poly’) –0.0243 –0.0227
mc-Si 3 ESS™ [Yor12a] 0.0132 0.0027
mc-Si 4 (‘ESS 1’) 0.0064 0.0059
mc-Si 5 (‘ESS 2’) 0.0031 0.0062
mc-Si aged (‘GPV’) 0.0290 0.0223

In order to assess the magnitudes of the two major terms in Eq. (7.3), one may
define and evaluate the following non-physical quantity:

nNS v 0  R S I M ,STC
k 1+  (7.5)
VM ,STC

Note the big difference between Eqs. (7.5) and (7.3). Unlike k1, which can take both
positive and negative values, as well as zero (depending on the balance between the two
terms in the numerator), k 1+ is always positive. In Table VII.4, the absolute difference
|Δk1| between the theoretical and fitted value of k1 is compared to k 1+ for each of the
248

studied modules. For most of them, the discrepancy between theory and data fit is
within 5-6 % of the combined magnitude of the terms proportional to n and RS in Eq.
(7.3). Therefore, it is concluded that the theory predicts the fitted values of k1 with a
reasonable accuracy, which is not obvious from Table VII.3. A situation is possible in
which k1 = 0 is obtained by fitting for a given PV module type, while a non-zero value
results from Eq. (7.3). In such a case, the relative error of the theoretical prediction
would be infinite (with respect to the fitted value), but could be negligibly small
compared to k 1+ – the proper gauge.

Table VII.4: |Δk1| vs. k 1+ for the ten c-Si modules.


Module k 1+ [ - ] |Δk1| [ - ] |Δk1|/ k 1+ [%]
‘RS mono’ 0.1349 0.0081 6.0
IBC (‘Sunc.’) 0.0877 0.0042 4.8
HIT 0.1053 0.0042 4.0
BC aged (‘BP’) 0.1345 0.0024 1.8
mc-Si 1 Polysilicon [Yor12a] 0.1218 0.0128 10.5
mc-Si 2 (‘RS poly’) 0.1707 0.0016 0.9
mc-Si 3 ESS™ [Yor12a] 0.1178 0.0105 8.9
mc-Si 4 (‘ESS 1’) 0.1189 0.0005 0.4
mc-Si 5 (‘ESS 2’) 0.1185 0.0031 2.6
mc-Si aged (‘GPV’) 0.1225 0.0067 5.5

For |ln(G/G0)|≪1, the following expression for k2 is derived in Appendix G,


which explains its always negative sign:

0.5 R S ISC ,STC I


k2    M ,STC R S (7.6)
VOC ,STC  nNS v 0 ln 1      R S ISC ,STC 2VM ,STC

When |ln(G/G0)| becomes comparable to 1 (i.e., when G is much lower or much


higher than 1000 W/m2), the powers of ln(G/G0) higher than the second cannot be
neglected and the model given by Eq. (7.2) becomes mathematically inaccurate.
Therefore, the fitted k2 must compensate for the missing terms, and thus it may differ
significantly from the value predicted by Eq. (7.6), which is confirmed by the results
obtained in the present work. According to Eq. (H.11) in Appendix H, all additional
terms have negative signs, which result in even lower ηREL at intermediate and low
irradiances than the one resulting when Eq. (7.6) is used in the calculation. In addition,
the ratio μ=IM/ISC starts to decrease at low irradiances (see Fig. 7.9), whereas the I-V
curve starts to deviate from the one-exponential model (that is, the MPP moves into a
249

region of higher local ideality factors). Note that for some of the c-Si devices tested in
the present work, the (RS-corrected) local ideality factor at the MPP is well above the
value of n determined from the flat region already at high irradiances (see Figs. 6.39,
6.41 and 6.42). Therefore, using Eq. (7.2) to fit experimental data obtained at too wide a
range of irradiances may give somewhat inaccurate values for k1 and k2, as well as for
the remaining coefficients. Detailed investigation of such effects is outside the scope of
the thesis, and may be done in future. Nevertheless, Eq. (7.6) seems to predict rather
accurately the low-light performance of c-Si modules. For example, the relative
efficiency curve of the module with IBC cells (denoted ‘Sunc.’ thereafter) in Fig. 7.1
does not decline as sharply at low irradiances as those of the other devices. This
corresponds well to its very low RS which, according to Eq. (7.6), should result in a k2
of relatively low magnitude. Indeed, the value listed for that module in Table VII.2 has
a notably lower magnitude (by a factor of 4 to 8) compared to the estimates for the
remaining c-Si devices.

Figure 7.9: The ratio μ=IM/ISC for the aged BC mono-Si module denoted ‘BP’.

The general validity of the theory behind Eq. (7.3) has been confirmed also by
fitting with Eq. (7.2) of data which are not corrected for the resistance of the cables
interconnecting the module and its load. Thus, for the modules denoted ‘mc-Si 1’ and
‘mc-Si 3”, notably lower values of k1 (–0.0092 and –0.0149, respectively) have been
obtained. The reductions have been predicted reasonably well with Eq. (7.3). This is
important in the context of the PVGIS, because especially the large PV systems involve
a lot of additional cabling, and thus some extra series resistance. According to Eqs. (7.3)
and (7.6), both k1 and k2 (which, together with k4 and k5, determine the irradiance
250

dependency of ηREL) must be duly modified when the total RS of a PV system is


significantly higher than the combined intrinsic resistance of its modules. The system-
level losses (including the extra cabling) are accounted for separately in the PVGIS, and
can be changed by the user from the default value to a certain percentage of the
system’s annual energy yield. Nevertheless, the temperature coefficient of the additional
leads’ resistance can modify the temperature dependency of performance defined in
terms of ηREL of one module.
In the case of k2, the theoretical values are very different from the fitted ones
(see Table VII.5). Such discrepancies could be expected, as already discussed above.
The main reason is believed to be that Eq. (7.2) becomes inaccurate when |ln(G/G0)| is
comparable to, or higher than unity (i.e., at intermediate and low irradiances, or at
extreme overirradiances). The related consequences for k1 are much milder, as seen
from Table VII.4. For most of the modules, Eq. (7.6) predicts a much lower (i.e. worse)
value of k2 than the fitted one except for the aged mc-Si device. In Chapter 5, its
ideality factor was found (indirectly) to increase at low levels of illumination.
Therefore, the corresponding VM is lower than the value derived from the one-
exponential I-V curve model and results in a degraded ηREL. The latter leads to a rather
low fitted k2. Shunting in PV devices increases the local ideality factor at intermediate
voltages. Significant shunting of individual cells in a PV module is known to degrade its
performance at intermediate and especially at low irradiances [McM96]. Such is not
accounted for when deriving the theoretical expressions for k1 and k2, and is expected to
increase the magnitude of the latter coefficient when obtained from data fitting.

Table VII.5: Fitted vs. predicted values of k2.


Module Fitted k2 [ - ] Theoretical k2 [ - ]
‘RS mono’ –0.0225 –0.0340
IBC (‘Sunc.’) –0.0040 –0.0120
HIT –0.0154 –0.0256
BC aged (‘BP’) –0.0309 –0.0346
mc-Si 1 Polysilicon [Yor12a] –0.0141 –0.0264
mc-Si 2 (‘RS poly’) –0.0302 –0.0464
mc-Si 3 ESS™ [Yor12a] –0.0130 –0.0278
mc-Si 4 (‘ESS 1’) –0.0178 –0.0262
mc-Si 5 (‘ESS 2’) –0.0185 –0.0270
mc-Si aged (‘GPV’) –0.0308 –0.0246

The performance results presented in the thesis refute the expectation expressed
by Photon International researchers that (at a fixed operating temperature T = 25°C) a
251

module should generally perform better at lower irradiances than at 1000 W/m2 due to
the effect of RS [Pho12b]. For most of the c-Si PV types tested in the present work, the
relative efficiency declines at intermediate irradiances (see Fig. 7.1) due to a positive or
small negative k1 and a negative (as a general rule) k2. It seems that the 3-busbar cell
design (resulting in reduced resistive losses) is preferred by the industry for large-area
screen-printed c-Si cells of the mainstream PESC type. Most of these are characterized
with a relatively low n between about 1.1 and 1.2, very close to the unity limit for low-
injection SRH recombination that dominates non-concentrating c-Si PV. Eventually, the
fine balance in the numerator in Eq. (7.3) between the term containing n and the one
proportional to RS results in a positive k1 for the present state-of-the art technology. A
negative k1 is obtained only for the module with screen-printed cells having only two
busbars. Using more than two busbars in the front-side metallization is beneficial in the
context of product safety by providing more redundant current paths, thus preventing
series arcing failures due to open-circuiting which can cause fires [Rei12, Woh12,
Joh12a, Joh12b, Joh12c]. For the tested ‘non-generic’ c-Si PV types, k1 is positive for
the module with IBC cells (denoted ‘Sunc.’ throughout the thesis) and the one with HIT
cells. This can be expected from Eq. (7.3) for the former technology, which is
characterized with very low RS, whereas n is well above unity (see Figs. 6.14 and 6.42).
The HIT device has a similarly high value of n, which eventually dominates over RS in
Eq. (7.3). The second of the only two negative k1 obtained in the study of ten c-Si
devices belongs to the aged module made with buried-contact cells (vintage ca. 1997) –
a technology which is no longer in production to the knowledge of the present author.
On the other hand, in the averaged coefficient sets obtained in [Hul10, Hul11] for
‘generic’ c-Si modules, a negative k1 is listed which may (or may not) be explained by
testing of devices of predominantly older vintages at the represented research
laboratories. This can be discussed with the authors of the referenced papers in future.

7.4.3 FURTHER VERIFICATION OF THE MODEL


It can be argued whether such a complex performance model as the one given in
[Hul10, Hul11] is really needed for accurate performance modeling and energy yield
assessments. Are all the coefficients k4 through k6 really necessary, or some (if not all)
of them can be dropped? One can analyze their relative significance by re-writing Eq.
(7.2) in the form:

 k4 k k 
REL  1  k 1lnG ' k 2  lnG '  k 3T ' 1  lnG ' 5  lnG '  6 T '
2 2
(7.7)
 k3 k3 k3 

and then comparing to unity the sum of the last three terms in the square brackets
calculated for extreme (non-zero) values of T’ and lnG’. Doing this for the values listed
252

in Table VII.2 is left for future work. If the contribution of these terms is negligible
even for the worst-case scenarios, the last three model coefficients are not needed for
the particular PV device. Limiting the model to just the first three coefficients would
mean that the efficiency-overirradiance curves at various operating temperatures could
be obtained by simple translation along the vertical axis of those defined at T = 25°C
(shown in Fig. 7.1). Fig. 7.10 contains a proof that this is not always the case. Namely,
the curves for the higher temperature are flatter near 1000 W/m2 than those for 25°C.

Figure 7.10: Relative efficiency curves for different operating temperatures of the
two mc-Si modules studied in [Yor10b, Yor12a].

Another way to analyze the significance of k4 through k6 is by re-writing Eq.


(7.2) in the form:

  k 6   k   k 
REL  G ', T '  1  k 3T ' 1  T '    k 1 1  4 T '  lnG ' k 2 1  5 T '   lnG '  (7.8)
2

  k 3   k1   k2 

and then comparing (k4/k1)T’ and (k5/k2)T’ to unity for extreme non-zero magnitudes
of T’, as well as by comparing (k6/k3)T’ to unity for lnG’=0 (i.e. at standard irradiance).
253

Figure 7.11: Accuracy testing of the ηREL model for the module ‘RS mono’.

Figure 7.12: Accuracy testing of the ηREL model for the module ‘RS poly’.

Another aspect of the model’s accuracy verification is to test if experimental


efficiency-overirradiance curves at a fixed T’ are fitted well with a quadratic
polynomial with respect to lnG’ as predicted by Eq. (7.2). Two examples are given in
Figs. 7.11 and 7.12. For the modules denoted ‘RS mono’ and ‘RS poly’, ηREL data
limited to a fixed (within ±1°C) operating temperature are fitted with a quadratic
polynomial. For the former device, a very satisfactory fit is obtained, whereas for the
254

latter one the fit is not as good at irradiances close to 1000 W/m 2. In the case of ‘RS
poly’, one should bear in mind that the ECT calculation is based on a non-standard
equation (discussed in Section 5.4) using a significantly varying n, which is a probable
source of errors. Poorer fit quality is obtained also for the CIGS module, for which even
stronger ideality factor variability is obtained. Otherwise, efficiency-overirradiance
curves (for a fixed ECT) of the other tested c-Si devices are fitted reasonably well with
a quadratic polynomial. The corresponding plots are skipped for the sake of brevity of
the discussion. This finding suggests that the studied performance model is very
accurate for most c-Si PV types.

7.4.4 PREDICTED VS. MEASURED ENERGY YIELDS FOR YEAR 2011

TABLE VII.6: Yearly averages of the modules’ relative efficiencies.


Module ηREL,AVG [ - ]
‘RS mono’ 0.955
IBC (‘Sunc.’) 0.963
HIT 0.962
BC aged (‘BP’) 0.957
mc-Si 1 Polysilicon [Yor12a] 0.966
mc-Si 2 (‘RS poly’) 0.957
mc-Si 3 ESS™ [Yor12a] 0.968
mc-Si 4 (‘ESS 1’) 0.966
mc-Si 5 (‘ESS 2’) 0.963
mc-Si aged (‘GPV’) 0.935

In real outdoor conditions, a module’s efficiency at a given level of illumination


varies significantly due to the wide range of possible operating temperatures. Fig. 7.8
shows relative efficiencies of the tested modules averaged for different irradiances. The
modules denoted ‘mc-Si 1 and ‘mc-Si 3 ESS’ are part of a test setup with non-optimal
geometry [Yor11b]. Their plots are created using their individual performance models,
but assuming the same tilt angle as that of the eight modules installed in Grimstad and
operating temperatures equal to those of ‘mc-Si 4 ESS’ (denoted ‘ESS 1’ throughout the
thesis). Overirradiances are usually associated with rather low module temperatures,
which explain the increasing efficiencies above about 1250 W/m2. The contribution to
the annual irradiation of each irradiance level is given in Fig. 4.40 in Chapter 4. The
annual relative efficiency of each module (given in Table VII.6) is obtained by
averaging its efficiency values given in Fig. 7.8 weighted with the irradiation shares
(according to the thermopile pyranometer) given in Fig. 4.40. A theoretical estimate of
module’s energy yield (per kilowatt-peak of rated power) is then obtained by
multiplying its yearly-averaged ηREL with the total in-plane irradiation (1200 kWh/m2 in
2011) divided by the standard irradiance G0 (1.0 kW/m2). The predicted yields are in
255

very good agreement with the measured ones based on the 1-minute data set (corrected
for the added series resistance effects of the extra cabling and connectors), see Table
VII.7.

TABLE VII.7: Final yields of the modules in [kWh/kWP].


Module Theoretical Empirical Δ [%]
‘RS mono’ 1146 1134 –1.1
IBC (‘Sunc.’) 1156 1153 –0.3
HIT 1155 1147 –0.7
BC aged (‘BP’) 1149 1132 –1.5
mc-Si 1 Polysilicon [Yor12a] 1159 N/A* N/A
mc-Si 2 (‘RS poly’) 1149 1138 –1.0
mc-Si 3 ESS™ [Yor12a] 1162 N/A* N/A
mc-Si 4 (‘ESS 1’) 1160 1141 –1.6
mc-Si 5 (‘ESS 2’) 1156 1139 –1.5
mc-Si aged (‘GPV’) 1122 1114 –0.7
*The two modules have not been measured continuously in 2011. They are installed in a
different location at a non-optimum tilt angle. In addition, one of them suffers
systematic partial shadowing during the cold season [Yor11b].

The model’s overestimation of the yields for most of the specimens can be
partially explained by increased reflectance losses at flat angles of incidence, which are
not accounted for. These are within a few per cent for optimally inclined, south-facing
c-Si modules, and vary slightly with geographical location and with the type of anti-
reflection coating used in the cells [Mar01, Hul11]. The model coefficients have been
determined using self-referenced irradiance, whereas the predicted energy yields are
based on the annual solar irradiation measured with the thermopile pyranometer. The
latter is unaffected by the reflectance losses at flat AOI typical for c-Si modules. In
addition, frost and partial snow/ice cover in winter could not be completely eliminated
for all modules (especially for the aged BC module), despite the regular cleaning. On
the other hand, there are uncertainties associated with the quantities used in the model.
The typical uncertainty of irradiance measurements is in the order of 2-3 % or more.
Another source of uncertainty is the varying spectral composition of sunlight. A yield
increase of 0.7 % associated with spectral effects in c-Si modules is expected for a
Central European climate [Hul11]. A somewhat higher value is possible in Southern
Norway due to the lower solar elevations and the resulting redder spectra of direct
sunlight. Initial degradation is also not accounted for by the model. The eight c-Si
modules in Grimstad have been installed outdoors in December, 2010. Their initial
performance degradation is expected to have affected mainly the yields measured in
January, 2011 when only about 4 % of the annual solar irradiation was received.
Therefore, the effect of initial degradation on the final yields is considered negligible.
256

The individual model of the aged mc-Si module predicts a notably lower yield
compared to those of the other modules (by about 3 %), and also its measured yield is
lower. Therefore it can be expected that an averaged performance model would lead to
systematically overestimated yield predictions for some aged PV modules. The poor
performance of this particular module corresponds well to its reduced fill factor at lower
irradiances, see Fig. 6.1. It has shown a relatively poor weak-light performance as new
[Mid10]. This module was made in 2004 with cells of size 10×10 cm2 based on older
processing technology, which may suffer from photocurrent losses (leakages) at non-
insulated edges [Pho12a]. Such edges are non-Ohmic shunts with diode-like current-
voltage characteristics (I-V curves) with ideality factor of about 2 and reverse saturation
current proportional to the cell’s perimeter [McI01, Bre07a]. In smaller-size cells, the
perimeter-to-area ratio is larger, which increases the detrimental effect of edge leakage
on cell performance at low illumination levels. On the other hand, the performance
model of the aged BC mono-Si module does not predict much lower yield compared to
those of the new devices. Note that the relative efficiency definition in Eq. (7.1) uses the
estimated (i.e. degraded) STC power rating P0 instead of the label value.
In general, the tested performance model is found to predict the annual energy
yield very accurately for c-Si PV. The measured annual relative efficiencies of new
devices are among the highest in Europe. This is attributed to rather low operating
temperatures of the modules in Southern Norway due to cool ambient air (see Fig. 5.3)
and frequent windy conditions in coastal locations.

TABLE VII.8: Comparison of energy yields (in kWh/kWP) calculated from the MPPT
and the 1-minute data set (not corrected for cable losses).
Module MPPT Yield 1-minute Data Set Δ [%]
BC aged (‘BP’) 1101 1115 –1.3
mc-Si aged (‘GPV’) 1095 1106 –1.0
IBC (‘Sunc.’) 1121 1134 –1.2
‘RS mono’ 1105 1118 –1.2
‘RS poly’ 1114 1124 –0.9
HIT 1120 1135 –1.3
‘ESS 1’ 1111 1124 –1.2
‘ESS 2’ 1107 1119 –1.1
CIGS 1115 1127 –1.1

For the c-Si modules and the CIGS module installed in Grimstad, Table VII.8
compares the energy yields in year 2011 (estimated from the ‘real-time’ operating
powers recorded during the MPPT between the I-V curve sweeps) to those estimated
from the maximum powers extracted from I-V curves (the 1-minute data set). The latter
values are lower than the corresponding ones given in Table VII.7, because the results
257

in Table VII.8 are not corrected for resistive losses in the additional cables and
connectors. The MPP tracking efficiency for each module cannot be determined by
simply dividing the value in the 1st column in Table VII.8 by the value in the 2nd
column. This is because sampling of module’s maximum power with 1-minute
resolution introduces additional uncertainty to the estimated final yield, as already
discussed in Section 4.6 in the context of estimating the monthly and annual solar
irradiation. To a 0th approximation, a PV module’s maximum power is proportional to
the in-plane irradiance. Therefore, the extra uncertainty of its estimated energy yield due
to 1-minute sampling is of similar order of magnitude to that of the annual irradiation
(i.e. well below 1 % when averaged over the whole year). It is concluded that the MPP
tracking algorithm implemented in the data acquisition software is very efficient (with
98 % or higher annual efficiency).

7.5 CHAPTER SUMMARY


This chapter of the thesis provides input to the testing and validation of the
performance model for c-Si PV devices adopted by the European Commission’s Joint
Research Centre (JRC). Averaged model coefficients are derived from outdoor data of
seven modules made with screen-printed cells. The model is found to predict the annual
energy yield very accurately. The measured annual relative efficiencies of new
specimens are among the highest in Europe. This is attributed to rather low operating
temperatures of the modules in Southern Norway due to the cool ambient air and
frequent windy conditions in coastal locations. The first two model coefficients, which
predetermine the irradiance dependency of efficiency at module temperature of 25°C,
are expressed through I-V curve parameters assuming a one-exponential equation. For
the first coefficient (which defines a module’s performance at intermediate irradiances),
predicted and fitted values agree well for most of the tested PV modules. The derived
theoretical relationships provide module manufacturers with quantitative guidelines for
achieving high annual energy yields of their products by optimization of device
parameters. Future work on the performance model may include analysis of the relative
significance of the last three coefficients for the studied c-Si and CIGS modules, as well
as for other PV types that will be acquired.
258

CHAPTER 8 CONCLUSIONS
The general aim of investigating photovoltaic technology as a potential source of
electricity in Southern Norway, and the sun as the provider of energy, has been reached.
A methodology for testing of such technologies in-situ in the local context has been
demonstrated, and interesting peculiarities regarding the solar resource have been
discovered and documented in this work. In addition, the total irradiation and the energy
yields from the various technologies have been recorded for a full year (2011), and are
presented in the thesis. The results include the irradiation profile and an estimate of the
cloud ‘resource’. Measurements show that the energy from the sun was 1.20 MWh/m2
at the test site in Grimstad for the year 2011, whereas the estimate from the PVGIS is
1.04 MWh/m2. It has been derived in this work that the total loss in energy due to cloud
cover was 44 % for the year 2011. These values will naturally vary somewhat from year
to year, but are nevertheless a starting point for a long-term investigation of the solar
resource in Southern Norway, and also a good starting point for increased end use of
photovoltaic technology in Norway.
The other general objective of this work, to identify and understand causes for
performance differences between various crystalline-silicon (c-Si) PV technologies, has
also been reached, and the results are documented in the thesis. Some major findings are
repeated below.
Ten c-Si PV modules based on different cell designs and one thin-film CIGS
module have been characterized. Their performances over a practical range of
irradiances and operating temperatures have been modeled in terms of relative
efficiency with respect to STC. A prominent performance model has been used to fit
experimental data, thus obtaining a set of six empirical coefficients for each module.
For the seven c-Si modules with generic screen-printed cells, a representative set of
averaged coefficients has been obtained and compared to results for the same
technology found in the literature. Agreement has been observed only for the third
model coefficient, which is identical to the relative temperature coefficient of the
maximum power at STC. This coefficient is usually very similar for modules of this
type, and hence the result is not surprising. However, notable disagreements have been
obtained for the remaining five coefficients, which in addition have varied significantly
among the modules tested in the present work. Also the irradiance dependencies of the
efficiencies of some modules of the same type have shown very different behavior at the
intermediate and low illumination levels. The conclusion is that the present level of
understanding of PV power performance as given in the available scientific literature is
not complete. This work attempts to fill some of the gaps by giving quantitative
physical explanations of the differences.
259

8.1 MAIN SCIENTIFIC CONTRIBUTIONS


In this author’s assessment, the main scientific contributions of the present work
are as follows:
 A methodology for in-situ testing of photovoltaic modules that includes ‘real-
time’ irradiance data with sub-second resolution, and I-V curves containing a
few thousand data points recorded in as short a time as possible. These high-
quality data have been essential for achieving a deeper understanding of the
solar resource, in particular overirradiance events, and for understanding and
identifying parameters of the photovoltaic modules. The methodology is
presented in detail in Chapter 3.
 One full year of measured solar irradiance and energy yield from the modules
tested in Grimstad. The results are presented in detail in Chapters 4 and 7.
 The identification of peak irradiance values that occur naturally on tilted, south-
facing surfaces which are higher than those assumed by a large part of the
European outdoor PV testing community and by some irradiance sensor makers.
The maximal recorded overirradiance magnitude was 1528 W/m2. Thus, the true
range of possible outdoor operating conditions of PV modules and systems has
been brought to the attention of the scientific community. The results are
presented in Chapter 4. The author believes that, as a result of the effective
communication of these findings, more solar energy researchers will adopt either
sub-second or 1-second resolution in their measurements, and many new studies
on overirradiances from different parts of the Globe will be reported in future.
 Methodologies for parameter identification of photovoltaic modules. These
include two new methods for determination of the lumped series resistance RS,
the ideality factor n and the reverse saturation current I0 from one I-V curve.
The methods are presented in Chapter 6.
 An improvement of an existing differential technique for I-V data analysis by
defining a RS-corrected local ideality factor and the corresponding m-V and m-I
plots, and the demonstration of the advantages of the novel plots over the classic
ones. These findings are presented in Chapter 6.
 A proposal of an equation for ECT calculation for PV devices with varying
ideality factors. This innovation is particularly useful in the case of thin-film
CIS/CIGS and CdTe modules (usually having a glass-glass structure), for which
the back-of-module temperature can differ significantly from that of the cells.
The equation is derived in detail in Chapter 5.
 The identification of limits of the applicability of the widely used 1- and 2-
exponential I-V curve models for different c-Si module types, as well as for thin-
film CIGS devices. This analysis is presented in Chapter 6.
 The coefficient sets of a widely used PV performance model for eight c-Si PV
modules and one CIGS module. These coefficients are useful for the validation
260

and testing of the model. New insight into the efficiency’s dependency on
irradiance has been achieved by relating two of the main model coefficients to
the I-V curve parameters. The individual effects of series resistance and ideality
factor on the overall performance are revealed. Thus, the PV module
manufacturers are provided with a quantitative guideline for designing products
with better energy yields per kilowatt-peak of installed power. The results and
equations are presented and discussed in Chapter 7.

The author believes that the present work is useful for the PV research and
testing community, solar resource researchers, irradiance sensor makers, as well as to
designers of power electronics for PV systems, who should take particular interest in
overirradiance events. The work has also provided input to Task 13 within the
International Energy Agency’s Photovoltaic Power Systems Programme (IEA PVPS)
[Bey11, Bey12, IEA12].

8.2 MAIN CONTRIBUTIONS TO PV INDUSTRY


The present work can also be useful for the PV module manufacturers and PV
system builders. The author has derived equations relating the first two coefficients of
the performance model used in the PVGIS to I-V curve parameters such as the series
resistance and the ideality factor. The predicted and fitted values of the first coefficient
(which governs the efficiency’s behavior at intermediate irradiances) have been found
to agree very well for most of the studied c-Si modules. The disagreement obtained for
the second coefficient (governing the low-light performance) has been traced to the
model’s deviation from the true physics at low irradiances. Nevertheless, the equation
derived for that coefficient explains its always negative sign (regardless of the PV
technology) in both the present work and prior studies on the model. Thus, the author
has discovered quantitative and qualitative links between device physics and
performance. These findings may benefit PV makers and system designers, since they
show a path to maximizing the energy yield. For example, a clear conclusion is that a
PV system developer should prefer modules with screen-printed cells having only two
front busbars (despite the higher series resistance) to modules with cells having three
busbars.
261

APPENDIX A: List of papers

G.H. Yordanov, O.-M. Midtgård, T.O. Saetre, H.K. Nielsen, L.E. Norum:
Overirradiance (cloud enhancement) events at high latitudes. IEEE Journal of
Photovoltaics (invited publication), DOI: 10.1109/JPHOTOV.2012.2213581, 2012.

G.H. Yordanov, O.-M. Midtgård, T.O. Saetre: Series resistance determination and
further characterization of c-Si PV modules. Renewable Energy 46, pp. 72–80, 2012.

G.H. Yordanov, T.O. Saetre, O.-M. Midtgård: 100-millisecond resolution for accurate
overirradiance measurements. Submitted to the IEEE Journal of Photovoltaics, 2012.

G.H. Yordanov, O.-M. Midtgård, T.O. Saetre, H.K. Nielsen, L.E. Norum: Over-
irradiance (cloud enhancement) events at high latitudes. Presented at the 38th IEEE
PVSC, Austin, TX, US, 2012.

G.H. Yordanov, O.-M. Midtgård, T.O. Saetre, H.K. Nielsen, L.E. Norum: Over-
irradiance (cloud enhancement) events at high latitudes. Proc. 27th EUPVSEC,
Frankfurt, Germany, 2012, pp. 3709–3713.

G.H. Yordanov, O.-M. Midtgård, J.-O. Odden, T.O. Saetre: Test of the European JRC
performance model for c-Si PV modules. Proc. 27th EUPVSEC, Frankfurt, Germany,
2012, pp. 3306–3312.

G.H. Yordanov, M. Tayyib, O.-M. Midtgård, J.-O. Odden, T.O. Saetre: Test of the
European Joint Research Centre performance model for c-Si PV modules. Proc. 38th
IEEE PVSC, Austin, TX, US, 2012, pp. 002382–002387.

G.H. Yordanov, O.-M. Midtgård, T.O. Saetre: Equivalent cell temperature calculation
for PV modules with variable ideality factors. Proc. 38th IEEE PVSC, Austin, TX, US,
2012, pp. 000505–000508.

G.H. Yordanov, O.-M. Midtgård, T.O. Saetre: PV modules with variable ideality
factors. Presented at the 38th IEEE PVSC, Austin, TX, US, 2012, pp. 002362–002367.

G.H. Yordanov, O.-M. Midtgård: Modeling and parameter identification of crystalline


silicon photovoltaic devices. Proc. 37th IEEE PVSC, Seattle, WA, US, 2011, pp.
002972–002976.

G.H. Yordanov, O.-M. Midtgård: Physically-consistent parameterization in the


modeling of solar photovoltaic devices. Proc. 2011 IEEE PowerTech, Trondheim,
Norway, DOI: 10.1109/PTC.2011.6019232, 4 pp.
262

G.H. Yordanov, O.-M. Midtgård, L. Norum: Modeling and parameter identification of


crystalline silicon photovoltaic devices. Proc. 2011 ICCEP, Ischia, Italy, 2011, pp. 574–
577.

G.H. Yordanov, O.-M. Midtgård, T.O. Saetre: Extracting parameters from semi-log
plots of polycrystalline silicon PV modules outdoor data: Double-exponential model
revisited. Proc. 35th IEEE PVSC, Honolulu, HI, US, 2010, pp. 002756–002761.

G.H. Yordanov, O.-M. Midtgård, T.O. Saetre, J.O. Odden, T. Buseth, A.G. Imenes,
C.L. Nge, L. Norum: Compensated SoG-Si from a metallurgical route: High latitude
outdoor performance. Proc. 25th EUPVSEC, Valencia, Spain, 2010, pp. 4289–4293.

G.H. Yordanov, O.-M. Midtgård, T.O. Saetre: Two-diode model revisited: Parameters
extraction from semi-log plots of I-V data. Proc. 25th EUPVSEC, Valencia, Spain, 2010,
pp. 4156–4163.

T.O. Saetre, O.-M. Midtgård, G.H. Yordanov: A new analytical solar cell I–V curve
model. Renewable Energy 36, pp. 2171–2176, 2011.

O.-M. Midtgard, T.O. Saetre, G. Yordanov, A.G. Imenes, C.L. Nge: A qualitative
examination of performance and energy yield of photovoltaic modules in Southern
Norway. Renewable Energy 35, pp. 1266–1274, 2010.

H.G. Beyer, G.H. Yordanov: Stability of the performance of thin film modules during
one year of operation. Proc. 38th IEEE PVSC, Austin, TX, US, 2012, pp. 002391–
002394.

H.G. Beyer, A. McKinley, G.H. Yordanov: Gleichmäßigkeit des Betriebsverhaltens


von Dünnschicht-Modulen. Proc. 27. Symposium Photovoltaische Solarenergie, 29.2.-
02.3.2012, Bad Staffelstein, Germany, ISBN 978-3-941785-82-3.

A.G. Imenes, G.H. Yordanov, O.M. Midtgård, T.O. Sætre: Development of a test
station for accurate in situ I-V curve measurements of photovoltaic modules in Southern
Norway. Proc. 37th IEEE PVSC, Seattle, WA, US, 2011, pp. 003153–003158.

H.G. Beyer, G.H. Yordanov, O.-M. Midtgård, T.O. Saetre, A.G. Imenes: Contributions
to the knowledge base on PV performance: Evaluation of the operation of PV systems
using different technologies installed in Southern Norway, Proc. 37th IEEE PVSC,
Seattle, WA, US, 2011, pp. 003103–003108.

T.O. Saetre, G.H. Yordanov, O.-M. Midtgård, J.O. Odden: A theoretical study of the
effect of degradation caused by diffusion over a 30-year lifespan in c-Si PV devices.
Proc. 25th EUPVSEC, Valencia, Spain, 2010, pp. 1732–1736.
263

C.L. Nge, O.-M. Midtgård, G. Yordanov, L. Norum, T. O. Sætre: A new maximum


power point tracking approach for partial shading conditions. Proc. 25th EUPVSEC,
Valencia, Spain, 2010, pp. 4476–4478.

A.G. Imenes, T. Buseth, J.-O. Odden, G.H. Yordanov, O.-M. Midtgård: Field stations
in Norway and Kenya for comparative analysis of compensated and standard solar-
grade polysilicon modules. Proc. 25th EUPVSEC, Valencia, Spain, 2010, pp. 4294–
4300.

C.L. Nge, G. Yordanov, O.-M. Midtgård, T. O. Sætre, L. Norum: A comparative


simulation analysis of maximum power point tracking approaches. Proc. 24th
EUPVSEC, Hamburg, Germany, 2009, pp. 3668–3672.
264

APPENDIX B: Matlab™ code used to calculate the solar position, the clear-sky air
mass (AM) and the angle of incidence (AOI) of direct sunlight on the tested modules.

close all
clear all
R = dlmread('210312.txt','\t',1,0); % Update file name!!!
[N n] = size(R);
% Calculate decimal hour from the 1st column:
t = R(:,1);
d = 1e6*floor(t/1e6); % Date
t = floor((t–d)/1e4) + floor((t–d – floor((t–d)/1e4)*1e4)/100)/60 + (t–d – floor((t–
d)/1e4)*1e4 – 100*floor((t–d – 1e4*floor((t–d)/1e4))/100))/3600;

[Ele_ Azi_ AM_ AOI_] = EleAziAMAOI(t,d,N);

function [Ele_ Azi_ AM_ AOI_] = EleAziAMAOI(t,d,N)

Long = 8.58; % Eastern Longitude


Lat = 58.35; % Latitude
Lat = Lat*pi/180; % [rad]
Y = d(1)/1e6 – 100*floor(d(1)/1e8); % Year (within the century)
LeapY = (Y – 4*floor(Y/4)==0); % 1 for a leap year, otherwise 0
delta = Y+51;
leap = floor(delta/4);
Months = [31 28+LeapY 31 30 31 30 31 31 30 31 30 31];
Day = floor(d(1)/1e10);
Month = floor((d(1) – Day*1e10)/1e8);
DayNo = Day + sum(Months(1:(Month))) – Months(Month); % Day of the year
TimeZone = 1;
ST2011 =
Y==11&&(Month>3||(Month==3&&Day>26))&&(Month<10||(Month==10&&Day<30
));
ST2012 =
Y==12&&(Month>3||(Month==3&&Day>24))&&(Month<10||(Month==10&&Day<28
));
SummerTime = ST2011 || ST2012; % 1 for Summer Time, 0 for Standard Time
hour_UT = t – SummerTime – TimeZone; % Universal time
JD = 2432916.5 + delta*365 + leap + DayNo + hour_UT/24; % Julian Date
nn = JD – 2451545.0;
gmst = 6.697375 + 0.0657098242*nn + hour_UT; % Greenwich mean sidereal time
gmst = gmst – 24*floor(gmst/24); % 0 <= gmst <= 24 h
lmst = gmst+Long/15; % [h] local mean sidereal time
lmst = lmst – 24*floor(lmst/24); % 0 <= lmst <= 24 h
ep = (23.439 – 4e-7*nn)*pi/180; % Obliquity of the ecliptic [rad]
LL = 280.46 + 0.9856474*nn; % Mean longitude
LL = LL – 360*floor(LL/360); % 0 <= LL <= 360°
265

gg = 357.528 + 0.9856003*nn; % Mean anomaly


gg = gg - 360*floor(gg/360); % 0 <= gg <= 360°
gg = gg*pi/180; % [rad]
eLo = LL + 1.915.*sin(gg) + 0.02.*sin(2*gg); % Ecliptic Longitude
eLo = eLo – 360*floor(eLo/360); % 0 <= eLo <= 360°
eLo = eLo*pi/180; % [rad]
ra = atan(cos(ep).*sin(eLo)./cos(eLo)); % Right Ascension
ra = ra*180/pi; % [deg]
for n=1:N
if cos(eLo(n))<0 % Determine the true quadrant
ra(n) = ra(n) + 180; % 0 <= ra <= 360°
end
end
ra = ra/15; % [h]
dec = asin(sin(ep).*sin(eLo));
ha = lmst – ra; % Hour angle [h]
for n=1:N
if ha(n) > 12
ha(n) = ha(n) – 24;
else
if ha(n) <= –12
ha(n) = ha(n) + 24;
end
end
end % –12 < ha <= 12 h
ha = 15*ha; % 15° in one hour
ha = ha*pi/180; % [rad]
Ele_ = asin(sin(dec).*sin(Lat)+cos(dec).*cos(Lat).*cos(ha)); % Solar elevation
Ele_Critical = asin(sin(dec)./sin(Lat));
Azi_ = asin(sin(ha).*cos(dec)./cos(Ele_)); % Solar azimuth
Azi_ = Azi_*180/pi; % [deg]
for n=1:N
if Ele_(n) < Ele_Critical(n) % Determine the true quadrant
if Azi_(n)<0
Azi_(n) = –180 – Azi_(n);
else
Azi_(n) = 180 – Azi_(n);
end
end
end
Ele_ = Ele_*180/pi; % [deg]
Alt = 60; % [m] Altitude of PV roof, UiA-Grimstad
AM_ = exp(–0.0001184*Alt)./((Ele_+6.07995).^(–1.6364)*0.50572 +
sin(Ele_.*pi/180)); % Air Mass (pressure-corrected)
Tilt = pi*39/180; % [rad] Average Tilt Angle PV Modules (39°)
AZmod = –7; % [deg] Module azimuth; –90°=East, 0°=South, +90°=West
266

AOI_ =
180/pi.*acos(cos(Tilt)*sin(pi.*Ele_/180)+sin(Tilt)*cos(pi.*Ele_/180).*cos(pi.*(Azi_–
AZmod)/180)); % [deg] Angle of incidence of direct sunlight on the modules
end
267

APPENDIX C: ECT error analysis with respect to the ideality factor n

The equation for ECT (in °C) as a function of n given in [IEC93], after some
manipulation, reads:
VOC2  VOC1 NS nk  273.15  G1 
T1   ln  
 q  G2 
ECT  (C.1)
N nk  G 
1  S ln  1 
q  G2 
Before the partial differentiation with respect to n, it is convenient to simplify
the above equation by adding 273.15°C on both sides:
V  VOC1
T1  273.15  OC2

ECT  273.15  (C.2)
NSnk  G1 
1 ln  
q  G2 
In the thesis, the used reference conditions (G1,T1) are the STC (G1 = G0 = 1
kW/m2; T2=25°C), whereas G (in kW/m2) is used instead of G2. Also, VOC and VOC,STC
are used instead on VOC2 and VOC1, respectively. Thus, Eq. (C.1) simplifies to a form
which is much easier to differentiate with respect to n:
V  VOC ,STC
298.15  OC

ECT  273.15  (C.3)
NSnk
1 ln G
q
Note the sign change in the denominator. An error Δn in the value used for the ideality
factor would result in the following error in ECT:
V  VOC ,STC
298.15  OC
  NS k 
ECTn   2  q ln G  n
 NSnk   
 1  q ln G 
 
(C.4)
NS k  ECT  273.15 ln G
 n
q  NSnk ln G
Note that Eq. (C.3) has been used in order to simplify the intermediate result in Eq.
(C.4). Note also that [IEC93] used 273 K (not 273.15 K) as the zero of the Celsius scale.
However, this small difference is of very little significance, causing only a 0.05 % error
of the thermal voltage defined at 0°C in the numerator in Eq. (C.1).
268

APPENDIX D: Matlab™ functions used to extract RS, n and I0 from one I-V curve

function [Rs_out Io n RR]=SeriesR(R,Isc,Ns,RsHi,T)


% R(:,1) is current; R(:,2) is voltage; R(1,1) is 0 Amps, R(1,2) is Voc;
RsLo = 0.0; % Lowest possible physical value of the series resistance
% RsHi = 0.6; % Can be higher for some modules! Too high value gives error!
Range = RsHi – RsLo;
K = 7; % Number of trial (mock) Rs values = K+1
[L m] = size(R);
% U_LowLim = (Voc+Vmp)/2/Ns
% Try a U_LowLim based on current, i.e. corresponding to Isc/2:
M=1; % Open circuit
while R(M,1) < Isc/2
M=M+1;
end
% U_HiLim corresponding to 0.6 Amps of current
N=1;
while R(N,1) < 0.6
N=N+1;
end
while Range > 0.00001
r = [];
Rs56 = [];
for k = 0:K
Rs = RsLo+Range*k/K;
Rs56 = [Rs56 Rs];
U = (R(:,2) + R(:,1)*Rs)/Ns;
U_LowLim = U(M); % Corresponding to current Isc/2
U_HiLim = U(N); % Corresponding to current 0.6 A
I = log(abs(Isc*ones(L,1) – R(:,1))/Isc);
h = createFitOC(U,I,U_LowLim,U_HiLim);
u = U(N:M); % The fitted portion
i56 = h( u );
r = [r sqrt((I(N:M) – i56)'*(I(N:M) – i56))]; % The Least Squared Error
end %for k=0:K
[RR n] = min(r);
Rs = Rs56(n);
r(n) = [];
Rs56(n) = [];
[RRR n] = min(r);
if Rs < Rs56(n)
RsLo = 2*Rs–Rs56(n);
RsHi = Rs56(n);
else
RsLo = Rs56(n);
RsHi = 2*Rs–Rs56(n);
end % if
269

Old = Range; % Check if the while loop is infinite


Range = RsHi–RsLo;
if Range == Old
break
end % if
end % while
Rs_out = (RsLo+RsHi)/2;
% Once we know the true Rs, we can fit the linearized semi-log plot and get n and Io:
U = (R(:,2) + R(:,1)*Rs_out)/Ns;
U_LowLim = U(M); % Corresponding to current Isc/2
U_HiLim = U(N); % Corresponding to current 0.6 A
I = log(abs(Isc*ones(L,1) – R(:,1))/Isc);
h = createFitOC(U,I,U_LowLim,U_HiLim);

Io = Isc*exp(h( 0 ));
n = –U(1)/h( 0 )*16020/(T+273.15)/1.38065; % Input T is in Centigrade!!!
RR = RR/(M–N); % A per-data-point least squares in order to allow
% comparison between different I-V curves
end % function SeriesR

function out=createFitOC(U56,I56,U_LowLim,U_HiLim)
% REQUIRES THE CURVE-FITTING TOOLBOX!!!
% Apply exclusion rule "U > U_LowLim":
ex_ = (U56 <= U_LowLim | U56 >= U_HiLim);
ok_ = isfinite(U56) & isfinite(I56);
if ~all( ok_ )
warning( 'GenerateMFile:IgnoringNansAndInfs', ...
'Ignoring NaNs and Infs in data' );
end
ft_ = fittype('poly1');
if sum(~ex_(ok_))<2 %% too many points excluded
error('Not enough data left to fit ''%s'' after applying exclusion rule ''%s''.','fit 1','U >
0.58')
else
cf_ = fit(U56(ok_),I56(ok_),ft_,'Exclude',ex_(ok_));
end
out=cf_;
end % function createFitOC
270

APPENDIX E: Matlab™ scripts used to determine RS(I) with the double-light method

clear all
Ns = 60; % Number of cells in series
% Temp. coefficients of Isc:
aSC = 0.00167; % [A/K]
% Temperature coefficients of Voc [Volt/Kelvin] at STC:
beta = –0.1238; % [V/K]
% Open-circuit voltage at STC:
Voc0 = 37.03 + 0.14; % Label value with correction
% Variable Ideality factors, n = n0 + a*(G-1000) + b*(TBACK-25°):
n0 = 1.182; a = 0.45E–4; b = 0.0;
R = dlmread('210611.txt','\t',1,0); % Daily parameters
[N n] = size(R);
Gc = R(:,2); % Irradiance, amplified mc-Si cell
Isc = R(:,37); % Short-circuit currents
T = R(:,41); % Back-of-module temperatures
Voc = R(:,38); % Open-circuit voltages
PM = R(:,40); % Maximum Powers
FF = 100*PM./Isc./Voc; % Fill Factor (FF)
G = Isc*1000/7.93; % Self-referenced (effective) irradiances from Isc
Gc = Gc*1.0272; % Correction derived at Air Mass 1.5±0.05
G = G*1.000; % Correction derived at Air Mass 1.5±0.05
% Equivalent Cell Temperatures (ECTs) for a variable ideality factor:
ECT_ = ECT(Voc,Voc0,beta,Ns,G,T,n0,a,b); % Function call
G = G./(1 + aSC*(ECT_ – 25)); % Temperature correction to self-irradiances
ECT_ = ECT(Voc,Voc0,beta,Ns,G,T,n0,a,b); % With T-corrected self-irradiances
I0 = []; Rs0 = []; % Store multiple Rs(I) here
files = dir('Raw*.lvm'); % Measured I-Vs
[NN p] = size(files);
if N~=NN
'The number of data rows in 210611.txt must equal the number of Raw*.lvm files!'
break
else
n = 1;
while Gc(n) < 900
n=n + 1;
end
while Gc(n) >= 900
R = RawImport(n – 1); % First file is named 'Raw.lvm'
R(:,1) = [];
Gcc = 1000*R(:,22)/7.8; % Irradiance during I-V sweep
K = numel(Gcc);
tt = (1:K)/K; % Normalized sweep time
h = SmoothGc(tt',Gcc); % Function call; requires curve fitting toolbox
Gcc = h( tt );
ISF = (max(Gcc) – min(Gcc))*2/mean(Gcc); % Stability factor
271

if ISF <= 0.003 && FF(n) >= 0.99*(FF(n – 1) + FF(n + 1))/2


VV = R(:,15)*59.6/9.95; % Higher irradiance I-V curve
II = 2*R(:,14); % Higher irradiance I-V curve
p = n + 1;
while Gc(p) >= Gc(n) – 50
p=p + 1;
end
while Gc(p) >= 0.9*Gc(n)
if abs(ECT_(n) – ECT_(p)) >= 0.1 % [°C]
p=p + 1;
else
R = RawImport(p – 1);
R(:,1) = [];
Gcc = 1000*R(:,22)/7.8; % Irradiance during I-V sweep
K = numel(Gcc);
tt = (1:K)/K;
h = SmoothGc(tt',Gcc);
Gcc = h( tt );
ISF = (max(Gcc) – min(Gcc))*2/mean(Gcc);
if ISF <= 0.003 && FF(p) >= 0.99*(FF(p – 1) + FF(p + 1))/2
V = R(:,15)*59.6/9.95; % Lower irradiance I-V curve
I = 2*R(:,14); % Lower irradiance I-V curve
[n – 1 p – 1 round(Gc(n)) round(Gc(p))] % Progress indicator
[I Rs Isc] = Rs_2IV(I,V,II,VV);
I0 = [I0; I];
Rs0 = [Rs0; Rs];
end
p=p + 1;
end
end
end
n=n + 1;
end
end % if
save Rs_over_I.mat I0 Rs0
figure; plot(I0,Rs0,'b.'); grid on
axis([0 8 0 1]) % Ranges depend on module and irradiance
xlabel('I [A]'); ylabel('R_S [\Omega]')

function [I_ Rs Isc] = Rs_2IV(I,V,II,VV)


SmoothParam = 0.99;
h = SmartSmooth(V,I,SmoothParam); % Function call; requires curve fitting toolbox
i_ = h( V );
V = sort(V,'descend');
i_ = sort(i_);
tangent = Tangent(V,i_,5,3); % Function call
272

IscLo = tangent( 0 ); % Fitting near short circuit


[V i_ N1] = RemoveRedundantDP(V,i_); % Function call
hh = SmartSmooth(VV,II,SmoothParam); % FUNCTION call, separate M-file!!!
ii = hh( VV );
VV = sort(VV,'descend'); % Fitting does not sort the voltages in descending order
ii = sort(ii); % Fitting does not sort the currents in ascending order
tangent = Tangent(VV,ii,5,3);
IscHi = tangent( 0 );
[VV ii N2] = RemoveRedundantDP(VV,ii);
j_ = IscLo – i_;
jj = IscHi – ii;
% The following requires the curve fitting toolbox:
fo_ = fitoptions('method','SmoothingSpline','SmoothingParam',1);
ok_ = isfinite(jj) & isfinite(VV);
if ~all( ok_ )
warning( 'GenerateMFile:IgnoringNansAndInfs',...
'Ignoring NaNs and Infs in data.' );
end
ft_ = fittype('smoothingspline');
h = fit(jj(ok_),VV(ok_),ft_,fo_);
v = h( j_ );
ii = IscHi – j_;
dV = V – v;
Rs = dV/(IscHi – IscLo);
I_ = (i_ + ii)/2;
Isc = (IscLo + IscHi)/2;
end
273

APPENDIX F: Derivation of Eq. (6.5) from Eq. (18) in [Gre03]

Green’s Eq. (18) reads [Gre03]:

f df
 d
nOC  (F.1)
kT 1 dA f df
1
q A dV  d

where the brackets are defined as weighting the quantity of interest. For example,

 x i A i T i  i  dfi d i 
N

 x  i 1
(F.2)
 A i T i  i  dfi d i 
N
i 1

The quantities Ai(V) and T i account for the voltage- and temperature-dependency of
the reverse saturation current of the i-th recombination mechanism in the PV cell. The
generalized I-V curve model introduced in [Gre03] reads
N
I  IL  V ,T    A i  V  T i fi  i  (F.3)
i 1

where IL is the photocurrent; fi is a linear function in the case of low-injection SRH


recombination and a square-root function for recombination in depletion regions; and

np  E 
i  2
exp   G ,i  (F.4)
nie  kT 

where n and p are the excess electron and hole density, respectively (only in the context
of this Appendix!); nie is the effective intrinsic carrier density; and EG,i is the bandgap
for the i-th recombination mechanism. At open circuit, Eq. (F.4) simplifies to

 qVOC  EG ,i 
 i  exp   (F.5)
 kT 

The numerator in Eq. (F.1) is the ‘raw’ ideality factor of the dominant
recombination process at open circuit [Gre03]. For low-injection SRH recombination,
f=αξ where α is a constant, and therefore

f d 1
 1 (F.6)
 df 

The corresponding weighting factor in Eq. (F.2) equals


274

df
AT   AT   AT f   (F.7)
d

which, according to Eq. (F.3), is the SRH recombination current. For depletion-region
recombination (which is negligible in c-Si cells), f    where χ is another constant,
and therefore

f d  2 
 2 (F.8)
 df  

The corresponding recombination current is ignored as negligibly small when weighting


the ideality factors as per Eq. (F.2). Therefore, the numerator in Eq. (F.1) equals the
ideality factor of low-injection SRH recombination. The bracketed term in the
denominator in Eq. (F.1) is being similarly weighted, and for the same reason the final
result contains only the SRH contribution:

1
nOC  (F.9)
kT 1 dA
1
q A dV
275

APPENDIX G: Effect of the temperature coefficient of ISC on the relative efficiency


ηREL

In Section 7.4, the temperature coefficient of each module’s ISC was neglected in
the analysis of the relative efficiency ηREL. The associated worst-case errors in ηREL are
here assessed in the cases of high and low irradiance. From Eq. (7.2), the error in the
relative efficiency due to an erroneous self-irradiance equals

REL   k1  k 4T'    ln G'   2  k 2  k 5T'  ln G'   ln G'  (G.1)

At irradiances close to GSTC, G’ is approximately 1 and therefore lnG’ is close


to 0. Thus, the 2nd term in the above equation loses significance at such levels of
illumination. The maximal values of T’=ECT–25°C in year 2011 for the modules
studied in the thesis have been close to 40°C, and therefore this value is used in the
worst-case error estimation. A rather high temperature coefficient of ISC of a c-Si
module of about +0.216 %/K would bias the calculated self-irradiance by +8.6 % at
T’=40°C. The resulting error Δ(lnG’) equals 0.086 at standard irradiance. The extreme
values obtained in this thesis for k1 (+0.029) and k4 (±0.00045 K-1), result in a worst-
case error ΔηREL=+0.004 at standard irradiance level. The latter value should be
compared to ηREL(G0,40°C)=1+40k3+402k6 which equals 0.786 when taking the
extreme values obtained for k3 and k6. Thus, the worst-case error at irradiances close to
the STC value amounts to only about +0.5 % and is negligibly small, even when
considering a module’s maximum possible power output at the present test location.
At irradiances as low as 50 W/m2, T’ typically does not exceed 0°C, but can be
as low as –40°C. In this extreme case, lnG’ ≈ –3 and hence both terms in Eq. (G.1) are
significant. A self-irradiance bias of –8.6 % translates into an error of –0.086 in lnG’.
Thus, the worst-case error in ηREL amounts to only –0.0013, which should be compared
to ηREL(50 W/m2, –40°C) = 0.9486. Again, the extreme values obtained for k1 through
k6 in the present work are used in the calculations. The relative error is only –0.14 %
which is negligible, especially when considering the small power output from a PV
module at such a low irradiance.
276

APPENDIX H: Derivation of the coefficients k1 and k2 of the relative efficiency model

The relative efficiency ηREL is defined in Eq. (7.1) in Chapter 7 and its model is
given in Eq. (7.2). For T = T0 = 25°C (i.e., T’ = T–T0 = 0°C), the model simplifies to

PMAX  G,25C G0
REL  G',T'  0   1  k 1 ln G'  k 2  ln G' 
2
(H.1)
G P0

where P0 ≡ PMAX,STC, G’ = G/G0, and G0 = 1000 W/m2. Assuming that ISC is


proportional to G, one may substitute ISC/ISC,STC for G/G0 at a given T. The following
assumptions are also made: IM = µISC where µ is a constant close to but less than unity
over the whole irradiance range of interest; the I-V curve between open circuit and the
MPP is governed by the 1-exponential model with zero shunt current (i.e., RP = ∞), as
well as constant RS and n. This gives:

V I R   V   ISCR S 
IM   ISC  ISC  I0 exp  M M S   ISC  I0 exp  M  (H.2)
 NSnv T   NSnv T 

 V 
0  I SC  I 0 exp  OC  (H.3)
 NSnv T 

Considering only cases with T = T0 = 25°C, one can define the constants
‘standard thermal voltage’ v0 = k(T0+273.15)/q and ‘standard reverse saturation current’
I00 = I0(T0). Eqs. (H.2) and (H.3), when applied at STC, give:

 V 
I 00  I SC ,STC exp   OC ,STC  (H.4)
 NSnv 0 

 VM ,STC   ISC ,STCR S  VOC ,STC 


 ISC ,STC  ISC ,STC  ISC ,STC exp   (H.5)
 NSnv 0 

VM ,STC  VOC ,STC  nNS v0 ln 1      RSISC ,STC (H.6)

For G ≠ G0, from Eqs. (H.2) and (H.4) one obtains:

 G 
 VM   ISC ,STC  VOC ,STC 
G G G0
 ISC ,STC  ISC ,STC  ISC ,STC exp   (H.7)
G0 G0  NSnv 0 
 
 
277

VM  VOC ,STC  nNS v0 ln 1     nNS v0 ln G'  ISC ,STCRSG' (H.8)

One must express ηREL(G’,T’=0) in Eq. (H.1) as a polynomial with respect to


lnG’, and then identify k1 and k2 as the multipliers of lnG’ and (lnG’)2, respectively:

G
 ISC ,STC VM  T  25C 
G0 G0 VM  T  25C 
REL  G',T'  0   (H.9)
 ISC ,STC VM ,STC G VM ,STC

Substituting Eqs. (H.6) and (H.8) in the above equation gives:

VOC ,STC  nNS v 0 ln 1     nNS v 0 ln G'   ISC ,STCR SG'


REL  G',T'  0  
VOC ,STC  nNS v 0 ln 1      R SISC ,STC
(H.10)
 ISC ,STCR S 1  G'   nNS v 0 ln G'
 1
VOC ,STC  nNS v 0 ln 1      R S ISC ,STC

One may substitute x = lnG’ and then expand 1–G’ ≡ 1–eX as a Maclaurin series
around x=0:

x 2 x3 x 4
1  e X  x     ... (H.11)
2! 3! 4!

For |lnG’| ≡ |x| ≪ 1, one may drop the terms with powers of x higher than two as
negligibly small in magnitude compared to x, which gives:

1
1  G '   ln G '   ln G ' 
2
(H.12)
2

The above expression is then substituted in Eq. (H.10):

REL  G',T'  0  1
 nN v S 0   ISC ,STCR S  ln G'  0.5 ISC ,STCR S  ln G' 
2

(H.13)
VOC ,STC  nNS v 0 ln 1      R SISC ,STC

By comparing Eqs. (H.13) and (H.1), one obtains (for |lnG’| ≪ 1):

nNS v 0   R S I SC ,STC
k1  (H.14)
VOC ,STC  nNS v 0 ln 1      R S ISC ,STC

0.5 R S ISC ,STC


k2  (H.15)
VOC ,STC  nNS v 0 ln 1      R S ISC ,STC
278

By substituting Eq. (H.6) and IM,STC = µISC,STC in the two expressions above,
one obtains simplified (and more intuitional) formulae:

nNS v 0  R S I M ,STC
k1  (H.16)
VM ,STC

I M ,STC
k2   RS (H.17)
2VM ,STC

as well as the relationship:

nNS v 0
k 1  2k 2  (H.18)
VM ,STC
279

APPENDIX I: List of symbols

<nRAW> In the wider context of Eq. (6.5) and Appendix F, the weighted “raw”
ideality factors of all the current-conduction mechanisms in a PV device [Gre03]
α A constant used in the model of the reverse saturation current in Eq. (6.12b)
αSC Temperature coefficient of the short-circuit current
a In Eq. (5.13), a parameter describing the ideality factor dependency on
irradiance; in IEC 60904-5:2011, the “diode thermal voltage” used for
calculation of the ECT of a PV device from its open-circuit voltage; in the
context of I-V curve interpolation/extrapolation after [Tsu09], the
interpolation/extrapolation factor
A Usually, the total area of a PV device; in Eq. (5.15), a constant having units of
temperature; in Eq. (6.3), the total area under the I-V curve
A(V), A In Eq. (6.5) and in Appendix F, the voltage-dependent multiplier in the
reverse saturation current [Gre03]
AM Air mass
AOI Angle of incidence of the direct (beam) component of sunlight on a PV device
with respect to the normal
β Temperature coefficient of the open-circuit voltage
b A parameter describing the irradiance dependency of the ideality factor in Eq.
(5.13)
B In Eq. (5.15), a constant used to model the low-irradiance behavior of the ECT;
in Eq. (6.12a), a constant used to model the temperature dependency of I0
c A constant used in Eq. (5.3) which relates ECT to the back-of-module
temperature and the plane-of-module irradiance
cT A constant used in Eq. (3.1) which relates a module’s temperature to the ambient
temperature and the plane-of-module irradiance
C Thermal capacitance of a PV module used in Eq. (3.3)
C0 through C3 Coefficients of the Sandia PV performance model
ΔGMIN The minimum difference between the two illumination levels used in the
double-light method for evaluation of a current-dependent series resistance RS(I)
of a PV device
D The “diode thermal voltage” used in IEC 60904-5:1993 for calculation of the
ECT of a PV device from its open-circuit voltage
η The instantaneous efficiency of a PV device
η0 The efficiency of a PV device at standard testing conditions (STC)
ηREL The relative efficiency of a PV device with respect to η0, see Eq. (7.1)
ηREL,AVG The yearly-averaged ηREL of a PV device in a given location
ECT The equivalent cell temperature of a PV device defined in IEC 60904-5
280

EG The bandgap energy of a semiconductor material


FF The fill factor of an I-V curve
FF STC Fill factor at STC
γ Usually, the temperature coefficient of PMAX; in Eq. (6.12a), a dimensionless
exponent used to model the temperature dependency of I0
g In Eq. (5.15), a constant used to model the low-irradiance behavior of the ECT;
G Solar irradiance in the plane of a PV module
G0 Solar irradiance in the plane of a PV module at STC, 1000 W/m2 (1 sun)
G’ The ratio G/G0
G1, G2 The two irradiance levels in the equation for ECT calculation given in
IEC 60904-5:1993
GB The beam (direct) component of the solar irradiance
GD The diffuse component of the solar irradiance
GMIN, GMAX, GMEAN The minimal, maximal and average irradiance in the plane of a
PV module during an I-V curve sweep; In the context of Eq. (3.3), GMAX is the
upper threshold used in the quality control of irradiance data
GO In the context of Eq. (3.3), the normal extraterrestrial irradiance
GS The self-referenced irradiance estimated from the ISC of a PV device
GHI The global horizontal irradiance
iM,STC The ratio IM,STC/ISC,STC, a figure of merit proposed in [Car07]
I The terminal current of a PV device
I0 The reverse saturation current of a PV device used in the 1-exponential I-V
curve model
I00 The reverse saturation current of a PV device at STC
I0,STC The same as I00
I01, I02 The reverse saturation currents of a PV device used in the 2-exponential
I-V curve model
I 1 , I2 , I 3 In the context of I-V curve interpolation/extrapolation after [Tsu09]: the
currents at the corresponding points in the two reference curves and in the
translated curve
ID The diode current of a PV device, ISC – I
IL The photocurrent of a PV device
IM The terminal current of a PV device at the maximum power point (MPP)
IM,STC The MPP current at STC
IP In general, the total shunt current in a PV device; in the context of Fig. 6.6, the
apparent shunt current
281

ISC The short-circuit current


ISC1, ISC2 The short-circuit currents of the two I-V curves used in the context of:
the double-light method for evaluation of a current-dependent series resistance
RS(I) of a PV device; IEC 60904-5; and I-V curve interpolation/extrapolation
after [Tsu09]
ISC,STC The short-circuit current at STC
ISF The irradiance stability factor during an I-V curve sweep
J The terminal current density of a PV device
J0 The reverse saturation current density of a PV device used in the 1-exponential
J-V curve model
J01, J02 The reverse saturation current densities of a PV device used in the 2-
exponential J-V curve model
J0k The reverse saturation current densities of a PV device used in a generalized
multi-exponential J-V curve model where k = 1, 2, …, K is an integer number
JL The photocurrent density of a PV device
JSC The short-circuit current density of a PV device
k The Boltzmann’s constant, 1.38065×10-23 J/K
k1 through k6 The coefficients of the performance model given in Eq. (7.2)
k 1+ In Eq. (7.5), a non-physical quantity used to assess the discrepancy between the
predicted and empirical value of the coefficient k1
kSC A proportionality coefficient relating ISC to G (in the case of linearity)
λ In Eq. (6.12a), a dimensionless constant used to model the temperature
dependency of I0
L The number of data points in a daily irradiance profile (in the 1-minute data set)
µ The ratio IM/ISC
m The classic local ideality factor along the I-V curve of a PV device defined by
Eq. (6.7)
mSR The RS-corrected local ideality factor defined by Eq. (6.11)
n Usually, the ideality factor of a PV device used in the 1-exponential I-V (J-V)
curve model; In semiconductor physics, the concentration of free electrons
n0 The ideality factor at standard irradiance G0 and TMOD = 25°C defined in Eq.
(5.13)
n1, n2 The ideality factors of a PV device used in the 2-exponential I-V (J-V)
curve model
nk The ideality factors of a PV device used in a generalized multi-exponential J-V
curve model where k = 1, 2, …, K is an integer number
282

nRAW In Eq. (6.5) and in Appendix F, the “raw” ideality factor of a current-conduction
mechanism in a PV device [Gre03]
N A number of irradiance-sensitive devices (irradiance sensors and PV modules)
used in the in-house calibration procedure described in Section 4.7
NS The number of series-connected cells in a PV module
p In semiconductor physics, the concentration of free holes; in the context of
Matlab’s Curve Fitting Toolbox, the smoothing parameter (0 ≤ p ≤ 1)
P The output power of a PV device
P0 The output power maximum of a PV device at STC
Pin The solar power incident on a PV device (= GA)
PMAX The output power of a PV device at the MPP
PMAX,SIT The MPP power of a PV device at standard irradiance and (module)
temperature (SIT); assumed equal to PMAX,STC when using the self-referenced
irradiance calculated from the ISC
PMAX,STC The same as P0
q The elementary charge, 1.602×10-19 C
r The reflectance of a PV module
2
r In the context of part 5.5.2, the goodness of a linear fit
R General notation of a resistance (either electrical or thermal)
RMSE The root mean square error of a linear fit in the context of part 5.5.2
RMSE0 The optimal RMSE in the context of part 5.5.2
RP The total parallel (shunt) resistance of a PV device
RS The total series resistance of a PV device; In Figs. 6.8 through 6.10, a trial
(mock) value
RS,DARK The total series resistance of a PV device in the dark
RS,LIGHT The total series resistance of a PV device under illumination
σ The standard deviation of the calibration errors of all the irradiance sensors with
a calibration uncertainty of ±3 %;
σG The standard deviation of the irradiance during an I-V curve sweep
σk1 trough σk6 The standard deviations of the coefficients k1 through k6 of the
JRC performance model given in Eq. (7.2)
θZ The solar zenith angle
τ The time constant of the equivalent circuit given in Fig. 3.25 (=RC)
τe The lifetime of excess free electrons in a semiconductor
τp The lifetime of excess free holes in a semiconductor
t The time
T The operating temperature of a PV device
283

T0 The operating temperature of a PV device at STC (25°C)


T’ T – T0
T1, T2 The two module temperatures in the equation for ECT calculation (ECT ≡ T2)
given in IEC 60904-5:1993
TA The ambient air temperature
TMOD The back-of-module temperature
U95 The uncertainty of a variable’s estimate at 95 % confidence level
v0 The thermal voltage of a PV device at STC, 25.7 mV
vM,STC The ratio VM,STC/VOC,STC, a figure of merit proposed in [Car07]
V The terminal voltage of a PV device
VA The ratio A/ISC in Eq. (6.3)
VM The terminal voltage of a PV device at the MPP
VM,STC The MPP voltage of a PV device at STC
VOC The open-circuit voltage
VOC1, VOC2 The open-circuit voltages at irradiances G1 and G2, respectively used in
the equation for ECT calculation given in IEC 60904-5:1993
VOC,STC The open-circuit voltage of a PV device at STC
VT The thermal voltage of a PV device, kT/q (where T is in absolute units)
284

BIBLIOGRAPHY
[Abe93a] A.G. Aberle, S.R. Wenham, and M.A. Green, “A new method for accurate
measurements of the lumped series resistance of solar cells”, Proc. 23rd IEEE
PVSC, Louisville, pp. 133–139, 1993.

[Abe93b] A.G. Aberle, S.J. Robinson, A. Wang, J. Zhao, S.R. Wenham, and M.A.
Green, “High-efficiency silicon solar cells: Fill factor limitations and non-ideal
diode behaviour due to voltage-dependent rear surface recombination
velocity”, Progress in Photovoltaics: Research and Applications 1, pp. 133–
143, 1993.

[Abe00] A.G. Aberle, “Surface passivation of crystalline silicon solar cells: A review”,
Progress in Photovoltaics: Research and Applications 8, pp. 473–487, 2000.

[Abe03] M.A. Abella, E. Lorenzo, and F. Chenlo, “Effective irradiance estimation for
PV applications”, Proc. 3rd IEEE WCPEC, Osaka, vol. 2, pp. 2085–2089,
2003.

[Ale11] G.B. Alers, J. Zhou, C. Deline, P. Hacke, and S.R. Kurtz, “Degradation of
individual cells in a module measured with differential IV analysis”, Progress
in Photovoltaics: Research and Applications 19, pp. 977–982, 2011.

[Alm05] J. Almorox, C. Hontoria, and M. Benito, “Statistical validation of daylight


definitions for estimation of global solar radiation in Toledo, Spain”, Energy
Conservation and Management 46, pp. 1465–1471, 2005.

[Alt96] P.P. Altermatt, G. Heiser, A.G. Aberle, A. Wang, J. Zhao, S.J. Robinson, S.
Bowden, and M.A. Green, “Spatially resolved analysis and minimization of
resistive losses in high-efficiency Si solar cells”, Progress in Photovoltaics:
Research and Applications 4, pp. 399–414, 1996.

[Alt02] P.P. Altermatt, A.G. Aberle, J. Zhao, A. Wang, and G. Heiser, “A numerical
model of p-n junctions bordering surfaces”, Solar Energy Materials and Solar
Cells 74, pp. 165–174, 2002.

[Alt05] P.P. Altermatt, “Silicon solar cells”, in: “Optoelectronic Devices”, J. Piprek,
Ed., pp. 313–341, DOI: 10.1007/0-387-27256-9_11, 2005.

[And00] D.G. Andrews, “An introduction to atmospheric physics”, Cambridge


University Press, Cambridge, p. 5, 2000.

[And12] J. Andersen (Instrumenttjenesten AS) and T. Rafoss (Bioforsk), private


communication dated 01/03/2012.
285

[App93] J. Appelbaum, A. Chait, and D. Thompson, “Parameter estimation and


screening of solar cells”, Progress in Photovoltaics: Research and Applications
1, pp. 93–106, 1993.

[Ara82] G.L. Araujo and E. Sánchez, “A new method for experimental determination of
the series resistance of a solar cell”, IEEE Transactions on Electron Devices
29, pp. 1511–1513, 1982.

[Ara86] G.L. Araújo, A. Cuevas, and J.M. Ruiz, “The effect of distributed series
resistance on the dark and illuminated current-voltage characteristics of solar
cells”, IEEE Transactions on Electron Devices 33, pp. 391–401, 1986.

[Bas94] P.A. Basore, “Defining terms for crystalline silicon solar cells”, Progress in
Photovoltaics: Research and Applications 2, pp. 177–179, 1994.

[Bel11] P.-R. Beljean, A. Lo, M. Despeisse, Y. Riesen, Y. Pelet, V. Fakhfouri, N.


Wyrsch, and C. Ballif, “I/V measurement of high capacitance cells with
various methods”, Proc. 26th EUPVSEC, Hamburg, pp. 3275–3278, 2011.

[Bet04] T.R. Betts, R. Gottschalg, and D.G. Infield, “Spectral irradiance correction for
PV system yield calculations”, Proc. 19th EUPVSEC, vol. 2, Paris, pp. 2533–
2536, 2004.

[Bet09] T.R. Betts, S.R. Wenham, and R. Gottschalg, “Outdoor characterization of PV


modules for energy rating”, Proc. 24th EUPVSEC, Hamburg, pp. 3471–3474,
2009.

[Bey11] H.G. Beyer, G.H. Yordanov, O.-M. Midtgård, T.O. Saetre, and A.G. Imenes,
“Contributions to the knowledge base on PV performance: Evaluation of the
operation of PV systems using different technologies installed in Southern
Norway”, Proc. 37th IEEE PVSC, Seattle, pp. 003103–003108, 2011.

[Bey12] H.G. Beyer and G.H. Yordanov, “Stability of the performance of thin film
modules during one year of operation”, Proc. 38th IEEE PVSC, Austin, pp.
002391–002394, 2012.

[Bio12] Bioforsk AgroMetBase, http://lmt.bioforsk.no/agrometbase/getweatherdata.php,


last accessed 27/02/2012.

[Bla89] A.W. Blakers, A. Wang, A.M. Milne, J. Zhao, and M.A. Green, “22.8%
efficient silicon solar cell”, Applied Physics Letters 55, pp. 1363–1365, 1989.

[Ble08] B. Bletterie, R. Bründlinger, and H. Häberlin, “Redefinition of the European


efficiency – finding the compromise between simplicity and accuracy”, Proc.
23rd EUPVSEC, Valencia, pp. 2735–2742, 2008.
286

[Bou06] A. Bouthors, F. Neyret, and S. Lefebvre, “Real-time realistic illumination and


shadowing of stratiform clouds”, Proc. Eurographics Workshop on Natural
Phenomena, pp. 41–50, 2006.

[BPS10] BP Solar’s distributor in UK Wind and Sun Ltd., private communication dated
04/02/2010.

[Bra08] A.F.B. Braga, S.P. Moreira, P.R. Zampieri, J.M.G. Bacchin, and P.R. Mei,
“New processes for the production of solar-grade polycrystalline silicon: A
review”, Solar Energy Materials and Solar Cells 92, pp. 418–424, 2008.

[Bra10] V. Lo Brano, A. Orioli, G. Ciulla, and A. Di Gangi, “An improved five-


parameter model for photovoltaic modules”, Solar Energy Materials and Solar
Cells 94, pp. 1358–1370, 2010.

[Bre03] O. Breitenstein, J.P. Rakotoniaina, and M.H. Al Rifai, “Quantitative evaluation


of shunts in solar cells by lock-in thermography”, Progress in Photovoltaics:
Research and Applications 11, pp. 511–526, 2003.

[Bre04] O. Breitenstein, J.P. Rakotoniaina, M.H. Al Rifai, and M. Werner, “Shunt types
in crystalline silicon solar cells”, Progress in Photovoltaics: Research and
Applications 12, pp. 529–538, 2004.

[Bre06] O. Breitenstein, P. Altermatt, K. Ramspeck, and A. Schenk, “The origin of


ideality factors n>2 of shunts and surfaces in the dark I-V curves of Si solar
cells”, Proc. 21st EUPVSEC, Dresden, pp. 625–628, 2006.

[Bre07a] O. Breitenstein, “Understanding shunting mechanisms in silicon cells: A


review”, Proc. 17th NREL Workshop, pp. 61–70, 2007.

[Bre07b] O. Breitenstein, J. Bauer, and J.P. Rakotoniaina, “Material-induced shunts in


multicrystalline silicon solar cells”, Semiconductors 41, pp. 440–443, 2007.

[Bre10a] C. Breyer and J. Schmid, “Global distribution of optimal tilt angles for fixed
tilted PV systems”, Proc. 25th EUPVSEC, Valencia, pp. 4715–4721, 2010.

[Bre10b] O. Breitenstein, J. Bauer, P.P. Altermatt, and K. Ramspeck, “Influence of


defects on solar cell characteristics”, Solid State Phenomena 156-158, pp. 1–
10, 2010.

[Bre10c] O. Breitenstein, A. Khanna, and W. Warta, “Quantitative description of dark


current–voltage characteristics of multicrystalline silicon solar cells based on
lock-in thermography measurements”, Physica Status Solidi A 207, pp. 2159–
2163, 2010.
287

[Bre11a] O. Breitenstein, “Nondestructive local analysis of current–voltage


characteristics of solar cells by lock-in thermography”, Solar Energy Materials
and Solar Cells 95, pp. 2933–2936, 2011.

[Bre11b] O. Breitenstein, “Separate imaging of space charge and bulk recombination


current distribution by lock-in thermography”, Proc. 26th EUPVSEC,
Hamburg, pp. 1179–1181, 2011.

[Bub12] BubbleLevel freeware for Windows Mobile Phone,


http://www.freewarepocketpc.net/ppc-download-bubblelevel.html, last
accessed 23/01/2012.

[Bur96] A.R. Burgers, J.A. Eikelboom, A. Schönecker, and W.C. Sinke, “Improved
treatment of the strongly varying slope in fitting solar cell I-V curves”, Proc.
25th IEEE PVSC, Washington, pp. 569–572, 1996.

[Bur05] B. Burger and R. Rüther, “Site-dependent system performance and optimal


inverter sizing of grid-connected PV systems”, Proc. 31st IEEE PVSC, Lake
Buena Vista, pp. 1675–1678, 2005.

[Bus11] T. Buseth (Elkem Solar AS), private communication dated 09/05/2011.

[Car07] C. Carrero, J. Amador, and S. Arnaltes, “A single procedure for helping PV


designers to select silicon PV modules and evaluate the loss resistances”,
Renewable Energy 32, pp. 2579–2589, 2007.

[Car09a] J. Carstensen, A. Abdollahinia, A. Schütt, and H. Föll, “Characterization of the


grid design by fitting of the distributed serial grid resistance to CELLO
resistance maps and global I-V curves”, Proc. 24th EUPVSEC, Hamburg, pp.
516–519, 2009.

[Car09b] L. Carnel, P.C. Hjemås, T. Lu, J. Nyhus, K. Helland, and Ø. Gjerstad,


“Influence of wafer quality on cell performance”, Proc. 34th IEEE PVSC,
Philadelphia, pp. 000036–000038, 2009.

[Cha85] J.-P. Charles, G. Bordure, A. Khoury, and P. Mialhe, “Consistency of the


double exponential model with the physical mechanisms of conduction for a
solar cell under illumination”, Journal of Physics D: Applied Physics 18, pp.
2261–2268, 1985.

[Cha87] D.S.H. Chan and J.C.H. Phang, “Analytical methods for the extraction of solar-
cell single- and double-diode model parameters from I-V characteristics”,
IEEE Transactions on Electron Devices 53, pp. 286–293, 1987.
288

[Che01] M. Chegaar, Z. Ouennoughi, and A. Hoffmann, “A new method for evaluating


illuminated solar cell parameters”, Solid-state Electronics 45, pp. 293–296,
2001.

[Che10] S. Chen, P. Li, D. Brady, and B. Lehman, “The impact of irradiance time
behaviors on inverter sizing and design”, Proc. 12th IEEE COMPEL Workshop,
DOI: 10.1109/COMPEL.2010.5562387, 5 pp., 2010.

[Che11] Y. Chen, X. Wang, D. Li, R. Hong, and H. Shen, “Parameters extraction from
commercial solar cells I–V characteristics and shunt analysis”, Applied Energy
88, pp. 2239–2244, 2011.

[Chu11a] S.K. Chunduri, “Preparing for bigger loads”, Photon International, pp. 236–
259, June, 2011.

[Chu11b] S.K. Chunduri, “The quest for quasi”, Photon International, pp. 260–264,
June, 2011.

[Coo97] S. Coors and M. Böhm, “Validation and comparison of curve correction


procedures for silicon solar cells”, Proc. 14th EUPVSEC, Barcelona, pp. 220–
223, 1997.

[Cou06a] P.J. Cousins and J.E. Cotter, “The influence of diffusion-induced dislocations
on high-efficiency silicon solar cells”, IEEE Transactions on Electron Devices
53, pp. 457–464, 2006.

[Cou06b] P.J. Cousins, N.B. Mason, and J.E. Cotter, “Loss analysis of DSBC solar cells
on FZ(B), MCZ(B), CZ(Ga), and CZ(B) wafers”, IEEE Transactions on
Electron Devices 53, pp. 797–807, 2006.

[Cra95] J.C. Craparo and E.F. Thacher, “A solar-electric vehicle simulation code”,
Solar Energy 55, pp. 221–234, 1995.

[Cue98] J.A. del Cueto, “Method for analyzing series resistance and diode quality
factors from field data of photovoltaic modules”, Solar Energy Materials and
Solar Cells 55, pp. 291–297, 1998.

[Cue99] J.A. del Cueto, “Method for analyzing series resistance and diode quality
factors from field data Part II: Applications to crystalline silicon”, Solar
Energy Materials and Solar Cells 59, pp. 393–405, 1998.

[Cue08] J.A. del Cueto, S. Rummel, B. Kroposki, C. Osterwald, and A. Anderberg,


“Stability of CIS/CIGS modules at the outdoor test facility over two decades”,
Proc. 33rd IEEE PVSC, San Diego, 6 pp., DOI: 10.1109/PVSC.2008.4922772,
2008.
289

[Cue09] J.A. del Cueto, C.A. Deline, D.S. Albin, S.R. Rumme, and A. Anderberg,
“Striving for a standard protocol for preconditioning or stabilization of
polycrystalline thin film photovoltaic modules,” Proc. SPIE 2009 Solar Energy
and Technology Conference, San Diego, 12 pp., 2009.

[Cue10] J.A. del Cueto, C.A. Deline, S.R. Rummel, and A. Anderberg, “Progress
toward a stabilization and preconditioning protocol for polycrystalline thin-film
photovoltaic modules”, Proc. 35th IEEE PVSC, Honolulu, pp. 002423–002428,
2010.

[Das11] A.K. Das, “An explicit J–V model of a solar cell for simple fill factor
calculation”, Solar Energy 85, pp. 1906–1909, 2011.

[Das12] A.K. Das, “Analytical derivation of explicit J–V model of a solar cell from
physics based implicit model”, Solar Energy 86, pp. 26–30, 2012.

[Dau54] T.M. Dauphinee and H. Preston-Thomas, “A copper resistance temperature


scale”, The Review of Scientific Instruments 25, pp. 884–886, 1954.

[Dei64] D. Deirmendjian, “Scattering and polarization properties of water clouds and


hazes in the visible and infrared”, Applied Optics 3, pp. 187-196, 1964.

[Del11] C.A. Deline, J.A. del Cueto, D.S. Albin, and S.R. Rummel, “Metastable
electrical characteristics of polycrystalline thin-film photovoltaic modules upon
exposure and stabilization”, http://www.nrel.gov/docs/fy11osti/52441.pdf, 14
pp., last accessed 15/12/2011.

[Dir12] Direct Industry Catalog Search, “BP MSX 120: 120-Watt multicrystalline
photovoltaic module”, http://pdf.directindustry.com/pdf/bp-solar/bp-msx-120-
solar-module/Show/15873-68158.html, last accessed 20/03/2012.

[Dit10] S. Dittmann, G. Friesen, S. Williams, T.R. Betts, R. Gottschalg, H.G. Beyer, A.


Guerin de Montgareuil, N.v.d. Borg, A.R. Burgers, T. Huld, B. Müller, C.
Reise, J. Kurnik, M. Topic, T. Zdanowicz, and F. Fabero, “Results of the 3rd
modelling round robin within the European project “Performance” –
comparison of module energy rating methods”, Proc. 25th EUPVSEC,
Valencia, pp.4333–4338, 2010.

[Dol11] J.A. Dolan, R. Lee, Y.-H. Yeh, C. Yeh, D.Y. Nguyen, S. Ben-Menahem, and
A.K. Ishihara, “Neural network estimation of photovoltaic I-V curves under
partially shaded conditions”, Proc. International Joint Conference on Neural
Networks, San Jose, pp. 1358–1365, 2011.

[Dom10] D. Dominé, A. Jagomägi, A.G. de Montgareuil, G. Friesen, E. Mõttus, H.-D.


Mohring, D. Stellbogen, T. Betts, R. Gottschalg, T. Zdanowicz, M. Prorok, F.
290

Fabero, D. Faiman, and W. Herrmann, “Uncertainties of PV module – long-


term outdoor testing”, Proc. 25th EUPVSEC, Valencia, pp. 4172–4178, 2010.

[Don10] M. Donovan, B. Bourne, and J. Roche, “Efficiency vs. irradiance


characterization of PV modules requires angle-of-incidence and spectral
corrections”, Proc. 35th IEEE PVSC, Honolulu, pp. 002301– 002305, 2010.

[Dun03] E.D. Dunlop, “Lifetime performance of crystalline silicon PV modules”, Proc.


3rd IEEE WCPEC, Osaka, vol. 3, pp. 2927–2930, 2003.

[Dun12] L. Dunn and M. Gostein, “Light soaking measurements of commercially


available CIGS PV modules”, Proc. 38th IEEE PVSC, Austin, pp. 001260–
001265, 2012.

[Dyk04] E.E. van Dyk and E.L. Meyer, “Analysis of the effect of parasitic resistances
on the performance of photovoltaic modules”, Renewable Energy 29, pp. 333–
344, 2004.

[Emc08] P. Emck and M. Richter, “An upper threshold of enhanced global shortwave
irradiance in the troposphere derived from field measurements in tropical
mountains”, Journal of Applied Meteorology and Climatology 47, pp. 2828–
2845, 2008.

[Eme03] K. Emery, “Measurement and characterization of solar cells and modules”, in:
Handbook of photovoltaic science and engineering, A. Luque and S. Hegedus,
Eds., Wiley, pp. 701–752, 2003.

[Esp09] B. Espinar, L. Ramírez, A. Drews, H.G. Beyer, L.F. Zarzalejo, J. Polo, and L.
Martín, “Analysis of different comparison parameters applied to solar radiation
data from satellite and German radiometric stations”, Solar Energy 83, pp.
118–125, 2009.

[EUR01] The European Parliament and the Council of the European Union, “Directive
2000/84/EC of the European Parliament and of the Council of 19 January 2001
on summer-time arrangements”, Official Journal of the European
Communities, pp. L 31/21 – L 31/22, 02/02/2001.

[Fan86] J.C.C. Fan, “Theoretical temperature dependency of solar cell parameters”,


Solar Cells 17, pp. 309–315, 1986.

[Fis00] B. Fischer, P. Fath, and E. Bucher, “Evaluation of solar cell J(V)-measurements


with a distributed series resistance model”, Proc. 16th EUPVSEC, vol. 2,
Glasgow, pp. 1365–1368, 2000.
291

[Fon11a] K.C. Fong, K.R. McIntosh, A.W. Blakers, and E.T. Franklin, “Series
resistance as a function of current and its application in solar cell analysis”,
Proc. 37th IEEE PVSC, Seattle, pp. 002257–002261, 2011.

[Fon11b] K.C. Fong, K.R. McIntosh, and A.W. Blakers, “Accurate series resistance
measurement of solar cells”, Progress in Photovoltaics: Research and
Applications, DOI: 10.1002/pip.1216, 10 pp., 2011.

[For95] W.C. Forsythe, E.J. Rykiel Jr., R.S. Stahl, H. Wu, and R.M. Schoolfield, “A
model comparison for daylength as a function of latitude and day of the year”,
Ecological Modelling 80, pp. 87–95, 1995.

[Fri04] K. Friestad, C. Zahedi, E. Enebak, M.G. Dolmen, J. Heide, K. Engvoll, T.


Buseth, R. Tronstad, C. Dethloff, K. Peter, R. Kopecek, and I. Melnyk, “Solar
grade silicon from metallurgical route”, Proc. 19th EUPVSEC, vol. 1, Paris, pp.
568–571, 2004.

[Fri07] G. Friesen, R. Gottschalg, H.G. Beyer, S. Williams, A.G. de Montgareuil, N.v.d.


Borg, W.G.J.H.M.v. Sark, T. Huld, B. Müller, A.C. de Keizer, and Y. Niu,
“Intercomparison of different energy prediction methods within the European
project “Performance” – results of the 1st round robin”, Proc. 22nd EUPVSEC,
Milan, pp. 2659–2663, 2007.

[Fri08] P. Sánchez-Friera, F. Ropero, B. Lalaguna, L.J. Caballero, and J. Alonso,


“Power losses in crystalline silicon PV modules due to cell interconnection”,
Proc. 23rd EUPVSEC, Valencia, pp. 2701–2704, 2008.

[Gar93a] H.P. Garg and S.N. Garg, “Measurement of solar radiation – I. Radiation
instruments”, Renewable Energy 3, pp. 321–333, 1993.

[Gar93b] H.P. Garg and S.N. Garg, “Measurement of solar radiation – II. Calibration
and standardization”, Renewable Energy 3, pp. 335–348, 1993.

[Gat11] S. Gatz, T. Dullweber, and R. Brendel, “Evaluation of series resistance losses in


screen-printed solar cells with local rear contacts”, IEEE Journal of
Photovoltaics 1, pp. 37–42, 2011.

[Geu06] N. Geuder and V. Quaschning, “Soiling of irradiation sensors and methods for
soiling correction”, Solar Energy 80, pp. 1402–1409, 2006.

[Glo08] T. Glotzbach, B. Schulz, M. Zehner, P. Fritze, M. Schlatterer, C. Vodermayer,


G. Wotruba, and M. Mayer, “Round-robin-test of irradiance sensors”, Proc.
23rd EUPVSEC, Valencia, pp. 3144–3147, 2008.
292

[Goe03] A. Goetzberger, C. Hebling, and H.-W. Schock, “Photovoltaic materials,


history, status and outlook”, Materials Science and Engineering R 40, pp. 1–
46, 2003.

[Goo09] C. Good, H. Persson, Ø. Kleven, W. Sulkowski, K. Ellingsen, and T. Boström,


“Solar cells above the Arctic Circle – measuring characteristics of solar panels
under real operating conditions”, Proc. 24th EUPVSEC, Hamburg, pp. 4094–
4098, 2009.

[Goo12] Google Earth web site, http://www.google.com/earth/index.html, last accessed


23/01/2012.

[Gos11] M. Gostein and L. Dunn, “Light soaking effects on photovoltaic modules:


Overview and literature review”, Proc. 37th IEEE PVSC, Seattle, pp. 003126–
003131, 2011.

[Got97] R. Gottschalg, M. Rommel, D.G. Infield, and H. Ryssel, “Comparison of


different methods for the parameter determination of the solar cell double
exponential equation”, Proc. 14th EUPVSEC, vol. 1, Barcelona, pp. 321–324,
1997.

[Got10] R. Gottschalg, “Performance characterization of photovoltaic modules”, Proc.


35th IEEE PVSC, Honolulu, pp. 001265–001270, 2010.

[Gow99] J.A. Gow and C.D. Manning, “Development of a photovoltaic array model for
use in power-electronics simulation studies”, IEE Proceedings on Electric
Power Applications 146, pp. 193–200, 1999.

[Gre86] M.A. Green, A.W. Blakers, S. Narayanan, and M. Taouk, “Improvements in


silicon solar cell efficiency”, Solar Cells 17, pp. 75–83, 1986.

[Gre95] M.A. Green, “Silicon solar cells: Advanced principles and practice", UNSW,
1995.

[Gre03] M.A. Green, “General temperature dependency of solar cell performance and
implications for device modelling”, Progress in Photovoltaics: Research and
Applications 11, pp. 333–340, 2003.

[Gre09] J. Greulich, M. Glatthaar, A. Krieg, G. Emanuel, and S. Rein, “JV


characteristics of industrial silicon solar cells: Influence of distributed series
resistance and Shockley Read Hall recombination”, Proc. 24th EUPVSEC,
Hamburg, pp. 2065–2069, 2009.
293

[Gre11] M.A. Green, “Learning experience for thin-film solar modules: First Solar, Inc.
case study”, Progress in Photovoltaics: Research and Applications 19, pp.
498–500, 2011.

[Gue02] C.A. Gueymard, D. Myers, and K. Emery, “Proposed reference irradiance


spectra for solar energy systems testing”, Solar Energy 73, pp. 443–467, 2002.

[Gue09] C.A. Gueymard, “Direct and indirect uncertainties in the prediction of tilted
irradiance for solar engineering applications”, Solar Energy 83, pp. 432–444,
2009.

[Haa11] C. Haase and C. Podewils, “More of everything”, Photon International, pp.


174–221, February, 2011.

[Hae01] H. Haeberlin, “Evolution of inverters for grid connected PV-systems from


1989 to 2000”, Proc. 17th EUPVSEC, Munich, pp. 426–430, 2001.

[Ham12] A. Hammer and H.G. Beyer, “Spatial and temporal variability in solar
radiation”, in: R. A. Meyers, (Ed.), “Encyclopedia of sustainability science and
technology”, ISBN 978-1-4419-0851-3, DOI: 10.1007/978-1-4419-0851-3, 16
pp., 2012.

[Han10] C.W. Hansen, J.S. Stein, and A. Ellis, “Statistical criteria for characterizing
irradiance time series”, SAND2010-7314, Sandia National Laboratories
Report, 2010.

[Han12] C.W. Hansen (Sandia National Laboratories), private communication dated


21/05/2012.

[Hao05] M. Haouari-Merbah, M. Belhamel, I. Tobías, and J.M. Ruiz, “Extraction and


analysis of solar cell parameters from the illuminated current–voltage curve”,
Solar Energy Materials and Solar Cells 87, pp. 225–233, 2005.

[Hei05] A.S.H.v.d. Heide, A. Schönecker, J.H. Bultman, and W.C. Sinke, “Explanation
of high solar cell diode factors by nonuniform contact resistance”, Progress in
Photovoltaics: Research and Applications 13, pp. 3–16, 2005.

[Her11] G. Hering, “Year of the tiger”, Photon International, pp. 186–218, March 2011.

[Hez09] R. Hezel, “Commercial high-efficiency silicon solar cells”, in: High-Efficient


Low-Cost Photovoltaics: Recent Developments, V. Petrova-Koch, Ed.,
Springer, pp. 95–100, 2009.

[Hin75] J. Hine and G.B. Wetherill, “A programmed text in statistics”, Chapman Hall,
London, 1975.
294

[Hir11] W.P. Hirshman, “Cloudy forecast”, Photon International, pp. 58–63, November
2011.

[Hof06] W. Hoffmann, “PV solar electricity industry: Market growth and perspective”,
Solar Energy Materials and Solar Cells 90, pp. 3285–3311, 2006.

[Hoy08] C. Hoyer-Klick, H.G. Beyer, D. Dumortier, M. Schroedter-Homscheidt, L.


Wald, M. Martinoli, C. Schilings, B. Gschwind, L. Menard, E. Gaboardi, L.
Ramirez-Santigosa, J. Polo, T. Cebecauer, T. Huld, M. Suri, M. de Blas, E.
Lorenz, R. Pfatischer, J. Remund, P. Ineichen, A. Tsvetkov, and J. Hofierka,
“Management and exploitation of solar resource knowledge”, Proc. EUROSUN
2008, 7 pp., 2008.

[HTC12] HTC HD2 – Full phone specifications, http://www.gsmarena.com/htc_hd2-


2957.php, last accessed 23/01/2012.

[Hul08] T. Huld, M. Šuri, and E.D. Dunlop, “Geographical variation of the conversion
efficiency of crystalline silicon photovoltaic modules in Europe”, Progress in
Photovoltaics: Research and Applications 16, pp. 595–607, 2008.

[Hul10] T. Huld, R. Gottschalg, H.G. Beyer, and M. Topič, “Mapping the performance
of PV modules, effects of module type and data averaging”, Solar Energy 84,
pp. 324–338, 2010.

[Hul11] T. Huld, G. Friesen, A. Skoczek, R.P. Kenny, T. Sample, M. Field, and E.D.
Dunlop, “A power-rating model for crystalline silicon PV modules”, Solar
Energy Materials and Solar Cells 95, pp. 3359–3369, 2011.

[Hul12] T. Huld, R. Müller, and A. Gambardella, “A new solar radiation database for
estimating PV performance in Europe and Africa”, Solar Energy 86, pp. 1803–
1815, 2012.

[Hus04] F. Huster, S. Seren, G. Schubert, M. Kaes, G. Hahn, and O. Breitenstein,


“Shunts in silicon solar cells below screen printed silver contacts”, Proc. 19th
EUPVSEC, vol. 1, Paris, pp. 832–835, 2004.

[Hyv03] J. Hyvärinen and J. Karila, “New analysis method for crystalline silicon cells”,
Proc. 3rd IEEE WCPEC, Osaka, vol. 2, pp. 1521–1524, 2003.

[IEA12] IEA PVPS, Task 13 “Performance and reliability of photovoltaic systems”,


http://www.iea-pvps.org/index.php?id=57, last accessed 26/06/2012.

[IEC93] IEC 60904-5, “Photovoltaic devices – Part 5: Determination of the equivalent


cell temperature (ECT) of photovoltaic (PV) devices by the open-circuit
voltage method”, 1st Ed., 1993.
295

[IEC11] IEC 60904-5, “Photovoltaic devices – Part 5: Determination of the equivalent


cell temperature (ECT) of photovoltaic (PV) devices by the open-circuit
voltage method”, 2nd Ed., 2011.

[Iga04] N. Igawa, Y. Koga, T. Matsuzawa, and H. Nakamura, “Models of sky radiance


distribution and sky luminance distribution”, Solar Energy 77, pp. 137–157,
2004.

[Ime11] A.G. Imenes, G.H. Yordanov, O.M. Midtgård, and T.O. Saetre, “Development
of a test station for accurate in situ I-V curve measurements of photovoltaic
modules in Southern Norway”, Proc. 37th IEEE PVSC, Seattle, pp. 003153–
003158, 2011.

[IPC07] Intergovernmental Panel on Climate Change, “Summary for Policymakers”, in:


Climate Change 2007: The Physical Science Basis, Cambridge University
Press, 2007.

[Ish12] A.K. Ishihara, G. Alers, and D. Liddell, “Real-time parameter estimation for PV
modules using a polymer dispersed liquid crystal active shading screen”,
presented at the 38th IEEE PVSC, Austin, 2012.

[Jag09] A. Jagomägi, E. Mõttus, D. Stellbogen, H.-D. Mohring, T. Betts, R.


Gottschalg, T. Zdanowicz, W. Kołodenny, M. Prorok, G. Friesen, I. Pola, D.
Dominé, A.G. de Montgareuil, F. Fabero, D. Faiman, and W. Herrmann,
“European network of PV outdoor testing – steps towards harmonized
procedures”, Proc. 24th EUPVSEC, Hamburg, pp. 3432–3438, 2009.

[Joh12a] J. Johnson, M. Montoya, S. McCalmont, G. Katzir, F. Fuks, J. Earle, A.


Fresquez, S. Gonzalez, and J. Granata, “Differentiating series and parallel
photovoltaic arc-faults”, Proc. 38th IEEE PVSC, Austin, pp. 000720–000726,
2012.

[Joh12b] J. Johnson, C. Oberhauser, M. Montoya, A. Fresquez, S. Gonzalez, and A.


Patel, “Crosstalk nuisance trip testing of photovoltaic DC arc-fault detectors”,
Proc. 38th IEEE PVSC, Austin, pp. 001383–001387, 2012.

[Joh12c] J. Johnson and J. Kang, “Arc-fault detector algorithm evaluation method


utilizing prerecorded arcing signatures”, Proc. 38th IEEE PVSC, Austin, pp.
001378–001382, 2012.

[Jor10] D.C. Jordan, R.M. Smith, C.R. Osterwald, E. Gelak, and S.R. Kurtz, “Outdoor
PV degradation comparison”, Proc. 35th IEEE PVSC, Honolulu, pp. 002694–
002697, 2010.
296

[Jor11] D.C. Jordan and S.R. Kurtz, “Photovoltaic degradation rates – an analytical
review”, Progress in Photovoltaics: Research and Applications, DOI:
10.1002/pip.1182, 18 pp., 2011.

[Kal09] S.A. Kalogirou, “Solar energy engineering: Processes and systems”, Elsevier,
Inc., pp. 476–483, 2009.

[Kam98] A. Kaminski, J.J. Marchand, and A. Laugier, “Non ideal dark I-V curves
behavior of silicon solar cells”, Solar Energy Materials and Solar Cells 51, pp.
221–231, 1998.

[Kan01] KanEnergi AS, “Nye fornybare energikilder” (in Norwegian), F. Salvesen, Ed.,
ISBN 82-12-01621-8, p. 7, 2001.

[Kas89] F. Kasten and A.T. Young, “Revised optical air mass tables and approximation
formula”, Applied Optics 28, pp. 4735–4738, 1989.

[Ken06] R.P. Kenny, M. Nikolaeva-Dimitrova, and E.D. Dunlop, “Performance


measurements of CIS modules: Outdoor and pulsed simulator comparison for
power and energy rating”, Proc. 4th IEEE WCPEC, Waikoloa, pp. 2058–2061,
2006.

[Ker03] M.J. Kerr, A. Cuevas, and P. Campbell, “Limiting efficiency of crystalline


silicon solar cells due to Coulomb‐enhanced Auger recombination”, Progress
in Photovoltaics: Research and Applications 11, pp. 97–104, 2003.

[Ker06] E.V. Kerschaver and G. Beaucarne, “Back-contact solar cells: a review”,


Progress in Photovoltaics: Research and Applications 14, pp. 107–123, 2006.

[Kim10] W. Kim and W. Choi, “A novel parameter extraction method for the one-diode
solar cell model”, Solar Energy 84, pp. 1008–1019, 2010.

[Kin96] D.L. King and P.E. Eckert, “Characterizing (rating) the performance of large
photovoltaic arrays for all operating conditions”, Proc. 25th IEEE PVSC,
Washington, pp. 1385–1388, 1996.

[Kin97a] D.L. King, J.A. Kratochvil, and W.E. Boyson, “Temperature coefficients for
PV modules and arrays: Measurement methods, difficulties and results”, Proc.
26th IEEE PVSC, Anaheim, pp. 1183–1186, 1997.

[Kin97b] D.L. King and D.R. Myers, “Silicon-photodiode pyranometers: operational


characteristics, historical experiences, and new calibration procedures”, Proc.
26th IEEE PVSC, Anaheim, pp. 1285–1288, 1997.
297

[Kin97c] D.L. King, J.A. Kratochvil, and W.E. Boyson, “Measuring solar spectral and
angle-of-incidence effects on photovoltaic modules and solar irradiance
sensors”, Proc. 26th IEEE PVSC, Anaheim, pp. 1113–1116, 1997.

[Kin98a] D.L. King, W.E. Boyson, and B.R. Hansen, “Improved accuracy for low-cost
solar irradiance sensors”, Proc. 2nd WCEPSEC, Vienna, pp. 2001–2004, 1998.

[Kin98b] D.L. King, J.A. Kratochvil, and W.E. Boyson, “Field experience with a new
performance characterization procedure for photovoltaic arrays”, Proc. 2nd
WCEPSEC, Vienna, pp. 1947–1952, 1998.

[Kin00] D.L. King, J.A. Kratochvil, and W.E. Boyson, “Stabilization and performance
characteristics of commercial amorphous-silicon PV modules”, Proc. 28th
IEEE PVSC, Anchorage, pp. 1446–1449, 2000.

[Kin04] D.L. King, W.E. Boyson, and J.A. Kratochvil, “Photovoltaic array performance
model”, SAND2004-3535, Sandia National Laboratories Report, 2004.

[Kip10] Kipp and Zonen, calibration certificate for CMP 3 dated 01/09/2010.

[Kip12a] “Kipp and Zonen – Pyranometers”,


http://www.kippzonen.com/?productgroup/331/Pyranometers.aspx, last
accessed 28/02/2012.

[Kip12b] “Kipp and Zonen Solar Energy Guide”,


http://www.kippzonen.com/?download/415212/Solar+Energy+Guide.aspx, last
accessed 28/02/2012.

[Kip12c] “CMP 3”, http://www.kippzonen.com/?product/1131/CMP+3.aspx, last


accessed 01/03/2012.

[Kip12d] “AMPBOX”, http://www.kippzonen.com/?Product/37172/AMPBOX.aspx,


last accessed 01/03/2012.

[Kip12e] AMPBOX Brochure,


http://www.kippzonen.com/?download/1131/AMPBOX+Signal+Amplifier+-
+Brochure.aspx, last accessed 01/03/2012.

[Kle09] Ø. Kleven, H. Persson, C. Good, W. Sulkowski, and T. Boström, “Solar cells


above the Arctic Circle – a comparison between a two-axis tracking system and
simulations”, Proc. 24th EUPVSEC, Hamburg, pp. 4090–4093, 2009.

[Kok11] K. Koka, “Photovoltaic module performance and thermal characterizations:


Data collection and automation of data processing”, MSc thesis, Arizona State
University, 2011.
298

[Kra09] S. Krauter and A. Preiss, “Comparison of module temperature measurement


methods”, Proc. 34th IEEE PVSC, Philadelphia, pp. 000333–000338, 2009.

[Kui11] J. Kuitche, J. Oh, A. Brunger, T. Inoue, M. Muller, C. Bauerdick, J. Althaus, S.


Kiehn, V. Feng, U. Therhaag, and R. Struwe, “One year NOCT round-robin
testing per IEC 61215 standard”, Proc. 37th IEEE PVSC, Seattle, pp. 002380–
002385, 2011.

[Kun04] G. Kunz and A. Wagner, “Internal series resistance determination of only one
I-V curve under illumination”, Proc. 19th EUPVSEC, vol. 2, Paris, pp. 2671–
2674, 2004.

[Kuu08] J. Kuune, A. Tolvanen, and J. Hyvärinen, “Sweep time, spectral mismatch and
light soaking in thin film module measurements”, Proc. 23rd EUPVSEC,
Valencia, pp. 2855–2857, 2008.

[Lad10] R. Lad, J. Wohlgemuth, and G. Tamizhmani, “Outdoor energy ratings and


spectral effects of photovoltaic modules”, Proc. 35th IEEE PVSC, Honolulu,
pp. 002827– 002832, 2010.

[Lea05] J. Lean, “Living with a variable sun”, Physics Today, pp. 32–38, 06/2005.

[Lee10] M.-K. Lee, J.-C. Wang, S.-F. Horng, and H.-F. Meng, “Extraction of solar cell
series resistance without presumed current–voltage functional form”, Solar
Energy Materials and Solar Cells 94, pp. 578–582, 2010.

[Luo12] J. Luoma, J. Kleissl, and K. Murray, “Optimal inverter sizing considering cloud
enhancement”, Solar Energy 86, pp. 421–429, 2012.

[Luq03a] A. Luque and S. Hegedus, editors. “Handbook of photovoltaic science and


engineering”, Wiley, ISBN 0-471-49196-9, Section 20.10 “PV generator
behaviour under real operation conditions”, pp. 947–956, 2003.

[Luq03b] A. Luque and S. Hegedus, editors. “Handbook of photovoltaic science and


engineering”, Wiley, ISBN 0-471-49196-9, Chapter 3 “The physics of the solar
cell”, pp. 61–112, 2003.

[Liu60] B.Y.H. Liu and R.C. Jordan, “The interrelationship and characteristic
distribution of direct, diffuse and total solar radiation”, Solar Energy 4, pp. 1–
19, 1960.

[Lof93] J.J. Loferski, “The first forty years: A brief history of the modern photovoltaic
age”, Progress in Photovoltaics: Research and Applications 1, pp. 67–78,
1993.
299

[Lub11] W.D. Lubitz, “Effect of manual tilt adjustments on incident irradiance on fixed
and tracking solar panels”, Applied Energy 88, pp. 1710–1719, 2011.

[Mac00] D. Macdonald and A. Cuevas, “Reduced fill factors in multicrystalline silicon


solar cells due to injection-level dependent bulk recombination lifetimes”,
Progress in Photovoltaics: Research and Applications 8, pp. 363–375, 2000.

[Mac01] D.H. Macdonald, “Recombination and trapping in multicrystalline silicon solar


cells”, PhD Thesis, the Australian National University, Canberra, 2001.

[Map12] “Jon Lilletuns vei 9, Grimstad, Norway”,


http://maps.google.com/maps?q=Jon+Lilletuns+vei+9,+Grimstad,+Norway&hl
=en&ll=58.313532,8.568592&spn=0.04567,0.132093&sll=58.334311,8.57718
4&sspn=0.002853,0.008256, last accessed 23/02/2012.

[Mar01] N. Martin and J.M. Ruiz, “Calculation of the PV modules angular losses under
field conditions by means of an analytical model”, Solar Energy Materials and
Solar Cells 70, pp. 25–38, 2001.

[Mas04] N. Mason, A. Artigao, P. Banda, R. Bueno, J.M. Fernandez, C. Morilla, and R.


Russell, The technology and performance of the latest generation buried
contact solar cell manufactured in BP Solar's Tres Cantos facility”, Proc. 19th
EUPVSEC, vol. 2, Paris, pp. 2653–2655, 2004.

[Mat12] MathWorks, “Smoothing Splines: Interpolation and Smoothing (Curve Fitting


Toolbox™)”, http://www.mathworks.se/help/toolbox/curvefit/bsz6bn_-1.html,
last accessed 19/02/2012.

[McA04] L.J.B. McArthur, “Baseline Surface Radiation Network (BSRN) – Operations


Manual, Version 2.1”,
http://www.bsrn.awi.de/fileadmin/user_upload/Home/Publications/McArthur.p
df, 2004.

[McG08] B. McGuire, “Seven years to save the Planet: The questions and the answers”,
Weidenfeld & Nicolson, 2008.

[McI99] K.R. McIntosh and C.B. Honsberg, “A new technique for characterizing
floating-junction-passivated solar cells from their dark IV curves”, Progress in
Photovoltaics: Research and Applications 7, pp. 363–378, 1999.

[McI00] K.R McIntosh, P.P. Altermatt, and G. Heiser, “Depletion-region recombination


in silicon solar cells: When does mDR = 2?”, Proc. 16th EUPVSEC, vol. 1,
Glasgow, pp. 251–254, 2000.
300

[McI01] K.R. McIntosh, “Lumps, humps and bumps: Three detrimental effects in the
current-voltage curve of silicon solar cells”, PhD thesis, University of New
South Wales, Sydney, 2001.

[McM96] T.J. McMahon, T.S. Basso, and S.R. Rummel, “Cell shunt resistance and
photovoltaic module performance”, Proc. 25th IEEE PVSC, Washington, pp.
1291–1294, 1996.

[Mer98] J. Merten, J.M. Asensi, C. Voz, A.V. Shah, R. Platz, and J. Andreu, “Improved
equivalent circuit and analytical model for amorphous silicon solar cells and
modules”, IEEE Transactions on Electron Devices 45, pp. 423–429, 1998.

[Mer10] A. Mermoud and T. Lejeune, “Performance assessment of a simulation model


for PV modules of any available technology”, Proc. 25th EUPVSEC, Valencia,
pp. 4786–4791, 2010.

[Met07a] A. Mette, D. Pysch, G. Emanuel, D. Erath, R. Preu, and S. W. Glunz, “Series


resistance characterization of industrial silicon solar cells with screen-printed
contacts using hotmelt paste”, Progress in Photovoltaics: Research and
Applications 15, pp. 493–505, 2007.

[Met07b] A. Mette, “New concepts for front side metallization of industrial silicon solar
cells”, PhD thesis, Fraunhofer Institute for Solar Energy Systems, Freiburg,
2007.

[Mey04] E.L. Meyer and E. Ernest van Dyk, “Assessing the reliability and degradation
of photovoltaic module performance parameters”, IEEE Transactions on
Reliability 53, pp. 83–92, 2004.

[Mia83] P. Mialhe and J. Charette, “Experimental analysis of I–V characteristics of


solar cells”, American Journal of Physics 51, pp. 68–70, 1983.

[Mia84] P. Mialhe, A. Khoury, and J.P. Charles, “A review of techniques to determine


the series resistance of solar cells”, Physica Status Solidi A 83, pp. 403–409,
1984.

[Mic88] J.J. Michalsky, “The Astronomical Almanac’s algorithm for approximate solar
position (1950–2050)”, Solar Energy 40, pp. 227–235, 1988.

[Mid07] O.-M. Midtgård, “A simple photovoltaic simulator for testing of power


electronics”, Proc. European Conference on Power Electronics and
Applications, Aalborg, DOI: 10.1109/EPE.2007.4417450, 10 pp., 2007.
301

[Mid08] O.M. Midtgård and T.O. Saetre, “A comparative simulation study of two-
terminal multijunction cells versus connected individual cells in a spectrum
splitting scheme”, Proc. 23rd EUPVSEC, Valencia, pp. 253–256, 2008.

[Mid10] O.-M. Midtgard, T.O. Sætre, G. Yordanov, A.G. Imenes, and C.L. Nge, “A
qualitative examination of performance and energy yield of photovoltaic
modules in Southern Norway”, Renewable Energy 35, pp. 1266–1274, 2010.

[Min10] P. Mints, “The commercialization of thin film technologies: Past, present and
future”, Proc. 35th IEEE PVSC, Honolulu, pp. 002400–002404, 2010.

[Mit10] L. Mitchell, J. Seidel, R. Ocampo, S. Wilson, and J. Andrews, “Impact of


microclimates on solar resource: Case study of the solar resource in San
Francisco”, Proc. 35th IEEE PVSC, Honolulu, pp. 001091–001094, 2010.

[Moh03a] N. Mohan, T.M. Undeland, and W.P. Robbins, “Power Electronics –


Converters, Applications, and Design”, 3rd Ed., Wiley, pp. 733–737, 2003.

[Moh03b] N. Mohan, T.M. Undeland, and W.P. Robbins, “Power Electronics –


Converters, Applications, and Design”, 3rd Ed., Wiley, pp. 172–173, 2003.

[Moh03c] N. Mohan, T.M. Undeland, and W.P. Robbins, “Power Electronics –


Converters, Applications, and Design”, 3rd Ed., Wiley, pp. 531–534, 543, 626–
631, 2003.

[Moh10] H.-D. Mohring, D. Stellbogen, A. Jagomägi, E. Mõttus, T. Betts, R.


Gottschalg, T. Zdanowicz, M. Prorok, G. Friesen, D. Dominé, A.G. de
Montgareuil, F. Fabero, D. Faiman, and W. Herrmann, “Energy delivery of PV
devices – implementation of best practices for outdoor characterization and
testing”, Proc. 25th EUPVSEC, Valencia, pp. 4322–4325, 2010.

[Mul12] M. Muller, B. Marion, and J. Rodriguez, “Evaluating the IEC 61215 Ed.3
NMOT procedure against the existing NOCT procedure with PV modules in a
side-by-side configuration”, Proc. 38th IEEE PVSC, Austin, pp. 000697–
000702, 2012.

[Mun11] M.A. Munoz, F. Chenlo, and M.C. Alonso-García, “Influence of initial power
stabilization over crystalline-Si photovoltaic modules maximum power”,
Progress in Photovoltaics: Research and Applications 19, pp. 417–422, 2011.

[Mül04] H. Müllejans, J. Hyvärinen, J. Karila, and E.D. Dunlop, “Reliability of routine


2-diode model fitting of PV modules”, Proc. 19th EUPVSEC, vol. 2, Paris, pp.
2459–2462, 2004.
302

[Mül08] H. Müllejans, D. Pavanello, and A. Virtuani, “New correction procedure for I-


V curves measured under varying irradiance”, Proc. 23rd EUPVSEC, Valencia,
pp. 2818–2821, 2008.

[Mül09] H. Müllejans, W. Zaaiman, and R. Galleano, “Analysis and mitigation of


measurement uncertainties in the traceability chain for the calibration of
photovoltaic devices”, Measurement Science and Technology 20, Article ID:
075101, DOI: 10.1088/0957-0233/20/7/075101, 12 pp, 2009.

[Mye89] D.R. Myers, K.A. Emery, and T.L. Stoffel, “Uncertainty estimates for global
solar irradiance measurements used to evaluate PV device performance”, Solar
Cells 27, pp. 455–464, 1989.

[Mye02] D.R. Myers, T.L. Stoffel, I. Reda, S.M. Wilcox, and A.M. Andreas, “Recent
progress in reducing the uncertainty in and improving pyranometer
calibrations”, Journal of Solar Energy Engineering 124, pp. 44–50, 2002.

[Mye04] D.R. Myers, I. Reda, S.M. Wilcox, and A.M. Andreas, “Optical radiation
measurement for photovoltaic applications: instrumentation uncertainty and
performance”, http://www.nrel.gov/docs/gen/fy04/36321.pdf, last accessed
06/10/2011, 12 pp., 2004.

[Mye09a] D.R. Myers and S.M. Wilcox, “Relative accuracy of 1-minute and daily total
solar radiation data from 12 global and 4 direct beam solar radiometers”,
http://www.nrel.gov/docs/fy09osti/45374.pdf, last accessed 06/10/2011, 8 pp.,
2009.

[Mye09b] D.R. Myers, “Comparison of historical satellite-based estimates of solar


radiation resources with recent rotating shadowband radiometer
measurements”, http://www.nrel.gov/docs/fy09osti/45375.pdf, last accessed
24/02/2012, 11 pp., 2009.

[Nan91] S. Nann and C. Riordan, “Solar spectral irradiance under clear and cloudy
skies: Measurements and a semiempirical model”, Journal of Applied
Meteorology 30, pp. 447–462, 1991.

[Nas83] P.-M. Nast, “Measurements on the accuracy of pyranometers”, Solar Energy


31, pp. 279–282, 1983.

[NCR11] National Center for Renewable Energy (Energiparken),


http://energiparken.hia.no/, last accessed 30/10/2011.

[Nel03a] J. Nelson, “The physics of solar cells”, Imperial College Press, London, p. 19,
2003.
303

[Nel03b] J. Nelson, “The physics of solar cells”, Imperial College Press, London, p.
151, 2003.

[Neu07] D.-H. Neuhaus and A. Münzer, “Industrial silicon wafer solar cells”, Advances
in OptoElectronics, Article ID: 24521, DOI:10.1155/2007/24521, 15 pp., 2007.

[Nge09] C.L. Nge, G. Yordanov, O.-M. Midtgård, T.O. Sætre, and L. Norum, “A
comparative simulation analysis of maximum power point tracking
approaches”, Proc. 24th EUPVSEC, Hamburg, pp. 3668–3672, 2009.

[Nge10] C.L. Nge, O.-M. Midtgård, G. Yordanov, L. Norum, and T. O. Sætre, “A new
maximum power point tracking approach for partial shading conditions”, Proc.
25th EUPVSEC, Valencia, pp. 4476–4478, 2010.

[Nik10] S. Niki, M. Contreras, I. Repins, M. Powalla, K. Kushiya, S. Ishizuka, and K.


Matsubara, “CIGS absorbers and processes”, Progress in Photovoltaics:
Research and Applications 18, pp. 453–466, 2010.

[NRE12a] “2000 ASTM Standard Extraterrestrial Spectrum Reference E-490-00”,


http://rredc.nrel.gov/solar/spectra/am0/E490_00a_AM0.xls, last accessed
22/02/2012.

[NRE12b] NREL System Advisor Model (SAM) version 2.5.0.2 installer for Windows
XP/Vista/7, https://sam.nrel.gov/files/content/binaries/sam_install_2502.exe,
last accessed 20/05/2012.

[NRE12c] NREL System Advisor Model (SAM) version 3.0.3.0 installer for Windows
XP/Vista/7, https://sam.nrel.gov/files/content/binaries/sam_install_3.0.3.0.exe,
last accessed 20/05/2012.

[NRE12d] NREL System Advisor Model (SAM), https://sam.nrel.gov/, last accessed


19/06/2012.

[NVE07] NVE, Enova SF, the Norwegian Research Council, and Innovasjon Norge,
“Fornybar energi 2007” (in Norwegian), ISBN: 978-82-410-0632-6, p. 20,
2007.

[Ols86] J.A. Olseth and A. Skartveit, “The solar radiation climate of Norway”, Solar
Energy 37, pp. 423–428, 1986.

[Oue99] Z. Ouennoughi and M. Chegaar, “A simpler method for extracting solar cell
parameters using the conductance method”, Solid-state Electronics 43, pp.
1985–1988, 1999.
304

[Par04a] A.V. Parisi and N. Downs, “Variation of the enhanced biologically damaging
solar UV due to clouds”, Photochemical and Photobiological Sciences 3, pp.
643–647, 2004.

[Par04b] A.V. Parisi and N. Downs, “Cloud cover and horizontal plane eye damaging
solar UV exposures”, International Journal of Biometeorology 49, pp. 130–
136, 2004.

[Per93] R. Perez, R. Seals, and J. Michalsky, “All-weather model for sky luminance
distribution – preliminary configuration and validation”, Solar Energy 50, pp.
235–245, 1993.

[Per04] M. Perić, Z. Matić, and K. Trontl, “Valorisation of spectral influence and


reflection on photovoltaic modules – contribution to optimization of modules
position”, Proc. 19th EUPVSEC, vol. 2, Paris, pp. 2599–2602, 2004.

[Per07] J.J. Pérez-López, F. Fabero, and F. Chenlo, “Experimental solar spectral


irradiance until 2500 nm: Results and influence on the PV conversion of
different materials”, Progress in Photovoltaics: Research and Applications 15,
pp. 303–315, 2007.

[Phi05] E.M. Phillips and D.S. Pugh, “How to get a PhD”, 4th Ed., Open University
Press, Berkshire, 2005.

[Pho11a] “PHOTON’s Yield Measurement”,


http://www.photon.info/AxCMSwebLive_PremiumSample/photon_lab_modul
_ertragsm_en.photon, last accessed 19/12/2011.

[Pho11b] Photon International, “Cost parity is not enough”, interview with Dick
Swanson (founder of SunPower Corp.), pp. 30–31, 08/2011.

[Pho11c] Photon International, “Pasan introduces its HighLIGHT module tester”, p.


152, 08/2011.

[Pho11d] Photon International, “PHOTON Lab’s outdoor module tests – August


results”, pp. 108–111, 10/2011.

[Pho12a] A. Rosenberger and P. Welter, “No two modules alike”, Photon International,
pp. 92–97, 02/2012.

[Pho12b] A. Rosenberger, P. Welter, and C. Podewils, “Sharp curves”, Photon


International, pp. 98–102, 02/2012.

[Pho12c] Photon International, “Detailed results of Photon’s 2011 yield


measurements”, pp. 104–124, 02/2012.
305

[Pin10] S. Pingel, O. Frank, M. Winkler, S. Daryan, T. Geipel, H. Hoehne, and J.


Berghold, “Potential induced degradation of solar cells and panels”, Proc. 35th
IEEE PVSC, Honolulu, pp. 002817–002822, 2010.

[Pol11a] D. Polverini, G. Tzamalis, and H. Müllejans, “A validation study on series


resistance calculation according to IEC60891, ED 2.0”, Proc. 26th EUPVSEC,
Hamburg, pp. 3417–3420, 2011.

[Pol11b] D. Polverini, G. Tzamalis, and H. Müllejans, “A validation study of


photovoltaic module series resistance determination under various operating
conditions according to IEC 60891”, Progress in Photovoltaics: Research and
Applications, DOI: 10.1002/pip.1200, 11 pp., 2011.

[Pra10] L. Pratt and D.L. King, “The effect of uncertainty in modeling coefficients
used to predict energy production using the Sandia Array Performance Model”,
Proc. 35th IEEE PVSC, Honolulu, pp. 002718–002723, 2010.

[PRE11] PR Electronics Sales Department, private communication dated 08/02/2011.

[PRE12] PR Electronics 2204 isolation amplifier data sheet,


http://www.prelectronics.co.uk/prefiles/2204/Datablad/2204uk.pdf, last
accessed 01/03/2012.

[Pro08] M. Prorok, W. Kołodenny, T. Zdanowicz, R. Gottschalg, and D. Stellbogen,


“Reducing uncertainty of PV module temperature determination based on
analysis using data gained during outdoor monitoring”, Proc. 23rd EUPVSEC,
Valencia, pp. 2865–2871, 2008.

[PVGIS] http://re.jrc.ec.europa.eu/pvgis/ last accessed 10/11/2012.

[Pys07] D. Pysch, A. Mette, and S.W. Glunz, “A review and comparison of different
methods to determine the series resistance of solar cells”, Solar Energy
Materials and Solar Cells 91, pp. 1698–1706, 2007.

[Q-C11] Q-Cells, “Multicrystalline solar cell Q6LTT3 – G2”, http://www.q-


cells.com/uploads/tx_abdownloads/files/Q-Cells_Q6LTT3-
G2_Data_sheets_2011-03_Rev01_EN.pdf, last accessed 14/10/2011.

[Que09] H.J. Queisser, “Detailed balance limit for solar cell efficiency”, Materials
Science and Engineering B 159–160, pp. 322–328, 2009.

[REC12] REC AE Series Data Sheet, http://www.posharp.com/Businesses/742d3860-


481f-4339-a774-6b4489d22cf8/Panel/REC_AE_Series_ENG.pdf, last accessed
21/06/2012.
306

[Rei12] F. Reil, A. Sepanski, W. Herrmann, J. Althaus, W. Vaaßen, and H. Schmidt,


“Qualification of arcing risks in PV modules”, Proc. 38th IEEE PVSC, Austin,
pp. 000727–000730, 2012.

[Rob94a] S.J. Robinson, A.G. Aberle, and M.A. Green, “Departures from the principle
of superposition in silicon solar cells”, Journal of Applied Physics 76, pp.
7920–7930, 1994.

[Rob94b] S.J. Robinson, G.F. Zheng, W. Zhang, Z. Shi, M.A. Green, and R. Bergmann,
“Opto-electronic characterization of thin-film crystalline silicon solar cells
grown from metal solutions”, Proc. 12th EUPVSEC, vol. 2, Amsterdam, pp.
1831–1834, 1994.

[Rob94c] S.J. Robinson, A.G. Aberle, and M.A. Green, “Recombination saturation
effects in silicon solar cells”, IEEE Transactions on Electron Devices 41, pp.
1556–1569, 1994.

[Roy08] J. Roy, T.R. Betts, R. Gottschalg, S. Mau, S. Zamini, R.P. Kenny, H.


Müllejans, G. Friesen, S. Dittmann, H.G. Beyer, and A. Jagomägi, “Validation
of proposed photovoltaic energy rating standard and sensitivity to
environmental parameters”, Proc. 23rd EUPVSEC, Valencia, pp. 2728–2734,
2008.

[Rum06] S. Rummel, A. Anderberg, K. Emery, D. King, G. TamizhMani, T. Arends, G.


Atmaram, L. Demetrius, W. Zaaiman, N. Cereghetti, W. Herrmann, W. Warta,
F. Neuberger, K. Morita, and Y. Hishikawa, “Results from the second
international module intercomparison”, Proc. 4th IEEE WCPEC, Waikoloa, pp.
2034–2037, 2006.

[Sae11] T.O. Saetre, O.-M. Midtgård, and G.H. Yordanov, “A new analytical solar cell
I-V curve model”, Renewable Energy 36, pp. 2171–2176, 2011.

[Sae12] T.O. Saetre (University of Agder), private communication dated 13/05/2012.

[Sak03] S. Sakamoto and T. Oshiro, “Field test results on the stability of crystalline
silicon photovoltaic modules manufactured in the 1990’s”, Proc. 3rd IEEE
WCPEC, Osaka, pp. 1888–1891, 2003.

[San10] L. Sandrolini, M. Artioli, and U. Reggiani, “Numerical method for the


extraction of photovoltaic module double-diode model parameters through
cluster analysis”, Applied Energy 87, pp. 442–451, 2010.

[San11] SANYO Component Europe GmbH, “Sanyo Solar Photovoltaics-Europe: New


products”, http://www.sanyo-solar.eu/en/products/new-products/, last accessed
21/12/2011.
307

[Sas93] R.A. Sasala and J.R. Sites, “Time dependent voltage in CuInSe2 and CdTe solar
cells”, Proc. 23rd IEEE PVSC, Louisville, pp. 543–548, 1993.

[Sat12] “Satel-Light – The European Database of Daylight and Solar Radiation”,


http://www.satel-light.com/index2.htm, last accessed 24/02/2012.

[Sau12] K.J. Sauer and T. Roessler, “Systematic approaches to ensure correct


representation of measured multi-irradiance module performance in PV system
energy production forecasting software programs”, Proc. 38th IEEE PVSC,
Austin, pp. 000703–000709, 2012.

[Sch01] J. Schmidt, A. Cuevas, S. Rein, and S.W. Glunz, “Impact of light-induced


recombination centres on the current–voltage characteristic of Czochralski
silicon solar cells”, Progress in Photovoltaics: Research and Applications 9,
pp. 249–255, 2001.

[Sch12] Citations of [Gow99] according to the Google Scholar indexing engine,


http://scholar.google.no/scholar?cites=7199735306859938690&as_sdt=2005&
sciodt=0,5&hl=en, last accessed 08/07/2012.

[Ser07] D. Sera, R. Teodorescu, and P. Rodriguez, “PV panel model based on datasheet
values”, Proc. 2007 IEEE ISIE, pp. 2392–2396, 2007.

[Ser10a] D. Sera, “Series resistance monitoring for photovoltaic modules in the vicinity
of MPP”, Proc. 25th EUPVSEC, Valencia, pp. 4506–4510, 2010.

[Ser10b] D. Sera (University of Aalborg), private communication dated 11/2010.

[Ses10] C. Seshan, “CPV: Not just for hot deserts!”, Proc. 35th IEEE PVSC, Honolulu,
pp. 003075–003080, 2010.

[Sha04] A.V. Shah, H. Schade, M. Vanecek, J. Meier, E. Vallat-Sauvain, N. Wyrsch, U.


Kroll, C. Droz, and J. Bailat, “Thin-film silicon solar cell technology”,
Progress in Photovoltaics: Research and Applications 12, pp. 113–142, 2004.

[Sho49] W. Shockley, “The theory of p-n junctions in semiconductors and p-n junction
transistors”, The Bell System Technical Journal 28, pp. 435–489, 1949.

[Sho61] W. Shockley and H.J. Queisser, “Detailed balance limit of efficiency of p-n
junction solar cells”, Journal of Applied Physics 32, pp. 510–519, 1961.

[Sie10] S. Siebentritt, M. Igalson, C. Persson, and S. Lany, “The electronic structure of


chalcopyrites – bands, point defects and grain boundaries”, Progress in
Photovoltaics: Research and Applications 18, pp. 390–410, 2010.
308

[Sin09] W.C. Sinke, W.v. Hooff, G. Coletti, B. Ehlen, G. Hahn, S. Reber, J. John, G.
Beaucarne, E.v. Kerschaver, M. de Wild-Scholten, and A. Metz, “Wafer-based
crystalline silicon modules at 1 €/Wp: Final results from the
CRYSTALCLEAR integrated project”, Proc. 24th EUPVSEC, Hamburg, pp.
845–856, 2009.

[Sit89] J.R. Sites and P.H. Mauk, “Diode quality factor determination for thin-film
solar cells”, Solar Cells 27, pp. 411–417, 1989.

[Ska86] A. Skartveit and J.A. Olseth, “Modelling slope irradiance at high latitudes”,
Solar Energy 36, pp. 333–344, 1986.

[Sma85] D.J. Smalley, “A cloud shading direct solar radiation model”, MSc thesis,
Texas Tech University, 1985.

[Sol10] SolData Instruments, Calibration protocol for SolData 80spc dated 04/10/2010.

[Sol12] “SolData pyranometer 80spc user’s guide”, http://www.soldata.dk/PDF-


NYE/80spc-pyranometer.pdf, last accessed 29/02/2012.

[Ste11] S. Steingrube, H. Wagner, H. Hannebauer, S. Gatz, R. Chen, S.T. Dunham, T.


Dullweber, P.P. Altermatt, and R. Brendel, “Loss analysis and improvements
of industrially fabricated Cz-Si solar cells by means of process and device
simulations”, Energy Procedia 8, pp. 263–268, 2011.

[Str00] B.G. Streetman and S. Banerjee, “Solid State Electronic Devices”, 5th Ed.,
Prentice Hall, pp. 211–220, 2000.

[Sue88] H. Suehrcke and P.G. McCormick, "The frequency distribution of


instantaneous insolation values", Solar Energy 40, pp. 413–422, 1988.

[Sue90] H. Suehrcke, C.P. Ling, and P.G. McCormic, “The dynamic response of
instruments measuring instantaneous solar radiation”, Solar Energy 44, pp.
145–148, 1990.

[Sun11] SunPower Corporation, “SunPower E20 Series Solar Panels”,


http://www.sunpowercorp.co.uk/commercial/products-services/solar-
panels/e20/, last accessed 20/12/2011.

[Sun12] SunPower Corporation, “PVSim and solar energy system performance


modeling”,
http://us.sunpowercorp.com/cs/BlobServer?blobkey=id&blobwhere=13002585
33987&blobheadername2=Content-Disposition&blobheadername1=Content-
Type&blobheadervalue2=inline%3B+filename%3DPVSim%2Band%2BSolar
%2BEnergy%2BSystem%2BPerformance%2BModeling.pdf&blobheadervalue
309

1=application%2Fpdf&blobcol=urldata&blobtable=MungoBlobs, last accessed


19/06/2012.

[Sur07] M. Šúri, T.A. Huld, E.D. Dunlop, and H.A. Ossenbrink, “Potential of solar
electricity generation in the European Union member states and candidate
countries”, Solar Energy 81, pp. 1295–1305, 2007.

[Sur08a] M. Šúri, T. Huld, T. Cebecauer, and E.D. Dunlop, “Geographic aspects of


photovoltaics in Europe: Contribution of the PVGIS website”, IEEE Journal of
Selected Topics in Applied Earth Observations and Remote Sensing 1, pp. 34–
41, 2008.

[Sur08b] M. Šúri, J. Remund, T. Cebecauer, D. Dumortier, L. Wald, T. Huld, and P.


Blanc, “First steps in the cross-comparison of solar resource spatial products in
Europe”, Proc. EUROSUN 2008, 9 pp., 2008.

[Sze07a] S.M. Sze and K.K. Ng, “p-n Junctions”, in: Physics of Semiconductor Devices,
3rd Ed., Wiley, pp. 79–133, 2007.

[Sze07b] S.M. Sze and K.K. Ng, “Solar Cells”, in: Physics of Semiconductor Devices,
3rd Ed., Wiley, pp. 719–742, 2007.

[The10] D. Thevenard, A. Driesse, S. Pelland, D. Turcotte, and Y. Poissant,


“Uncertainty in long-term photovoltaic yield predictions”, Report #2010-122
(RP-TEC), CanmetENERGY, Varennes Research Center, National Resources
Canada, 52 pp., 2010.

[Tol12] F.J. Toledo, J.M. Blanes, A. Garrigós, and J.A. Martínez, “Analytical
resolution of the electrical four-parameters model of a photovoltaic module
using small perturbation around the operating point”, Renewable Energy 43,
pp. 83–89, 2012.

[Tre98] F. Treble, “Milestones in the development of crystalline silicon solar cells”,


Renewable Energy 15, pp. 473–478, 1998.

[Tru07] T. Trupke, E. Pink, R.A. Bardos, and M.D. Abbott, “Spatially resolved series
resistance of silicon solar cells obtained from luminescence imaging”, Applied
Physics Letters 90, Article ID: 093506, 3 pp., 2007.

[Tsu08] J. Tsutsui and K. Kurokawa, “Investigation to estimate the short circuit current
by applying the solar spectrum”, Progress in Photovoltaics: Research and
Applications 16, pp. 205–211, 2008.
310

[Tsu09] Y. Tsuno, Y. Hishikawa, and K. Kurokawa, “Modeling of the I-V curves of the
PV modules using linear interpolation/extrapolation”, Solar Energy Materials
and Solar Cells 93, pp. 1070–1073, 2009.

[Tul76] S.E. Tuller, “The relationship between diffuse, total and extra terrestrial solar
radiation”, Solar Energy 18, pp. 259–263, 1976.

[Tur06] D.J. Turnbull, A.V. Parisi, and N.J. Downs, “Effect of clouds on the diffuse
component of the solar terrestrial erythemal UV”, Radiation Protection in
Australasia 23, pp. 2–9, 2006.

[UOr11] University of Oregon, Solar Radiation Monitoring Laboratory, “Sun path chart
program”, created by Peter Harlan and Frank Vignola,
http://solardat.uoregon.edu/SunChartProgram.html, last accessed 28/10/2011.

[USD11] U.S. Department of Commerce, National Oceanic and Atmospheric


Administration, “Trends in Atmospheric Carbon Dioxide”,
http://www.esrl.noaa.gov/gmd/ccgg/trends/, last accessed 25/10/2011.

[Var00] E. Vartiainen, “A new approach to estimating the diffuse irradiance on inclined


surfaces”, Renewable Energy 20, pp. 45–64, 2000.

[Vei95] N. Veissid, D. Bonnet, and H. Richter, “Experimental investigation of the


double exponential model of a solar cell under illuminated conditions:
Considering the instrumental uncertainties in the current, voltage and
temperature values”, Solid-state Electronics 38, pp. 1937–1943, 1995.

[Ver12] D. Verma, M. Tayyib, T.O. Saetre, and O.-M. Midtgård, “Outdoor performance
of north-facing multicrystalline modules in Southern Norway”, Proc. 38th IEEE
PVSC, Austin, pp. 002372–002375, 2012.

[Vir08] A. Virtuani, H. Muellejans, F. Ponti, and E. Dunlop, “Comparison of indoor and


outdoor performance measurements of recent commercially available solar
modules”, Proc. 23rd EUPVSEC, Valencia, pp. 2713–2718, 2008.

[Vir12] A. Virtuani, G. Rigamonti, P. Beljean, G. Friesen, M. Pravettoni, and D.


Chianese, “A fast and accurate method for the performance testing of high-
efficiency c-Si photovoltaic modules using a 10 ms single-pulse solar
simulator”, Proc. 38th IEEE PVSC, Austin, pp. 000496–000500, 2012.

[Vod04] C. Vodermayer, M. Zehner, O. Grasnick, and G. Becker, “A new approach for


outdoor weather, PV-module or PV-generator data acquisition”, Proc. 19th
EUPVSEC, vol. 2, Paris, pp. 2418–2421, 2004.
311

[Wal01] G. Walker, “Evaluating MPPT converter topologies using a MATLAB PV


model,” Journal of Electrical and Electronics Engineers, Australia 21, pp. 49–
56, 2001.

[Wan90] A. Wang, J. Zhao, and M.A. Green, “24% efficient silicon solar cells”,
Applied Physics Letters 57, pp. 602–604, 1990.

[Wei09] D. Weinstock and J. Appelbaum, “Optimization of solar photovoltaic fields”,


Journal of Solar Energy Engineering 131, 9 pp., 2009.

[Wen01] G. Wen, R.F. Calahan, S.-C. Tsay, and L. Oreopoulos, “Impact of cumulus
cloud spacing on Landsat atmospheric correction and aerosol retrieval”,
Journal of Geophysical Research 106, pp. 12129–12138, 2001.

[Wer88] J.H. Werner, “Schottky barrier and pn-junction I/V plots – small signal
evaluation”, Applied Physics A 47, pp. 291–300, 1988.

[Wer08] B. Werner, W. Kolodenny, A. Dziedzic, and T. Zdanovicz, “Experimental


determination of physical parameters in II-(III)-VI thin-film solar modules”,
Proc. 23rd EUPVSEC, Valencia, pp. 2881–2884, 2008.

[Whi01] K. Whitfield and C.R. Osterwald, “Procedure for determining the uncertainty
of photovoltaic module outdoor electrical performance”, Progress in
Photovoltaics: Research and Applications 9, pp. 87–102, 2001.

[Wil04a] S.R. Williams, T.R. Betts, R. Gottchalg, and D.G. Infield, “Modelling real
annual PV module performance with consideration to spectral and incidence
angle effects”, Proc. 19th EUPVSEC, vol. 2, Paris, pp. 2645–2647, 2004.

[Wil04b] S.R. Williams, R. Gottschalg, and D.G. Infield, “PV modules real operating
performance in the UK, a temperate maritime climate”, Proc. 19th EUPVSEC,
vol. 2, Paris, pp. 2109–2112, 2004.

[Wil06] S.R. Williams, T.R. Betts, R. Gottschalg, D.G. Infield, H. de Moor, N.v.d.
Borg, A.R. Burgers, G. Friesen, D. Chianese, A.G. de Montgareuil, T.
Zdanowicz, D. Stellbogen, and W. Herrmann, “Accuracy of energy prediction
methodologies”, Proc. 4th IEEE WCPEC, Waikoloa, pp. 2206–2209, 2006.

[WMO08] World Meteorological Organization (WMO), “Guide to meteorological


instruments and methods of observation (WMO-No. 8)”, 7th Ed., 2008,
ftp://ftp.wmo.int/Documents/MediaPublic/Publications/WMO8_CIMOguide/.

[Woh12] J.H. Wohlgemuth and S.R. Kurtz, “How can we make PV modules safer?”,
Proc. 38th IEEE PVSC, Austin, pp. 003162–003165, 2012.
312

[Wol63] M. Wolf and H. Rauschenbach, “Series resistance effects on solar cell


measurements”, Advanced Energy Conversion 3, pp. 455–479, 1963.

[Wol77] M. Wolf, G.T. Noel, and R.J. Stirn, “Investigation of the double exponential in
the current-voltage characteristics of silicon solar cells”, IEEE Transactions on
Electron Devices 24, pp. 419–428, 1977.

[Woy03] A. Woyte, J. Nijs, and R. Belmans, “Partial shadowing of photovoltaic arrays


with different system configurations: Literature review and field test results”,
Solar Energy 74, pp. 217–233, 2003.

[Yor10a] G.H. Yordanov, O.-M. Midtgård, and T.O. Saetre, “Extracting parameters
from semi-log plots of polycrystalline silicon PV modules outdoor I-V data:
Double-exponential model revisited”, Proc. 35th IEEE PVSC, Honolulu, pp.
002756–002761, 2010.

[Yor10b] G.H. Yordanov, O.-M. Midtgård, and T.O. Saetre, “Two-diode model
revisited: Parameters extraction from semi-log plots of I-V data”, Proc. 25th
EUPVSEC, Valencia, pp. 4156–4163, 2010.

[Yor10c] G.H. Yordanov, O.M. Midtgård, T.O. Saetre, J.O. Odden, T. Buseth, A.G.
Imenes, C.L. Nge, and L. Norum, “Compensated SoG-Si from a merallurgical
route: High latitude outdoor performance”, Proc. 25th EUPVSEC, Valencia, pp.
4289–4293, 2010.

[Yor11a] G.H. Yordanov and O.-M. Midtgård, “Modeling and parameter identification
of crystalline silicon photovoltaic devices”, Proc. 37th IEEE PVSC, Seattle, pp.
002972–002976, 2011.

[Yor11b] G.H. Yordanov, O-M. Midtgård, and L. Norum, “Modeling and parameter
identification of crystalline silicon photovoltaic devices”, Proc. International
Conference on Clean Electrical Power, Ischia, pp. 574–577, 2011.

[Yor12a] G.H. Yordanov, O.-M. Midtgård, and T.O. Saetre, “Series resistance
determination and further characterization of c-Si PV modules”, Renewable
Energy 46, pp. 72–80, 2012.

[Yor12b] G.H. Yordanov, O.-M. Midtgård, T.O. Saetre, H.K. Nielsen, and L.E. Norum,
“Over-irradiance (cloud enhancement) events at high latitudes”, presented at
the 38th IEEE PVSC, Austin, 2012.

[Yor12c] G.H. Yordanov, O.-M. Midtgård, T.O. Saetre, H.K. Nielsen, and L.E. Norum,
“Overirradiance (cloud enhancement) events at high latitudes”, IEEE Journal
of Photovoltaics, DOI: 10.1109/JPHOTOV.2012.2213581, 2012.
313

[Yor12d] G.H. Yordanov, O.-M. Midtgård, and T.O. Saetre, “PV modules with variable
ideality factors”, Proc. 38th IEEE PVSC, Austin, pp. 002362–002367, 2012.

[Yor12e] G.H. Yordanov, O.-M. Midtgård, and T.O. Saetre, “Equivalent cell
temperature calculation for PV modules with variable ideality factors”, Proc.
38th IEEE PVSC, Austin, pp. 000505–000508, 2012.

[Yor12f] G.H. Yordanov, M. Tayyib, O.-M. Midtgård, J.-O. Odden, and T.O. Saetre,
“Test of the European Joint Research Centre performance model for c-Si PV
modules”, Proc. 38th IEEE PVSC, Austin, pp. 002382–002387, 2012.

[Zeg11] A. Zegaoui, P. Petit, M. Aillerie, J.P. Sawicki, A.W. Belarbi, M.D. Krachai,
and J.P. Charles, “Photovoltaic cell/panel/array characterizations and modeling
considering both reverse and direct modes”, Energy Procedia, 6, pp. 695–703,
2011.

[Zeh09] M. Zehner, P. Fritze, M. Schlatterer, T. Glotzbach, B. Schulz, C. Vodermayer,


M. Mayer, and G. Wotruba, “One year round robin testing of irradiation
sensors – measurement results and analyses”, Proc. 24th EUPVSEC, Hamburg,
pp. 3800–3805, 2009.

[Zeh10] M. Zehner, M. Hartmann, J. Weizenbeck, T. Gratzl, T. Weigl, B. Mayer, G.


Wirth, M. Krawczynski, T. Betts, R. Gottschalg, A. Hammer, B. Giesler, G.
Becker, and O. Mayer, “Systematic analysis of meteorological irradiation
effects”, Proc. 25th EUPVSEC, Valencia, pp. 4545–4548, 2010.

[Zeh11] M. Zehner, T. Weigl, M. Hartmann, S. Thaler, O. Schrank, M. Czakalla, B.


Mayer, T.R. Betts, R. Gottschalg, K. Behrens, G.K. Langlo, B. Giesler, G.
Becker, and O. Mayer, “Energy loss due to irradiance enhancement”, Proc.
26th EUPVSEC, Hamburg, pp. 3935–3938, 2011.

[Zha94] J. Zhao, A. Wang, P.P. Altermatt, S.R. Wenham, and M.A. Green, “24%
efficient silicon solar cells”, Proc. 24th IEEE PVSC, vol. 2, Waikoloa, pp.
1477–1480, 1994.

[Zho11] Q. Zhou, X. Hu, K. Al-Hemyari, K. McCarthy, L. Domash, and J.A. Hudgings,


“High spatial resolution characterization of silicon solar cells using
thermoreflectance imaging”, Journal of Applied Physics 110, ID: 053108, 6 pp,
DOI: 10.1063/1.3629979, 2011.

[Zhu09] J. Zhu, Y. Qui, T.R. Betts, and R. Gottschalg, “Outlier identification in outdoor
measurement data – effects of different strategies on the performance
descriptors of photovoltaic modules”, Proc. 34th IEEE PVSC, Philadelphia, pp.
000828–000833, 2009.
314

[Zhu11] J. Zhu, R. Bründlinger, T. Mühlberger, T. Betts, and R. Gottschalg, “Optimised


inverter sizing for photovoltaic systems in high-latitude maritime climates”,
IET Renewable Power Generation 5, pp. 58–66, 2011.

View publication stats

You might also like