You are on page 1of 153

Veikko Schepel

1
The Dutch PV Portal 2.0

By

Veikko Schepel

in partial fulfilment of the requirements for the degree of

Master of Science

in Sustainable Energy Technology

at the Delft University of Technology,

to be defended publicly on Thursday March 8, 2018 at 9:30 AM.

Supervisor: Dr. ir. O. Isabella

Thesis committee: Prof. dr. M. Zeman, TU Delft

Dr. ir. H. Ziar, TU Delft

Dr. ir. S. Tindemans, TU Delft

An electronic version of this thesis is available at http://repository.tudelft.nl/.

Front page image: https://www.videoblocks.com/video/netherlands-on-the-satellite-map-outlined-and-glowed-elements-of-


this-image-furnished-by-nasa-scwg6vv

2
ACKNOWLEDGMENTS
First of all I want to thank my thesis supervisor Dr. Olindo Isabella, who has provided me with
guidance and invaluable help in this project. I owe my gratitude to Arianna Tozzi, whose design of the
PV Portal 1.0 I had the privilege to build upon. I am also indebted to the Google Corporation, who
created the JavaScript chart and map APIs which I have found so useful for my website. Lastly, and
most importantly, I want to thank my family for supporting me throughout my academic career.

Delft, 2 March 2018

3
ABSTRACT
The growing importance of solar energy, or photovoltaic (PV) energy, worldwide in recent years
has increased the need for publically available, reliable information on PV technology. The internet
has become a channel for the communication of knowledge to a wide audience, which has led to the
development of numerous websites on solar energy. However, many of these sites lack a clear
scientific justification for the communicated information. In 2014, the Photovoltaic Materials and
Devices (PVMD) group of Delft University of Technology launched its own solar energy website titled
the Dutch PV Portal (PVP 1.0), which provides visitors with scientific information about PV together
with an interactive PV system performance calculator that uses real-time weather measurements for
the Netherlands.

Two years after its launch, an evaluation of the website's effectiveness culminated in a Master
thesis project to create a new website that would incorporate additional features into the Dutch PV
Portal framework, thereby improving upon the original website. The realization of this new website,
under the moniker Dutch PV Portal 2.0 (PVP 2.0), involved the execution of five objectives: (1)
creating a meteorological database storing real-time and annual data relevant to PV system modelling
of any location in the Netherlands, (2) creating a model to calculate the annual energy output of a
user-designed PV system, (3) creating a figure displaying the real-time efficiency losses of a PV
system, (4) creating a model to estimate the Dutch national solar energy production (NSEP), and (5)
creating an economic profitability analysis of user-designed PV systems.

The approach used to realize the objectives was to combine information from literature research
with multiple datasets of the characteristics of Dutch PV systems provided by external organizations
and incorporate the information in the website code. The validity of the resultant outputs of the
models was checked against literature. The website incorporated a visual representation of the
results of the models.

For each of the five objectives, the most important findings were as follows: (1) the spatial
granularity of the PVP 2.0 database improved significantly upon the PVP 1.0 by storing data for 46
weather stations throughout the country. A distinctive feature of the database is that the annual
datasets are updated every hour with real-time measurements, making the database dynamic instead
of static. (2) An annual energy calculation was implemented that incorporated two novel features: a
dynamic soiling loss calculation based on local rainfall patterns and an inverter sizing factor that
implements a larger or smaller inverter capacity based on the module orientation thereby increasing
the modelled DC to AC conversion efficiency. (3) The efficiency breakdown figure clearly displayed the
effect of real-time meteorological changes on the system efficiency. (4) An estimate of the NSEP was
made through a simulation of a set of 475 PV systems representing the characteristics of the Dutch
installed solar panel capacity. The PVP 2.0 model found higher values for real-time and annual energy
production than represented in the official government estimates. Future research is required to
further assess the degree of accuracy of the PVP 2.0 NSEP estimate. (5) An economic analysis showed
that the profitability of Dutch PV systems is in large part attributable to the governmental legislation
and subsidies supporting renewable energy. The uncertain future direction of this policy could harm
the adoption rate of PV systems in the Netherlands.

The successful execution of the five objectives have made the PVP 2.0 a website that collects and
displays relevant, scientifically accurate, and publically accessible information on solar energy within
the Netherlands. The website can be viewed at http://tudelft.nl/pvp2/.

4
TABLE OF CONTENTS
1 Introduction .......................................................................................................................... 18

1.1 Main introduction ............................................................................................................. 18

1.2 Website introduction ........................................................................................................ 21

2 Meteorological database ...................................................................................................... 23

2.1 Introduction ...................................................................................................................... 23

2.2 Motivation and objectives ................................................................................................ 23

2.3 Methodology for database creation ................................................................................. 26

2.3.1 Data source selection and database structure ........................................................ 26

2.3.2 Weather station locations........................................................................................ 27

2.3.3 Climate database construction ................................................................................ 30

2.3.4 Additional parameter calculation ............................................................................ 33

2.3.5 Weather data upload to database ........................................................................... 35

2.3.6 Real-time weather database and connection to climate database ......................... 35

2.4 Results .............................................................................................................................. 38

2.4.1 Climate dataset validation ....................................................................................... 38

2.4.2 Calculated parameter evaluation ............................................................................ 43

2.4.3 Integration in website .............................................................................................. 46

2.5 Discussion ......................................................................................................................... 47

2.6 Conclusion ........................................................................................................................ 48

3 PV system performance model ............................................................................................. 50

3.1 Introduction ...................................................................................................................... 50

3.2 Model description............................................................................................................. 50

3.3 Model input parameters ................................................................................................... 52

3.4 In-plane irradiance ............................................................................................................ 53

3.4.1 Tilted surface model ................................................................................................ 53

3.4.2 Light capture losses.................................................................................................. 56

3.5 PV system conversion efficiency....................................................................................... 59

5
3.5.1 Module conversion efficiency .................................................................................. 60

3.5.2 Temperature and irradiance effect on DC module efficiency .................................. 61

3.5.3 Additional component DC losses ............................................................................. 64

3.5.4 DC to AC conversion efficiency ................................................................................ 64

3.6 Overview of inputs in the performance model ................................................................ 66

3.7 Results and model evaluation .......................................................................................... 67

3.7.1 Overall model performance ..................................................................................... 68

3.7.2 Individual component evaluation ............................................................................ 70

3.7.3 Meteorological dataset comparison ........................................................................ 77

3.8 Discussion ......................................................................................................................... 87

3.9 Conclusion ........................................................................................................................ 88

4 System efficiency breakdown ............................................................................................... 90

4.1 Introduction ...................................................................................................................... 90

4.2 Efficiency losses overview ................................................................................................ 90

4.3 Efficiency loss determination ........................................................................................... 91

4.3.1 Light capture losses.................................................................................................. 91

4.3.2 Photovoltaic conversion losses ................................................................................ 91

4.3.3 Balance-of-system losses ......................................................................................... 95

4.4 Results .............................................................................................................................. 95

4.5 Discussion ......................................................................................................................... 96

4.6 Conclusion ........................................................................................................................ 97

5 National solar energy production model .............................................................................. 98

5.1 Introduction ...................................................................................................................... 98

5.2 Installed capacity and NSEP .............................................................................................. 99

5.3 Existing NSEP models ........................................................................................................ 99

5.3.1 CBS ........................................................................................................................... 99

5.3.2 TenneT ................................................................................................................... 100

5.3.3 iCarus/Alliander ..................................................................................................... 100

5.3.4 EnTranCe ................................................................................................................ 100

6
5.4 The PVP 2.0 NSEP model ................................................................................................ 101

5.4.1 Proposed method .................................................................................................. 101

5.4.2 Location ................................................................................................................. 102

5.4.3 System size ............................................................................................................. 103

5.4.4 Panel technology.................................................................................................... 105

5.4.5 System type............................................................................................................ 106

5.4.6 Orientation ............................................................................................................. 106

5.4.7 Installed capacity ................................................................................................... 112

5.4.8 Final model ............................................................................................................ 113

5.5 Results ............................................................................................................................ 114

5.5.1 Estimate of the national solar energy production ................................................. 114

5.5.2 Comparison with EnergieOpwek and ENTSO-E...................................................... 114

5.5.3 Website output ...................................................................................................... 117

5.6 Discussion ....................................................................................................................... 118

5.7 Conclusion ...................................................................................................................... 119

6 Economic profitability model .............................................................................................. 120

6.1 Introduction .................................................................................................................... 120

6.2 Economic profitability indicators .................................................................................... 120

6.2.1 Net present value................................................................................................... 121

6.2.2 Payback period....................................................................................................... 121

6.2.3 Compound annual growth rate .............................................................................. 122

6.2.4 Levelised Cost of Electricity ................................................................................... 122

6.2.5 Economic model flowchart .................................................................................... 123

6.3 Input parameter values .................................................................................................. 123

6.3.1 System lifetime ...................................................................................................... 124

6.3.2 Initial investment ................................................................................................... 124

6.3.3 Operation and maintenance costs ......................................................................... 129

6.3.4 Annual energy production ..................................................................................... 130

6.3.5 Annual revenue ...................................................................................................... 130

7
6.3.6 Discount rate.......................................................................................................... 132

6.3.7 Summary of applied values .................................................................................... 133

6.4 Results ............................................................................................................................ 133

6.5 Discussion ....................................................................................................................... 134

6.6 Conclusion ...................................................................................................................... 135

7 Overall conclusion ............................................................................................................... 136

7.1 Extent of achievement of objectives .............................................................................. 136

7.2 Future research .............................................................................................................. 137

8 References .......................................................................................................................... 139

9 Appendix ............................................................................................................................. 145

9.1 Selection of PHP scripts .................................................................................................. 145

9.1.1 Weather and climate database script .................................................................... 145

9.1.2 Economic model .................................................................................................... 146

9.2 Missing weather parameters .......................................................................................... 150

9.3 Summary of communication with Stedin and TenneT ................................................... 151

9.3.1 Telephone call with Marko Kruithof at Stedin (25/11/2016)................................. 151

9.3.2 Telephone call with Fieke ‘t Hoen at TenneT (18/01/2017) .................................. 151

8
LIST OF FIGURES
Figure 1: Prediction of growth in worldwide installed solar energy capacity up to 2050 according
to the ITRPV-report, from which the figure is taken [2]. The blue curve indicates actual panel
shipments since 2000. From 2017 onwards, the orange curve projects the annual market for PV, i.e.
the amount of capacity produced and shipped each year. The area curves show the total cumulative
production capacity installed in the respective regions. ........................................................................ 18

Figure 2: Increase in internet users worldwide since 1997 [5]........................................................ 19

Figure 3: Interrelation of the five objectives of this thesis. The weather and climate database is
fundamental for the realization of the PV system performance model, which in turn enables the final
three objectives. ..................................................................................................................................... 21

Figure 4: Layout of the PVP 2.0 website. Purple arrows denote hyperlinks, blue boxes denote
webpages, and orange boxes describe the content of the webpages. .................................................. 22

Figure 5: Overview of the real-time (weather) and historical (climate) database setup, as well as
the interconnection between the two databases. ................................................................................. 27

Figure 6: Location of the 46 KNMI weather stations used in the PVP 2.0. ...................................... 28

Figure 7: All locations in the Netherlands lie within a radius of 35.4 km around a weather station,
as can be seen by the complete coverage of the country. ..................................................................... 29

Figure 8: Method to create a weighted average from multiple values (this can be any amount of
values) and the respective number of measurements ('counts') that contributed to this value. The
method is applied to create a new climate average from the old climate average and new weather
measurements. ....................................................................................................................................... 32

Figure 9: Inclusion of the delayed measurements in the final provincial weather averages at a
given timestamp. The preliminary average is used as input in the real-time power model, as it is the
most recent value available. ................................................................................................................... 37

Figure 10: Annual sum of global horizontal irradiance for all 46 climate datasets of the PVP 2.0.
The position in the figure represents the station’s geographical coordinates. ...................................... 39

Figure 11: Annual sum of global horizontal irradiance over the period of 1981-2010 as calculated
2 2
by the KNMI. The units are in kJ/cm instead of kWh/m . The figure is taken from [21]. ..................... 40

Figure 12: Comparison of the PVP 2.0 climate average for ambient temperature and the monthly
temperature averages for 1981-2010 of the KNMI. ............................................................................... 41

Figure 13: Comparison between the daily sum of global horizontal irradiation values of
Meteonorm and the PVP 2.0 climate database...................................................................................... 42

Figure 14: Comparison between the hourly ambient temperature values of Meteonorm and the
PVP 2.0 database. ................................................................................................................................... 42

Figure 15: Contribution of diffuse and direct horizontal irradiance to GHI in the PVP 2.0 climate
dataset. It is clear that the ABD method results in much higher diffuse fractions than the DBA method.
................................................................................................................................................................ 43

9
Figure 16: Annual sum of diffuse and direct irradiation for the Meteonorm TMY, the KNMI TMY,
the ABD PVP 2.0 dataset (referred to as original climate), and the DBA dataset (referred to as adjusted
climate). .................................................................................................................................................. 44

Figure 17: Total annual horizontal irradiation for the KNMI weather station of Eelde. ................. 45

Figure 18: Direct and diffuse irradiance for every hour in the annual KNMI TMY and climate
dataset for Eelde. The Meteonorm TMY (which is not shown in this figure) showed a similar variability
in diffuse and direct irradiance............................................................................................................... 46

Figure 19: Screenshot of the climate data viewer in the PVP 2.0 is displayed. At the top of the
page other climate parameters can be selected. ................................................................................... 47

Figure 20: Overview of the PVP 2.0 PV system performance model, which will be discussed in
detail in the rest of this chapter. ............................................................................................................ 52

Figure 21: Relation between the irradiance vectors Gdirect, DNI and Gbeam via the module tilt θM,
angle of incidence θAOI and zenith angle θz. ........................................................................................... 55

Figure 22: Shading factor distribution of the 5398 simulated PV systems of Solar Monkey. ......... 57

Figure 23: Determination and long-term storage of the daily and climate rain-free period in the
PVP 2.0 database. RT stands for real-time values (as opposed to climate values). ............................... 59

Figure 24: Dependence of the module VOC on the ideality factor. The voltage dependence used in
the PVP 1.0 is displayed as well. The module characteristics of the Panasonic HIT module were used.
................................................................................................................................................................ 63

Figure 25: Relative sensitivity of the PVP 2.0 modules to low-light conditions. The top cluster
contains four modules, i.e. mono-Si, HIT and a-Si. The middle cluster is the CdTe and poly-Si, while
CIGS is most sensitive. ............................................................................................................................ 63

Figure 26: VOC dependence on irradiance as zoomed-in from the previous figure to distinguish the
different module lines. ........................................................................................................................... 64

Figure 27: Difference in module temperature dependence on irradiance for the BAPV and FS
systems. .................................................................................................................................................. 70

Figure 28: Hourly calculated module temperature for the BAPV and FS systems using the Eelde
climate dataset. ...................................................................................................................................... 71

Figure 29: Difference between TM for the BAPV and FS system under the same meteorological
conditions. .............................................................................................................................................. 71

Figure 30: Comparison of TM as calculated with the INOCT, DB, and FD model for a BAPV system.
................................................................................................................................................................ 72

Figure 31: Comparison of TM as calculated with the INOCT, DB, and FD model for a FS system. ... 73

Figure 32: Average wind speed for the Eelde climate dataset of the PVP 2.0. ............................... 74

Figure 33: Annual rainfall measured at the weather station of Eelde in 2016 and the average of all
measurements between 1991-2017. ..................................................................................................... 75

10
Figure 34: Rain-free period for 2016 and the average daily RFP for 1991-2017. An increase in RFP
for a set of days means that during this period daily rainfall does not exceed 2 mm. .......................... 75

Figure 35: Comparison of the PV system power output and the respective efficiencies of a small
inverter selected with the ISF and a large inverter with an input capacity equal to the installed PV
capacity. The module tilt of the system is 40° and the azimuth is 0° (north). ........................................ 77

Figure 36: PR breakdown for a BAPV system using the three meteorological datasets. ................ 78

Figure 37: PR breakdown for a FS PV system using the three meteorological datasets. ................ 79

Figure 38: Relation of module temperature deviation from ambient temperature and incident
irradiance................................................................................................................................................ 81

Figure 39: Distribution of module temperature for every hour in the PVP 2.0 climate dataset for
Eelde. ...................................................................................................................................................... 82

Figure 40: Distribution of module temperature for every hour in the KNMI TMY dataset for Eelde.
................................................................................................................................................................ 82

Figure 41: Distribution of module temperature for every hour in the Meteonorm TMY dataset for
Eelde. ...................................................................................................................................................... 83

Figure 42: Percentage difference between HPoA for modules with a 270° and 90° tilt for the 46
stations. .................................................................................................................................................. 84

Figure 43: DNI vs Sun elevation for the PVP 2.0 climate dataset of Vlieland, differing between
values for when the Sun is in the east vs. when it is in the west of the sky. .......................................... 85

Figure 44: Annual cloud coverage, taken from NASA satellite data measured between 2011-2016.
Figure created by [50]. ........................................................................................................................... 86

Figure 45: Annual cloud coverage, taken from KNMI weather station data measured between
2011-2016. Figure created by [50]. ........................................................................................................ 86

Figure 46: Efficiency losses considered in the system efficiency breakdown. The different box
colors categorize the losses. ................................................................................................................... 91

Figure 47: AM1.5 solar spectrum. The largest amount of power in sunlight occurs at wavelengths
below 1500 nm. ...................................................................................................................................... 93

Figure 48: Ultimate conversion efficiency under AM1.5. The maximum conversion efficiency of
49.1% occurs at a bandgap of 1.12 eV, exactly the bandgap of silicon. Note that the x-axis has changed
from wavelength to bandgap energy. .................................................................................................... 94

Figure 49: Efficiency breakdown for a CIGS PV system. The blue area displays the final system
efficiency. When the Sun is below the horizon, system efficiency goes to zero and the figure displays
only one value for “No sunlight”. ........................................................................................................... 96

Figure 50: PVP 2.0 NSEP real-time estimation model. EY refers to electricity yield in W/Wp. ..... 101

Figure 51: Klimaatmonitor distribution key of installed solar capacity over the twelve Dutch
provinces as shown on the PVP 2.0 website. ....................................................................................... 103

11
Figure 52: System size of residential PV systems in the Stedin territory. The 10 kWp bin contains
systems exceeding 10 kWp as well. They are grouped in this bin to reduce the number of system sizes
input into the PVP 2.0 design portfolio. ............................................................................................... 104

Figure 53: Tilt distribution of the three datasets employed for the PVP 2.0. ............................... 108

Figure 54: Closer look at the tilt distribution of the Solar Monkey and TU/e datasets. ................ 108

Figure 55: Azimuth distribution of the three datasets employed for the PVP 2.0. ....................... 109

Figure 56: Closer look at the azimuth distribution of the TU/e and iCarus datasets. ................... 110

Figure 57: Final tilt distribution, taken from the Solar Monkey dataset, is limited between 10 and
55°, resulting in 10 different options.................................................................................................... 111

Figure 58: Final azimuth distribution, taken from the iCarus dataset, is limited between 140 and
220°, resulting in 9 different options.................................................................................................... 111

Figure 59: Installed capacity as displayed on the PVP 2.0 website. Data until 2016 is taken from
the CBS, for 2017 and 2018 it is based on the Solar Solutions data. .................................................... 113

Figure 60: Comparison of the NSEP estimate of the PVP 2.0 and TenneT for 4 February, 2018. . 115

Figure 61: Comparison of the NSEP estimate of the PVP 2.0 and TenneT for 6 February, 2018. . 116

Figure 62: Comparison of the NSEP estimate of the PVP 2.0 and TenneT for 8 February, 2018. . 116

Figure 63: NSEP as calculated by the PVP 2.0 model for 8 February, 2018. .................................. 117

Figure 64: Breakdown by province of the NSEP for 6 February, 2018. The interactive date selector
allows a user to find any date of choice. .............................................................................................. 118

Figure 65: Overview of the general model to calculate four economic profitability indicators. Blue
boxes are numerical inputs and purple boxes are factors influencing these numerical inputs over the
system lifetime. .................................................................................................................................... 123

Figure 66: Correlation of installed capacity and specific system cost for the 1456 Solar Monkey
systems. No discernable relation between capacity and specific cost is observed. The displayed values
are excluding VAT. ................................................................................................................................ 125

Figure 67: Breakdown of the total specific system cost, based on the information provided by a
solar installation company in Delft. The displayed values are excluding VAT. ..................................... 126

Figure 68: Average consumer electricity price in the Netherlands for the period of 2010-2016
[103]. .................................................................................................................................................... 131

Figure 69: Screenshot of the economic profitability chart on the website. For every year, the total
profit is compared to the net present value at a 3% and 7% discount rate. The designed system is an
optimally-oriented residential BAPV system with Trina Solar mono-Si panels. ................................... 134

12
LIST OF TABLES
Table 1: Overview of the 46 KNMI stations in the Netherlands: the number per province and the
station code and location. Station 280 (Eelde) lies on the border between Groningen and Drenthe. .. 30

Table 2: Weather parameter overview for the KNMI and PVP 2.0. ................................................ 31

Table 3: Weather data averaging for 1-hr timeslots. Row values are averaged whilst absent or
invalid data are excluded........................................................................................................................ 31

Table 4: PV system design inputs in the PVP 2.0. ............................................................................ 52

Table 5: Overview of the solar panels used in the PVP 2.0. ............................................................ 53

Table 6: Optimal inverter sizing factors in %, as calculated by [42]. The installed PV capacity
should be divided by the ISF to find the best nominal inverter power. ................................................. 66

Table 7: Summary of the loss determination method and source applied to each step in the PV
system performance model. .................................................................................................................. 67

Table 8: Difference between BAPV and free-standing performance calculation in the PVP 2.0
model...................................................................................................................................................... 67

Table 9: PR for a BAPV and FS system in Eelde with a tilt of 30° and an azimuth of 180°. ............. 69

Table 10: PR as calculated by the PVP 2.0 and PV*Sol for a BAPV system in Herwijnen. ............... 69

Table 11: Non-STC irradiance and temperature PR losses. A negative loss indicates a performance
ratio gain, i.e. a positive effect on the PV system performance. ........................................................... 79

Table 12: Annual average values for module and ambient temperature of the three
meteorological datasets. ........................................................................................................................ 80

Table 13: GPoA calculated for Eelde during 7 normal weather years (leap years) of the KNMI. ...... 85

Table 14: Coverage factor for the PVP 2.0 panels. The values for cell and module area were taken
from the manufacturer datasheets. ....................................................................................................... 91

Table 15: Non-absorption, thermalization, and bandgap utilization efficiencies for the PVP 2.0
module technologies are displayed. ηv*FF was inferred from the values for ηtheor, ηabsand ηuse. .......... 95

Table 16: The installed capacity in each of the 12 Dutch provinces, as well as the installed capacity
normalized to the number of inhabitants per province in the Netherlands in 2015 [71]. ................... 102

Table 17: System size for the four portfolio subsets. .................................................................... 104

Table 18: Division of module capacity by technology. The data represents the produced or
installed capacity in a certain year. ...................................................................................................... 105

Table 19: Modules selected per technology and their associated efficiency. ............................... 106

Table 20: Division between residential/BAPV and commercial/FS PV systems in the Netherlands at
the end of 2017. ................................................................................................................................... 106

13
Table 21: Portfolio of the 475 PV systems representative of national installed capacity. ............ 114

Table 22: Specific system cost values found in the datasets and in literature. All residential
systems exclude the 21% VAT for the Netherlands. As commercial systems are not eligible for the VAT
tax exemption under Dutch law, these systems include VAT. ............................................................. 128

Table 23: Specific cost of the solar panels used in the PVP 2.0, excluding VAT. *The price of the
flexible a-Si module is unknown, but assumed equal to the CdTe panel. ............................................ 129

Table 24: Input values for the PVP 2.0 model, based on the values found in literature and datasets
supplied by solar panel companies. The inverter costs are divided by the inverter sizing factor to
include the economic benefit of inverter downsizing. ......................................................................... 133

Table 25: Absent measurement parameters for each KNMI weather station and the stations that
supplied replacement data. Station 323 was not included as one of the final 46 stations as it was
discontinued in 2014, but was used to provide replacement data for 5 stations. ............................... 150

14
LIST OF SYMBOLS AND
ABBREVIATIONS
GREEK SYMBOLS
α – surface roughness factor

δ - Sun declination

η - efficiency

θL - local mean sidereal time

θM - module tilt

θz - zenith angle

λ - latitude of the observer with respect to the Earth's equator

ν - photon frequency.

ρ - surface albedo

LATIN SYMBOLS AND ABBREVIATIONS


Ai - Anisotropy index

AM - Air mass.

AOI - Angle of incidence

APV - solar panel area

As - Sun azimuth.

as - sun altitude.

BAPV - Building-added photovoltaics.

BIPV - Building-integrated photovoltaics

C - cloud coverage

Ct - total cost in year t

CAGR - compound annual growth rate

CF - coverage factor

DF - deposition factor

15
DHI - diffuse horizontal irradiance

DNI - direct normal irradiance

DPBP - discounted payback period

DSO - distribution system operator

DSV - dust settling velocity

E0 - eccentricity correction factor

EPV - annual PV system energy production

EY - energy yield

FD - fluid-dynamic

FS - free-standing photovoltaics

Galbedo - ground-reflected irradiance component on a module surface

Gbeam - direct irradiance component on a module surface

Gclear-sky - Clear-sky irradiance

Gdiffuse- equivalent to/synonym of DHI

Gdiffuse-PoA - diffuse irradiance component on a module surface

Gdirect - direct horizontal irradiance

Gglobal - synonym of GHI

GHI - global horizontal irradiance

GPoA - global irradiance incident on the plane-of-array of a module

Greal - available sunlight, equivalent to GHI

h - hour angle

hM - module height

hP - Planck constant

href – reference height for wind speed measurements

k - clearness index

kB - Boltzmann constant

LCoE - levelized cost of electricity

Lt - lifetime

MySQL - My structured query language

16
n - ideality factor

NPV - net present value

NSEP - national solar energy production

PBP - payback period

PD - performance degradation

PHP - PHP hypertext preprocessor

PR - performance ratio

PVP - Dutch PV Portal

q - elementary charge

r – discount rate

Rt - total revenue in year t

RFP - rain-free period

S0 - solar constant

SF - soiling factor

STC - standard test conditions

Tamb - ambient temperature

TM - module temperature

TMY - typical meteorological year

Tsky - sky temperature

TSO - transmission system operator

z0 – terrain roughness

17
1 INTRODUCTION
1.1 MAIN INTRODUCTION

Photovoltaic energy conversion refers to the transformation of sunlight energy into useable
electricity. The devices that can achieve this conversion are called photovoltaic energy systems, or PV
systems for short [1]. PV energy is more commonly known as solar energy. The advantages of solar
energy as a clean energy source are increasingly being recognized worldwide. In the past decade the
installed capacity of PV systems, i.e. the power production potential in Watts, has grown
exponentially, as has the annual production of solar panels. The International Technology Roadmap
for Photovoltaics (ITRPV) has predicted in 2017 that the current increase in the global energy market
is just the start of an expanding growth curve for solar panel systems, as shown in Figure 1.

Figure 1: Prediction of growth in worldwide installed solar energy capacity up to 2050 according to
the ITRPV-report, from which the figure is taken [2]. The blue curve indicates actual panel
shipments since 2000. From 2017 onwards, the orange curve projects the annual market for PV, i.e.
the amount of capacity produced and shipped each year. The area curves show the total cumulative
production capacity installed in the respective regions.

The impressive reduction in solar panel production costs and the exponential growth of PV
system capacity has led the president of the China Photovoltaic Industry Association to hail solar
energy as the ‘next energy revolution’ in a recent publication of the World Economic Forum [3].
Similar talk of a ‘clean energy revolution’ has been expressed by the World Bank regarding the growth
of solar energy in India [4].

The rapid development and spread of PV technology call to mind another important
technological revolution that has taken place over the past two decades: the worldwide adoption of
internet communication. The number of people with access to the World Wide Web has risen
exponentially between 1995 and 2017 with growth rates of 8-15% in the past 10 years (see Figure 2).
As of 2017, over four billion people make use of the internet [5].

18
Figure 2: Increase in internet users worldwide since 1997 [5].

The internet revolution has introduced the possibility of global, instantaneous access to
information. As in recent years solar energy is quickly gaining importance in the public perception, the
public demand of accurate information on solar energy will rise in the coming years. In view of the
growing relevance of solar energy, numerous solar energy websites have already been developed that
aim to provide visitors with information on the technology. These websites tend to include interactive
elements that allow users to design a PV system and calculate the expected energy production of
such a system. Notable examples of such websites are PVGIS, a solar radiation database developed by
the European Commission, Sunny Design by the PV system inverter company SMA, PV Education (a
comprehensive scientific overview of PV technology), and, more recently, Project Sunroof developed
by Google.

With the increased relevance of solar energy, the necessity also arises for research institutes to
communicate scientific knowledge on this subject. This is useful for two purposes: firstly, the
presentation of reliable information on the technology can enhance public understanding of and
support for this energy alternative. Secondly, and perhaps more directly, when both citizens and
policy makers have the ability to learn about solar energy, they may be more inclined to opt for, or to
facilitate, the installation of a small-scale or large-scale PV system. It is therefore a priority that the
information communicated by a research institute is as scientifically accurate as possible, that the
website provides a clear methodology for the results, and that the information is communicated in an
clear manner. The research group on Photovoltaic Materials and Devices (PVMD group) in the Delft
University of Technology (TU Delft) in the Netherlands was early to recognize the potential of the
internet as a resource for communicating scientific information on solar energy in an interactive and
accessible way. In 2014 the Dutch Photovoltaic Energy Portal (Dutch PV Portal for short), developed
within the PVMD group by Master student Arianna Tozzi, was launched. The website was created with
three goals in mind [6]:

19
1. To estimate the potential for solar energy production in the Netherlands using real-time
weather measurements.
2. To offer website visitors the option to design small-scale to large-scale PV systems and
to view the real-time power production of these systems.
3. To provide information on the scientific background of PV and the developments in this
field in an understandable manner to both technically educated visitors and to laymen.

The distinguishing feature of the Dutch PV Portal compared to other solar energy websites is that
through a collaboration with the Royal Netherlands Meteorological Institute (Koninklijk Nederlands
Meteorologisch Instituut (KNMI)) it is able to include real-time weather measurements relevant to PV
system calculations [6]. As an additional attribute, the Dutch PV Portal includes a methodology
section providing a justification for the model results.

Two years after its launch in 2014, an evaluation of the effectiveness of the Dutch PV Portal was
made in the PVMD group, including a comparison to the information and features provided by other
solar energy websites. The following issues were identified:

1. An annual energy calculation was not possible, only the calculation of real-time power
production. Moreover, after sundown no system designs could be made as the website
requires sunlight measurements to make a calculation.
2. Weather data in the website were stored at a provincial spatial resolution, which
restricted the user location selection to only one of twelve provinces instead on any
location within the Netherlands.
3. Information on the contribution of solar energy to the total energy production in the
Netherlands was absent.
4. One of the most interesting aspects for people considering adopting PV technology, i.e.
the economic profitability of PV systems, was not addressed in the website.

To strengthen the website with regard to the issues mentioned above, the Master thesis project
of the Dutch PV Portal 2.0 (PVP 2.0) was instigated. As the name implies, the overarching goal of this
thesis project was to create a new version of the first Dutch PV Portal (referred to as PVP 1.0 from
now on) that expands upon the original website and integrates important educational features (that
are often quite hidden) of various other solar energy websites within one domain, taking the
Netherlands as the country of focus. Five specific objectives or features were to be realized for this
thesis project:

1. The creation of a meteorological database from which data can be retrieved that are
applicable to any location within the Netherlands, i.e. with a high geographical resolution.
2. The creation of a model to calculate the annual energy output of a PV system, based on
meteorological data and a user-input PV system design.
3. The creation of a figure to display, in real-time, the efficiency losses occurring within a PV
system.
4. The creation of a model to estimate the national solar energy production for the
Netherlands.
5. The creation of an economic profitability estimate for user-designed PV systems.

These objectives are in line with the three goals of the original PVP 1.0. The five objectives are
interrelated as shown in Figure 3. The meteorological information stored in the database is a
necessary input for the PV system performance model, as the power generation capability of a system
depends on the meteorological conditions. The results of the performance model can subsequently
be applied to the final three objectives. The modus operandi of the PVP 2.0 is to base the realization

20
of these features on recent scientific research, which also entails the incorporation of research
developed within the PVMD group.

Figure 3: Interrelation of the five objectives of this thesis. The weather and climate database is
fundamental for the realization of the PV system performance model, which in turn enables the
final three objectives.

Chapter 2-6 will describe the scientific background behind the realization of each of the five
thesis objectives. In the concluding chapter, the extent to which the five objectives have been
achieved will be evaluated and an outlook to future work on the Dutch PV Portal is given.

1.2 WEBSITE INTRODUCTION

Before the scientific research of this thesis is described, a brief introductory section is in place, to
provide some additional background on the final website that is the result of this Master thesis
project. The Dutch PV Portal 2.0 is now online at http://tudelft.nl/pvp2/ and the reader is encouraged
to browse the website to gain an understanding of the final product that was accomplished by the
research described in this report.

The domain of the Dutch PV Portal 2.0 was built in a LAMP stack, which is a collection of
components that are used together to create websites. LAMP stands for Linux, Apache, MySQL, and
PHP. PHP, an abbreviation of Personal Home Page: Hypertext Preprocessor, is a coding language which
allows for dynamic content. Unlike static HTML, calculations can be made in PHP code and therefore
the webpage information can change based on the input into the PHP code. Facebook is an example
of a dynamic website written in PHP. MySQL, or My Structured Query Language, is a database which
can store bulk information used in the website. Apache is the server technology that handles requests
from the website administrator and from users. The server can communicate with the MySQL
database and retrieve information that can be processed further in a PHP code. The server also runs
website code and creates HTML files that are displayed on the website user's browser. Finally, Linux is
the operating system on which the server runs, just like Windows is the operating system on which
most PCs run.

21
To perform all numerical calculations and communicate with the MySQL database, PHP codes
were created. To output the numerical data stored in the database and calculated in the PHP code in
a visual interface, the programming languages HTML, CSS, and JavaScript were employed. HTML, or
HyperText Markup Language, is the standard code language for websites which browsers use. CSS, or
Cascading Style Sheet, is code that is integrated in HTML to create a unique visual design for the
website. Finally, JavaScript, like PHP, is a dynamic language that can be used to perform calculations.
JavaScript is the language used for the application programming interface (API) of Google. Google has
provided free-to-use APIs for integration in websites. Specifically, the PVP 2.0 has used a Google Map
API to allow users to interactively select a geographic location, and Google Chart APIs which output
numerical data into figures such as bar or line charts that can be updated dynamically.

The website layout is displayed in Figure 4.

Figure 4: Layout of the PVP 2.0 website. Purple arrows denote hyperlinks, blue boxes denote
webpages, and orange boxes describe the content of the webpages.

22
2 METEOROLOGICAL DATABASE
2.1 INTRODUCTION

An important facet of the second version of the Dutch PV Portal (PVP 2.0) is the creation of a new
meteorological database in MySQL, which is compatible with and useful for the other newly
implemented features. The first Dutch PV Portal (PVP 1.0) website does contain a database which
stores weather information for the Netherlands. However, the PVP 1.0 database was deemed
insufficient to implement the additional objectives of the PVP 2.0, most notably the enhancement of
spatial granularity and calculation of annual energy production.

This chapter details the creation of a new database for storing meteorological information, to be
used in the PVP 2.0 website. First the necessity for such a database is discussed, as well as the
objectives to be achieved by the database. Thereafter the methodology for building the database is
described, followed by a presentation of the database in its final form and a discussion on its merits
and limitations.

2.2 MOTIVATION AND OBJECTIVES

For the PVP 1.0 a weather database was created to store real-time weather parameter data as
collected by weather stations in the Netherlands. The most important weather parameter for PV
system modelling is, of course, the amount of available sunlight as indicated by the irradiance
intensity, as this determines to a large extent how much electricity can be generated. Next to
irradiance, additional weather parameters were stored in the database for use as input in a so-called
fluid-dynamic model. The fluid-dynamic model, created in 1987, calculates the temperature of a flat
surface by simulating the heat exchange between the surface and its environment based on a set of
meteorological parameters and the surface characteristics [7]. The fluid-dynamic model can be
applied to PV modules to calculate the module temperature under illumination. The fluid-dynamic
model was incorporated into the PVP 1.0 model as detailed in the Master Thesis by Arianna Tozzi [6,
p. 33]. Using the outcome of the fluid-dynamic model, i.e. the module temperature, the effect of
module temperature on the efficiency of a PV system can be calculated. It follows that the storage of
real-time weather parameter values is a prerequisite for the proper functioning of the website. To this
end, the PVP 1.0 database stored average weather data for all twelve Dutch provinces over a one-day
period. At midnight, all database tables were cleared, after which the tables would be repopulated by
the data of the following day. In other words, no long-term storage of measurements takes place in
the PVP 1.0. As the PVP 1.0 database did not store weather data for a complete year, the PV system
efficiency could only be calculated in real-time, or for one day at most. One of the main objectives of
the PVP 2.0 thesis is to enable the calculation of the annual energy yield of a PV system design. The
prediction of the total PV system efficiency and power production over a full year would require the
storage of meteorological data for an entire year.

The PVP 1.0 does include an estimation of annual energy production for all the photovoltaic
technologies (i.e. monocrystalline/polycrystalline/amorphous silicon, copper-indium-gallium-sulphide
(CIGS) and cadmium telluride (CdTe)) that can be chosen on the website, given a 36° tilt and one of
the eight available system orientations (north, northwest, west etc.). The annual estimates are taken
1
from the PVGIS (which stands for Photovoltaic Geographic Information System) website , developed
by the European Commission Joint Research Centre (JRC). PVGIS utilizes a model that calculates the

1
http://re.jrc.ec.europa.eu/pvgis.html

23
2
amount of solar radiation (in kWh/m per day, month or year) falling on a geographic location from
satellite and weather station data. PVGIS can give an estimate of the amount of energy being
produced by a PV system by calculating the in-plane irradiance for a given tilt and orientation. The
system energy production is calculated by applying typical values for PV module efficiency (which are
dependent on the selected technology) and adjusting module efficiency according to losses due to
temperature and low irradiance. Efficiency losses for light reflectance are calculated dynamically
based on the Sun position. The module temperature is calculated as a function of ambient
temperature, wind speed, incident irradiance, and two factors dependent on the module technology.
For additional system losses such as inverter efficiency and Ohmic cabling losses a standard, fixed
value of 14% is taken [8]. The PVP 1.0 model in contrast includes the fluid-dynamic model as an
alternate temperature calculation and calculates inverter efficiency dynamically, based on the power
input and selected inverter ratings, as inverter efficiency has been shown to be dynamic instead of
static [1, Ch. 20]. In short, the PVP 1.0 provides a more accurate performance estimation given an
irradiance input. It is therefore desirable to adapt the PVP 1.0 model to calculate annual PV system
performance as well as real-time performance. The annual system performance calculation requires
the storage of accurate weather parameter values over an entire year. As an additional advantage the
PVP 2.0 can become more independent by no longer relying on calculations of an alternative website.

The second motive for restructuring the PVP database is to allow for a higher granularity in
weather data to be implemented in the website. The current website has a provincial-level accuracy
of weather data, as real-time data from weather stations spread throughout each of the 12 Dutch
provinces are averaged in order to smooth local fluctuations of irradiance or other parameters.
However, this averaging comes at the cost of spatial resolution. Since 46 real-time weather stations
are present in the Netherlands, this could potentially mean an almost 4-fold increase in spatial
granularity [9].

The final reason to create a new meteorological database is to implement memory features in
the website. In the PVP 1.0, real-time measurements are only stored for a day. With a long-term (i.e.
historical) database, real-time measurements can be averaged with previous measurements and
subsequently stored to form an annual dataset for PV performance modelling. The important
difference between weather and climate should be highlighted here. As stated by the World
Meteorological Organization (WMO), climate is the statistically averaged weather over a defined
period of time [10]. The climate can show the long-term meteorological conditions in a location which
are less apparent in weather because of its large day-to-day variability. The WMO has set the
standard timespan for climate determination at 30 years [10]. Consequently, the averaging of
weather measurements over a period of 30 years can lead to a view of the climate in a certain
location. By retrieving historical weather measurements and averaging these, a so-called base climate
dataset can be created. By continuously including new real-time measurements, the base climate
dataset can be updated. When looking at the PVP 2.0, the use of real-time measurements in this
manner would mean that real-time weather information collected from KNMI is not discarded after
one day, but utilized to build and improve a climatological dataset for each weather station in the
Netherlands (as well as for provinces as a whole). The result would be a comprehensive climatological
data resource for the Netherlands, of which an asset is that fluctuations in weather parameters are
smoothed on a temporal scale without a loss in spatial resolution. A second advantage is that the
continuous storage of new real-time measurements grants a dynamic nature to the climatological
(historical) database: new values lead to an adjustment (or ‘update’) of the older values. The database
therefore becomes more representative of the average weather (i.e. climate) in the Netherlands as
additional real-time values are being stored. This is an interesting feature: for example, nine of the
ten hottest years on record in the Netherlands occurred in the period of 1999-2016 [11]. By using
real-time data to create a moving average for climate, the change in climate can be incorporated into

24
the climate values used in the PVP 2.0. A dynamic climate database would distinguish the PVP 2.0
1
from PVGIS and Meteonorm , which use static climatological means for a defined period as a basis for
their weather datasets [12]. The radiation data used by PVGIS, for example, span 13 years, which is far
lower than the 30-year period defined by the WMO [13]. The 25-30 year climate timespan also
corresponds to the default lifetime estimated for PV systems [1, Ch. 21]. It is an interesting idea to use
average weather data over a 30-year period to determine the average expected annual output of PV
systems over a 25-30 year timespan. The premise behind this idea is that the timescale of the
meteorological data must be representative for the timescale on which the PV system is operative. If
the weather data of one year is taken to model the annual energy output of a PV system, then the
data should be as close as possible to the average meteorological conditions at the location over the
25-year PV system lifespan. With a climate dataset instead of a weather dataset, the long-term
meteorological conditions at a location would be represented in the most comprehensive manner.

The proposed method of creating a climate dataset instead of a weather dataset would entail a
departure from the standard meteorological inputs used in scientific modelling. As argued by Herrera
et al., a good meteorological data file should represent both normal and abnormal weather
conditions, as would occur during any year [14]. This is especially important in building modelling, in
which the researcher must ensure that the building's cooling or heating system is able to cope with
abnormally high or low temperatures [14]. The most common way to simulate the performance of a
PV system is to use the concept of a reference year (often referred to as a test reference year (TRY) or
typical meteorological year (TMY)) [15]. The TMY is constructed by first selecting a time period for
which the data should be representative. For this period, a mean is calculated, and subsequently the
measurement data which lie closest to this mean. This can either be done by individually selecting the
12 months that lie closest to each monthly mean, or to select a single year that is closest to the
annual average weather [16]. The advantage of this method is that the real fluctuations of weather
are still included in the dataset while the dataset represents the inter-annual average as well. A
disadvantage of this method is that it is clearly tied to a specific time period in the past, and therefore
requires the construction of new TMYs in the future [14]. As a noticeable example, the System
Advisory Model (SAM) of the U.S. National Renewable Energy Laboratory (NREL) is one of the most
prominent (offline) computer-aided design (CAD) programs for PV systems. The SAM software is
designed to use TMY datasets as meteorological input for modelling [17]. The explanatory website of
SAM lists databases that provide free or commercial weather datasets. All of these nine websites
provide single-year data (i.e. measurement data for a specific year) or TMY data, but none provide a
climate dataset [17]. This is a clear indication that the approach used in the PVP 2.0 is unique and
departs from the major solar resource databases that are currently available.

The motivation for this departure is to challenge the common format of meteorological datasets
for PV system modelling. By using an alternate approach, it is possible to test the hypothesis that a
dataset with climate averages will provide valid results for PV system performance despite the lower
variability than TMY datasets. The hypothesis lies in the reasoning that the inclusion of extreme
weather conditions is of large importance in building heating simulations, where the design needs to
be made resilient to the worst-case conditions (such as a heatwave). In PV system performance, while
high temperatures will surely adversely affect the power generation capability, these fluctuations will
occur sporadically and will be of a limited duration. In addition, an extreme cold spell in the same year
could have an opposing beneficial effect on solar energy generation, cancelling out the negative effect
of large positive temperature fluctuations. It could therefore be theorized that the difference in
power output between a simulation using TMY and a complete inter-annual average (with extreme

1
Meteonorm is a leading weather data software program, of which the data is often used in PV
system modelling.

25
weather conditions being smoothed by normal values) will be limited. The hypothesis is expanded
upon in the results and discussion of this chapter, and will be tested in the Annual energy calculation
chapter.

The main objectives of the redesigned database are summarized as follows:

1. Storage of datasets representing an entire year of meteorological measurements for annual


energy yield calculation.
2. Storage of measurements of all individual weather stations next to provincial averages in
order to increase spatial resolution of the PVP model.
3. Storage of real-time data measurements in a long-term database, together with previous
station measurements, forming a dynamic climatological dataset for the weather station
location.

2.3 METHODOLOGY FOR DATABASE CREATION

2.3.1 DATA SOURCE SELECTION AND DATABASE STRUCTURE

To create the desired meteorological database, several steps were required. First and foremost, a
reliable source of weather data was found. By reliable it is meant that the measured values should be
available for an extensive time period, and that the measured values should be accurate. The second
step was to determine which types of weather parameters are relevant for the PVP 1.0 model and
thus need to be stored in the database. The weather parameters to be used in the PVP 1.0 model
must be present in the source. If a reliable source for data with all relevant parameters present has
been determined, a method must be formed to collect all the data, adjust it to the proper format for
storage and upload it to the database.

The obvious candidate source for weather data is the Royal Dutch Meteorological Institute
(Koninklijk Nederlands Meteorologisch Instituut (KNMI)) collected from 63 stations throughout the
country. The PVP 1.0 uses near-real-time measurements (with a 10-minute delay between
measurement and availability on the server) collected by the KNMI. To build up the new database, it
makes sense to utilize the available data from the KNMI as the organization owns all Dutch weather
stations and publishes the data. An alternate candidate for weather data would be Meteonorm. The
advantage of KNMI over Meteonorm is that the KNMI has made historical hourly measurement values
for each weather station available on their website, spanning from the first measurement date of the
station until the current date. These hourly values are the average of the real-time measurements
that have a 10-minute time resolution (which are unfortunately not made available by the KNMI).
Meteonorm provides synthetic weather datasets with a 1-hour time resolution: the hourly values are
not directly measured, but generated from monthly station averages with a stochastic (random value
generation) model. Thus the Meteonorm datasets have an inherent inaccuracy compared to actual
hourly measurements as published by the KNMI. The inaccuracy of the Meteonorm stochastic model
is discussed in further detail in [18]. For the development of the PVP 2.0 database the decision was
made to use the measurements of the KNMI. For one weather station, namely weather station Eelde
in the province of Drenthe, the data of Meteonorm were downloaded for use as comparison with the
KNMI data of the Eelde station.

In the previous section, the distinction between real-time measurements representing weather
and historical, averaged measurements representing climate was highlighted. This distinction is
important since both types of data will be relevant for the features of the PVP 2.0. A measure of the
instantaneous power output and efficiency of systems, as present in the PVP 1.0, is a valuable feature

26
to maintain, especially to calculate the real-time solar energy production in the Netherlands (which is
one of the main objectives of the thesis). Climate data is an essential input to determine the expected
long-term performance of PV systems. Since both weather and climate data are to be used, albeit in
distinctive website features, it was decided to create a real-time (weather) database and a historical
(climate) database. The creation of a MySQL database entails the formation of data tables with a
number of columns in which weather parameters will be stored. Both databases have the same table
structure and thus the same weather parameters. The main difference is the temporal resolution of
the data. The real-time weather measurements have a time resolution of 10 minutes, whereas
historical KNMI measurements are only available with a 1-hour resolution. Due to a limitation upon
the storage capacity of the MySQL database and website, it is not an option to store data for a year
with a 10-minute resolution in the current format for all weather stations and 12 provinces.

In Figure 5 a setup of the database system in the PVP 2.0 is shown. The real-time measurements
are stored in the weather database, after which they are averaged with measurements in the climate
database. The resultant value is stored in the climate database. The initial data in the climate
database come from measurements of the KNMI.

Figure 5: Overview of the real-time (weather) and historical (climate) database setup, as well as the
interconnection between the two databases.

The first step towards the database system required determining the suitable weather stations
for which both real-time and historical data were available (section 2.3.2). Subsequently, the weather
parameters necessary for the PVP model were characterized, and the data for these parameters were
downloaded and averaged to form the initial historical measurements for all weather stations and
provinces (section 2.3.3). A method was developed for the calculation of additional weather
parameters that are not measured by KNMI, but are of importance for the PVP model (section 2.3.4).
The final climatological datasets were uploaded to the MYSQL database (section 2.3.5). A database
was also created for real-time weather measurements (section 2.3.6). With both databases in place,
the method to update the climate database with real-time measurements through averaging and
storage was created.

2.3.2 WEATHER STATION LOCATIONS

The KNMI collects real-time data from 63 weather stations within the Netherlands [9]. Of these
63 stations, there are 46 stations for which historical hourly weather values were also available on the

27
KNMI website. This prerequisite is necessary to allow for both a real-time and annual energy
production estimate in the PVP model. The station codes and names per province can be found in
Table 1. Station 279 (Eelde) lies on the border between Groningen and Drenthe and therefore is
assigned to both provinces. The location of all stations can be viewed on a geographical map shown in
1
Figure 6 .

Figure 6: Location of the 46 KNMI weather stations used in the PVP 2.0.

With online software developed by FreeMapTools, a radius was drawn around each weather
station to determine the maximum distance any location in the Netherlands would be removed from
a weather station [19]. For a radius of 35.4 km, a complete coverage of the Netherlands was possible.
The result is shown in Figure 7. It must be noted that most places in the Netherlands (especially the
west of the country) lie much closer to a weather station. It can therefore be said that the selected
weather data in the PVP 2.0 are at most 35.4 km removed from the actual real-time weather of the

1
An interactive map visualizing all the weather stations used in the model is accessible via this
URL:
https://www.google.com/maps/d/viewer?mid=1jjT7DhX5FkDE57nC58EGV1FK26I&ll=52.2605739403
1085%2C5.196000000000026&z=7

28
selected location. According to [18], weather data can reasonably be used within a radius of 50 km
around a weather station. A maximum distance of 35.4 km suggests a very good weather coverage for
the Netherlands by using the 46 selected weather stations.

Figure 7: All locations in the Netherlands lie within a radius of 35.4 km around a weather station, as
can be seen by the complete coverage of the country.

It was decided not to apply an interpolation technique between weather stations, given the
complexity of implementing a truly accurate interpolation method for meteorological data in the PVP
2.0 weather station selection model [20]. Instead of interpolation, the weather at a given location is
assumed to be equal to the measured weather at the nearest weather station. While this leads to an
inherent inaccuracy for real-time data, a distance of 35 km was deemed irrelevant in the
determination of a location's climate, as temporal weather fluctuations on this geographical scale will
have been smoothed by the measurements in previous and subsequent years. The climates of nearby
weather stations and the locations in-between will likely lie closer together than the real-time
weather.

29
Table 1: Overview of the 46 KNMI stations in the Netherlands: the number per province and the
station code and location. Station 280 (Eelde) lies on the border between Groningen and Drenthe.
Province Number of KNMI station code and name
stations
Drenthe 2 279 (Hoogeveen), 280 (Eelde)
Flevoland 3 258 (Houtribdijk), 269 (Lelystad), 273 (Marknesse)
Friesland 4 242 (Vlieland), 251 (Hoorn), 267 (Stavoren), 270 (Leeuwarden)
Gelderland 3 275 (Deelen), 283 (Hupsel), 356 (Herwijnen)
Groningen 4 277 (Lauwersoog), 280 (Eelde), 285 (Huibertgat), 286 (Nieuw Beerta)
Limburg 3 377 (Ell), 380 (Maastricht), 391 (Arcen)
Noord- 4 340 (Woensdrecht), 350 (Gilze-Rijen), 370 (Eindhoven), 375 (Volkel)
Brabant
Noord- 7 209 (Ijmond), 225 (Ijmuiden), 235 (De Kooy), 240 (Schiphol), 248
Holland (Wijdenes), 249 (Berkhout), 257 (Wijk aan Zee)
Overijssel 2 278 (Heino), 290 (Twenthe)
Utrecht 2 260 (De Bilt), 348 (Cabauw)
Zeeland 9 308 (Cadzand), 310 (Vlissingen), 312 (Oosterschelde), 313 (Vlakte v.d.
Raan), 315 (Hansweert), 316 (Schaar), 319 (Westdorpe), 324
(Stavenisse), 331 (Tholen)
Zuid- 4 215 (Voorschoten), 330 (Hoek van Holland), 343 (R'dam-Geulhaven), 344
Holland (Rotterdam)

The number of years available for a weather station is determined by the year in which
measurement operations were started. On the KNMI website, a data download tool is available where
a specified time range (e.g. 1 January 2011 to 31 December 2016) can be chosen and the relevant
parameters can be downloaded for this period. For this project, a time period spanning from 1
January 1991 to 29 January 2017 (the day upon which the data was downloaded) was chosen. The
starting point was chosen to lie between 25 and 30 years before the present date, which is an
approximation of the timespan needed to characterize a location's climate, with a low enough
number of years to allow for a larger influence of real-time measurements on the average climate
values compared to a 40- or 50-year period of historical data [10]. For the majority of weather
stations, measurements were available from 1 January 1991 onward. 15 stations had a later starting
point for measurements, ranging from 18 March 1991 to 15 July 2014.

2.3.3 CLIMATE DATABASE CONSTRUCTION

Weather parameters are measured in weather stations in the Netherlands in near-real-time. This
refers to the fact that 10 measurements at 1-min. intervals are averaged over the 10-min. period.
Next to the real-time data, aggregated one-hour values for each weather station are stored on the
KNMI website and available for download. No average climatological year with hourly values (i.e.
weather values averaged over a range of years to form an indication of the climate in a location) are
available on the KNMI website. The closest data to this are tables with monthly climate averages for
1981-2010 for a number of weather stations, published by the KNMI on the website Klimaatatlas.nl
[21]. However, the time resolution of the published data is not high enough for modelling purposes
and due to the manner of publishing, i.e. PDF documents, it was too labour-intensive to copy all
climate averages for 46 weather stations by hand. It was therefore necessary to create a
comprehensive climatological dataset for the PVP 2.0 from the one-hour measurement values.

In the PVP 1.0 model, six measured weather parameters were used in the power calculation of a
PV system [6]. Rainfall was added as a parameter for the PVP 2.0 bringing the required parameter
number to seven. The amount of rainfall is relevant for the degree to which soiling loss occurs on a

30
panel. Soiling refers to the accumulation of dust on a panel which blocks irradiation from entering the
panel. This topic and the use of rainfall data is elaborated upon in Chapter 3. The average annual sum
of rainfall was calculated for the selected locations and found to be in good agreement with the
average yearly rainfall communicated by the Planbureau voor de Leefomgeving [22]. An overview of
the required parameters, the data provided by KNMI, and any unit translations is given in Table 2.

Table 2: Weather parameter overview for the KNMI and PVP 2.0.
KNMI available data PVP required parameters Unit translation
2
Global irradiation [J/cm ] Global horizontal irradiance 10000[cm2 /m2 ]
2
(GHI) in W/m 3600[s/hr]
6-hr minimum temperature at Ground temperature (Tground) in -
10 cm above ground [°C] °C
Temperature at 1.50 m above Ambient temperature (Tambient) -
ground [°C] in °C
Cloud cover [okta] Cloud cover (N) in okta -
1-hr average wind speed [m/s] Wind speed (F) in m/s -
Pressure [Pa] Pressure (P) in Pa -
Rainfall [mm/hr] Rainfall in mm/hr -

In lieu of Tground, a minimum temperature at 10 cm above ground levels for each 6-hr period is
given by KNMI. This is an issue, since here only one value for each 6-hr period is given. The best
approximation in this case is to assume a linear change in ground temperature between the two
minima for the 6-hr period. This may still lead to a slight discrepancy, as the minimum instead of the
average ground temperature is given, leading to a structural underestimation of the ground
temperature.

Text files with the selected parameters were downloaded for the given period for each of the 46
KNMI weather stations used in the PVP 1.0 instantaneous data retrieval. Subsequently, the data were
imported into the simulation and modeling program MATLAB for data processing.

The next required step was to average each parameter over the time period to create an average
annual 1-hr interval dataset. The result would be an array with 8784 rows, one for each hour in the
year (leap days are accounted for). This was achieved by placing each measurement value in a matrix
where the row denotes the 1-hr period within a year (e.g. 3:00 on 14 October) and the column
denotes the year in which the measurement was taken. Subsequently, all row values were averaged,
yielding an average value for each timeslot. Table 3 denotes an example for wind speed. Non-existent
values were omitted from the averaging process. Leap years were accounted for by adding an
additional 24 empty rows after day 59 (February 28) in non-leap years. Since these rows contain no
values, only the actual measured values in leap years contribute to the average leap day weather
data.

Table 3: Weather data averaging for 1-hr timeslots. Row values are averaged whilst absent or
invalid data are excluded.
Time and Date \ Year 1996 1997 1998-2016 2017 Average wind speed [m/s]
1:30, May 3 2.0 4.4 Data from 1998-2016 3.2 3.4
2:30, May 3 2.5 4.6 3.3 3.6
3:30, May 3 2.4 5.2 2.9 3.3
4:30, May 3 1.7 5.5 3.0 3.8
5:30, May 3 1.5 5.0 3.2 3.0

When determining the climatological value of weather parameters, the measurement values for
the individual years is of importance, but it is also necessary to record the number of years (i.e.

31
timespan) over which the average value is calculated. As climate is defined as a 30-year average, an
average of 20 measurement years provides a better indication of the climate than an average taken
over 3 years, in which individual years with extreme weather could influence the average to a larger
degree. In MATLAB, the number of values present in the measurement matrix rows was calculated.
The number of measurements indicates the weight of the specific row, together forming a so-called
count array for each weather parameter. The use of count values in the climate database is twofold.
Firstly, provincial average weather parameters are calculated with a weighted average of the data of
the individual weather stations. The reason a weighted average was chosen is to account for the
stronger smoothing of extreme weather values that occur when measurements have occurred over a
longer period of time. Weather stations with a larger number of measurement years will have a
higher weight of their average annual values and are thus deemed to give a more reliable indication
of the actual weather in the selected region. For example, if weather station A has measured wind
speed for 14 years, the weight of the average wind speed value will be 14. Weather station B with a
longer measurement period of 27 years will hold a higher weight in the average. If the average wind
speed at 3:00 on 30 October is 4.3 m/s for station A and 5.6 m/s for station B, the average of these
two datasets will be ((4.3*14)+(5.6*27))/(14+27) = 5.156 m/s. This is a significant difference from the
non-weighted average of 4.95 m/s. The weighted average method is visualized in Figure 8.

Figure 8: Method to create a weighted average from multiple values (this can be any amount of
values) and the respective number of measurements ('counts') that contributed to this value. The
method is applied to create a new climate average from the old climate average and new weather
measurements.

Via the described method, per weather station and for each of the 12 provinces in the
Netherlands seven arrays of 8784 rows (24 hours over 366 days) were created, one for each weather
parameter. In addition, seven count arrays of 8784 rows, with the weight of each measurement
average were created.

In the data analysis, problems were encountered with partial data sets, for which next best
solutions (approximations) were implemented. An overview of the measured and absent values for all
weather stations was made, which can be found in Appendix 9.2. For absent variables the dataset
values of the nearest weather station with a complete annual parameter array were used. The
determination of the closest station was performed via application of the Pythagorean Theorem to
the geographical coordinates of the 46 stations. The latitude and longitude of a station with absent
measurements was subtracted from the latitude and longitude of all other 45 stations and the
latitude and longitude differences were squared and summed together. The station with the smallest
square root of this distance sum was determined to be closest to the data with absent
measurements. The reduction in accuracy is expected to be limited given the relatively small distance
between adjacent weather stations.

32
In line with the methodology applied by Reindl et al. [23], throughout the process of creating the
average arrays, the following validity calculations were performed: a check for incomplete datasets
and a check for unrealistic weather values (such as an instance of cloud coverage of 67 okta which
lead to a skewed average). After the two data checks it was concluded that the KNMI data were of
high quality for all weather stations in terms of reliable weather measurements, as no data gap was
present for the weather stations.

2.3.4 ADDITIONAL PARAMETER CALCULATION

To complete the required dataset for the database, the four additional parameters related to the
Sun position and irradiance needed to be calculated as input for the PVP 1.0 model:

1. Sun azimuth (As) in degrees. This is the position of the Sun at given latitude and time, with
respect to the North Pole (0°).
2. Sun altitude (as) in degrees. This is the position of the Sun at given latitude and time, with
respect to the horizon (0°).
2
3. Diffuse fraction of GHI (Gdiffuse) in W/m . This is equivalent to DHI (diffuse horizontal
irradiance).
2
4. Direct fraction of GHI (Gdirect) in W/m . Gdirect is different from direct normal irradiance (DNI),
as Gdirect is the projection of DNI on the horizontal plane. Therefore, Gdirect  DNI  sin(as ) [24].
This subsection details the method used to calculate the four additional parameters. The Sun
azimuth and altitude calculation was based on the Sun position algorithm described by Smets et al.
[1], Appendix E. The calculation of Gdiffuse and Gdirect was performed with an updated version of the
method used by Tozzi [6], based on the calculation of a clearness index k by Woyte et al. [25]. The
clearness index k (also known as kt) was developed through pioneering research of Liu and Jordan into
the diffuse and direct fractions of global irradiance, as summarized by Reindl et al. [23]. Using the
clearness index, an estimate of the diffuse fraction of sunlight can be made using the Liu and Jordan
relation Gdiffuse / Gglobal  f (k) , i.e. where the fraction of diffuse irradiance to global irradiance is a
function of the clearness index. The model used in the PVP 1.0 is based on research of Reindl,
Beckman and Duffie [23] which proposed a new relation of diffuse irradiance fraction and clearness
factor in order to reduce the root mean square error of the Liu and Jordan relation. [6] used the
following formula to calculate the clearness index k:

Gglobal
k
S0 E0  sin sin  cos cos  cos h  (2.1)

In which Gglobal is the GHI as calculated from the KNMI data, S0 is the solar constant (the irradiance
2
of sunlight at the top of the atmosphere, 1367 W/m [25]), E0 is the eccentricity correction factor
which uses the day angle, i.e. the position of the Earth on the elliptic path around the sun, to calculate
the effect of the distance of the Earth to the Sun on the incoming radiation [26, p. 3]. δ refers to the
Sun declination, i.e. the vertical position of the Sun in the celestial sphere, or the angle between the
Sun and the equatorial plane. When the Sun is in the vernal or autumnal equinox, declination is 0 [26,
p. 6]. λ refers to the latitude of the observer with respect to the Earth's equator. For the Netherlands,
the average latitude lies around 52°N. h indicates the hour angle, which is the difference between the
local mean sidereal time θL (the angle between the vernal equinox and the longitude of the observer)
and the right ascension of the Sun (the angle between the vernal equinox and the longitudinal
position of the Sun on the celestial sphere). A simplification can be made here if the altitude of the
Sun is already known, which is the case as Sun altitude and azimuth were calculated beforehand via
[1]. According to [26], the following equation holds:

33
cosz  sin sin  cos cos  cos h  sinas (2.2)

θz denotes the angle between the sun altitude as and the zenith. θz is therefore equal to 90-as. By
using equation (2.2), (2.1) can be replaced by:

Gglobal
k (2.3)
S0 E0 sinas

In this simplified formula, the various components that make up the clearness index can now be
discussed. Clearness index indicates the ratio of the available sunlight (Greal, the numerator) to that
which would be available for a clear-sky situation without the influence of clouds or scattering by the
Earth's atmosphere (Gclear-sky, the denominator). If the path of the Earth around the Sun would be
perfectly circular, E0 would have a constant value of 1. The factor 1 sinas is defined as the air mass
(AM) [1]. AM quantifies the length of the path sunlight has to travel through the atmosphere, which is
longer at low sun altitudes. An AM of 1 means that the sun is perpendicular to the Earth surface at the
observer latitude (the Sun is in zenith), and will have to travel the minimum distance through the
atmosphere to reach the surface. Here an additional correction to the formula applied by [6] can be
made: due to the Earth's curvature, the actual AM value for a given a s is slightly lower, as the top of
the atmosphere curves away from the sun rays [1]. Therefore, the sunlight will have to travel through
less air mass. The correction to AM as proposed by [27] is as follows:

1
AM(as ) 
sinas  0.50572(6.07995  as )1.6364 (2.4)

An additional advantage of this correction is that for altitude angles of 0 °, AM will no longer
extend to infinity, but instead reach an AM(0) of 37.9.

Using equations (2.3) and (2.4), clearness index k therefore is equivalent to

Greal Gglobal Gglobal 1


k   AM(as )  
Gclear sky S0E0 S0E0 sinas  0.50572(6.07995  as )1.6364 (2.5)

The clearness index takes both astronomical (as, E0) and climatological factors (Gglobal) into
account. In general, when the weather is cloudy, Greal or Gglobal will be lower than the expected value
for the given time of year and time of the day (Gclear-sky). In essence this means the sky is less clear
than during cloudless weather, and light will be more diffuse. [25] mentions that a bimodal pattern
can be recognized in which the most frequent sky clearness occurrences are either fully clear or
overcast. The occurrence of in-between, e.g. skies with scattered clouds is quite low compared to the
other two options. The division of the sky clearness index into three categories is the basis of the step
function applied by [23]. They determined intervals for k in which the correlation between G diffuse and
k was highest:

Gdiffuse  Gglobal (1.02  0.248  k) (0  k  0.3)


Gdiffuse  Gglobal (1.45  1.67  k) (0.3  k  0.78) (2.6)
Gdiffuse  Gglobal  0.147 (0.78  k  1)

34
Gdiffuse is the diffuse component of global radiation. The step function shown in equation (2.6) was
shown to be applicable for hourly irradiance values, which is the time resolution of the historic data
available from KNMI [23]. [23] provided other correlations that take more weather and Sun position
parameters into account. For intervals of k of 0-0.78, k was the largest predictor of Gdiffuse. However,
at high intervals (i.e. clear-sky conditions) as became the most relevant determinant since air mass
could largely increase the diffuse fraction of sunlight at low Sun altitudes. A more complete
correlation is therefore:

Gdiffuse  Gglobal (1.02  0.254  k  0.0123  sin as ) (0  k  0.3)


Gdiffuse  Gglobal (1.40  1.749  k  0.0177  sin as ) (0.3  k  0.78) (2.7)
Gdiffuse  Gglobal (0.486  k  0.182  sin as ) (0.78  k  1)

According to [23], equation (2.7) has a 9% lower composite residual sum squares compared to
equation (2.6), which is an indicator of the margin of error between the model and actual
measurements. This was for the complete dataset of that research, which included several locations
of varying latitude. Due to the higher proposed accuracy of the correlations in (2.7), equation (2.7)
was implemented in both the real-time and historical diffuse and direct component calculations of
the PVP 2.0, replacing step function (2.6) which was used in the PVP 1.0.

In the MATLAB model several additional checks were performed: if sun altitude becomes lower
than 0, air mass is set to zero as well to make sure a k of zero is arrived at (which indicates low
clearness). If k approaches zero, it is possible for G diffuse to exceed Gglobal in the step function. In this
case, Gdiffuse is set equal to Gglobal. Gdirect, the direct component of overall irradiance in the horizontal
plane is the complement of Gdiffuse and Gglobal.

2.3.5 WEATHER DATA UPLOAD TO DATABASE

After average data for all 46 weather stations were created, the data was uploaded to the
website MySQL database. Two additional fields were added to the database table next to the 10
weather parameters, in order to make data retrieval and analysis simpler. These are an identifier
field, with values ranging from 1-8784, and a datetime field, which contains the date and time of each
measurement row in a (YY-MM-DD HH:mm:ss) format. In the climate database, the year field is not
relevant as the measurement rows contain average values of multiple years, and therefore the year is
set to a standard value of 2000. The time range is set from 2000-01-01 00:30:00 (the middle of the
first hour timeslot (0:00-1:00)) to 2000-12-31 23:30:00 (the middle of the last hour timeslot of the
year (23:00-0:00)).

All time values are stored in UTC (Coordinated Universal Time), which has a 1- or 2- hour time
difference with the Netherlands (Central European (Summer) Time (CE(S)T)), dependent on whether
daylight saving time is active or not. UTC time values are used in all database tables to ensure that no
measurement time errors occur when the switch from winter to summer time and back is made, as
UTC times do not shift during daylight saving time. The UTC values are translated in PHP to CE(S)T
time values after retrieval from the database for correct display on the front-end of the website.

2.3.6 REAL-TIME WEATHER DATABASE AND CONNECTION TO CLIMATE DATABASE

As illustrated in Figure 5, a real-time weather database was required for the PVP 2.0. This real-
time weather database was created by designing 12 provincial and 46 individual station tables in
which to store real-time data. In addition, a script was written in the programming language PHP to

35
retrieve real-time KNMI measurements every 10 minutes. The real-time (10-min. average) weather
data are provided by the KNMI via an FTP connection. Every 10 minutes a CSV file on the FTP server of
the KNMI is populated with the most recent data of the weather stations. The data stored in this file is
retrieved from the FTP server and stored in a table in the PVP MySQL database. Subsequently, select
data can be fetched from the MySQL table and stored in PHP variables which are then uploaded to
the individual weather station tables via a MySQL query.

A distinction between the individual station and provincial tables must be made. The individual
station tables mostly rely on the measurement values stored in the KNMI table, whereas an averaging
of individual weather station measurements is required before values can be uploaded to the
provincial tables.

For the individual tables, the weather station codes are used to select the available parameters in
the KNMI table and upload the values to the relevant tables. However, only 20 of the 46 stations
measure all seven parameters necessary for the PVP 2.0 model. This was the main obstacle to
increase the spatial granularity of the website and the reason why the PVP 1.0 used provincial
averages instead of individual station values. In the PVP 2.0, a workaround was implemented to be
able to use all of the available measurements. In this approach, for each individual station, the
measured weather values are used where present. If one of the parameters is not measured at the
station, the provincial average value for that parameter at the local time is retrieved and used. In the
eventuality of absence of provincial average values, for example due to a server failure, the
parameter value of the individual station in the previous 10-minute period is inserted in the individual
table. In this approach, provincial averages are assumed to be somewhat representative for the
individual station. This way, all individual station measurements are combined with provincial
averages, allowing for database tables without any data gaps. Another issue concerns the speed of
measurement data transfer from the individual stations to the central KNMI server. There are 9
weather stations that lag 30 minutes in measurement data upload behind the other 37 stations. This
30-minute lag is recorded in the PVP 2.0 weather database by uploading the measurement time
denoted by the KNMI together with the parameter values. The latest data stored in the 9 lagging
station tables will be 30 minutes behind the latest data row of the other 37 most rapidly updated
stations.

For each of the 12 provinces, the averages of all provincial weather station parameter values are
retrieved (see Table 1 for an overview). To construct the provincial average, only the measurements
for which the measurement time is equal to the most recent measurement time are taken into
account. This is essential because of the lag of 9 of the real-time stations. Mixing measurements that
occurred at different times during the day would naturally result in an error in the calculated average
values (think of lower irradiance or temperature values at earlier times in the morning). In the
eventuality that no average data can be found for a parameter (for example because of an error in
the KNMI data retrieval from the weather stations), the last parameter value uploaded to the
province weather table is assigned to the current weather parameter. This ensures that the province
table in MySQL does not contain any data gaps.

After all weather parameters have been retrieved from the KNMI MySQL table, the values for the
diffuse and direct fraction of irradiance, and for the sun position (elevation and azimuth) are
calculated via the method described in the Additional parameter calculation section. The most recent
measurement time is used as an input for the sun position calculation, together with the average
station or province latitude and longitude coordinates. The measurement date and time in UTC,
together with the seven measured weather parameters and the four calculated weather parameters

36
are subsequently uploaded to the MySQL, yielding a complete weather measurement overview for
the Netherlands.

The real-time provincial averages do not include the measurements of the 9 delayed weather
stations. In order to include the delayed measurements in the final real-time province averages to be
uploaded into the PVP 2.0 climate database, the province values need to be averaged with the
delayed measurements when these become available 30 minutes after the real-time data. The
process is shown in Figure 9. 30 minutes after the initial real-time province average has been
calculated, a new, final, weather average is calculated, which is subsequently used to update the
values stored in the climate database.

Figure 9: Inclusion of the delayed measurements in the final provincial weather averages at a given
timestamp. The preliminary average is used as input in the real-time power model, as it is the most
recent value available.

In addition to values of the weather parameters, it is also of relevance for the PVP 2.0 climate
database how many weather stations contributed to the average weather values for a province at a
given measurement time. If for example 7 weather stations exist in Noord-Holland (which is the case),
but only 2 are equipped to measure cloud coverage in near-real-time, the average cloud coverage will
be less representative for the total province than for example the average of wind speed, which can
be measured in 5 weather stations. The number of stations that can measure a weather parameter is
important in achieving the dynamic update of the climate database. As explained in section 2.3.3, the
number of stations contributing to an average measurement is the weight that is assigned to this
measurement (referred to as the ‘count’ of the measurement). The measurement values and counts
for each 10-minute period in an hour can be averaged to yield values for the one-hour timeslot for the
specific month and day of the measurement. This is an equivalent format to the values that are stored
in the climate database. The provincial parameter values and counts include the delayed real-time
measurements. Naturally, all parameter measurements for the individual stations are assigned a
count of 1. The annual database values can then be averaged with the real-time values via the
following formulas:

Valuereal  Countreal   Valuehistory  Counthistory 


Valuehistory ,new  (2.8)
Countreal  Counthistory

37
Counthistory ,new  Countreal  Counthistory (2.9)

Equations (2.8) and (2.9) constitute the essence of the connection between the weather and
climate database as shown in Figure 5. The implementation of this feature in the PVP 2.0 has had two
consequences: the weather measurements are stored permanently through their contribution to the
climate average, and the climate database, originally constructed with historical KNMI data as
described in section 2.3.3, has become dynamic. The database tables will adjust to gradual changes in
climate over time, improving in accuracy as more measurements are stored. The script to dynamically
update the climate database with weather data can be found in Appendix 9.1.1.

2.4 RESULTS

In the results, the final climate datasets of the PVP 2.0 created through the methodology
described in the previous section are compared to the information on the Dutch climate made
available by the KNMI to check whether the results are valid, and the PVP 2.0 climate dataset is
compared to the TMY dataset of Meteonorm, and a TMY dataset generated from the KNMI data for
this thesis, to highlight the differences. An additional section evaluates the values of the calculated
parameters. Finally, the integration of these datasets in the front-end of the website is discussed.

2.4.1 CLIMATE DATASET VALIDATION

The first evaluation that must be made is to check whether the PVP 2.0 climate datasets match
the ‘official’ KNMI climate averages. The KNMI has published monthly climate data for the period of
1981-2010 for a select number of weather stations [21]. As a sample, the irradiation and ambient
temperature are taken as parameters for comparison. The results can be viewed in Figure 10-Figure
12. It can be seen that the PVP 2.0 data match the KNMI averages and that the PVP 2.0 datasets
compare well to the official KNMI estimates despite the different time periods under evaluation
(1981-2010 for the KNMI, 1991-2017 for the PVP 2.0).

38
Figure 10: Annual sum of global horizontal irradiance for all 46 climate datasets of the PVP 2.0. The
position in the figure represents the station’s geographical coordinates.

39
Figure 11: Annual sum of global horizontal irradiance over the period of 1981-2010 as calculated by
2 2
the KNMI. The units are in kJ/cm instead of kWh/m . The figure is taken from [21].

Both Figure 10 and Figure 11 show that the west of the Netherlands receives more irradiation
than the east. Figure 12 shows that the PVP 2.0 dataset for ambient temperature oscillates around
the monthly KNMI averages but that it still contains significant variability between hours.

40
Figure 12: Comparison of the PVP 2.0 climate average for ambient temperature and the monthly
temperature averages for 1981-2010 of the KNMI.

To highlight the differences between the results of the KNMI climate averaging and the
Meteonorm values, a comparison was made between data of Meteonorm and the climate KNMI data
results of the PVP 2.0 for the weather station in Eelde. As an additional comparison option, the KNMI
weather measurements for the year of 2006 were selected as a representative KNMI TMY to ensure
that the PVP 2.0 climate data could be compared with data from the same source as well. Eelde was
selected as the weather station has measured all relevant weather parameters from 1/1/1991
onward. As can be seen in Figure 13 and Figure 14, the Meteonorm data follows the climate trend of
the PVP 2.0, but contains more variable daily irradiation and temperatures. This is because the
climate averaging smooths fluctuations while the stochastic Meteonorm model uses average monthly
values but generates hourly values that mimic the normal weather fluctuations in any given year.

41
Figure 13: Comparison between the daily sum of global horizontal irradiation values of Meteonorm
and the PVP 2.0 climate database.

Figure 14: Comparison between the hourly ambient temperature values of Meteonorm and the
PVP 2.0 database.

42
2.4.2 CALCULATED PARAMETER EVALUATION

An additional check that is required is the evaluation of the calculated parameters, which are Sun
elevation and azimuth as well as the direct and diffuse horizontal irradiance. The Sun position
calculation was checked with the online position calculator of the U.S. governmental National Oceanic
1
& Atmospheric Administration (NOAA ) and the PVP 2.0 results were found to closely match the
NOAA calculations.

In a first calculation effort for the diffuse and direct irradiation, the annual sum of the diffuse
fraction in the PVP 2.0 dataset was found to be much higher than for the TMY dataset of Meteonorm
or for an annual weather dataset of the KNMI. The PVP 2.0 calculations found an annual diffuse
irradiation sum of >70% of global irradiation, whereas it was ~50% for the Meteonorm dataset. A
thorough evaluation of the PVP 2.0 calculation method led to the conclusion that the error occurred
because the Reindl decomposition method was applied to the climate average of the GHI. This
method is referred to as ‘averaging before decomposition’ (ABD). Instead, the correct approach is to
apply the decomposition method to the raw GHI measurements, yielding a list of diffuse and direct
hourly fractions for every hour between 1991 and 2017, and to subsequently average these direct and
diffuse arrays. This method is referred to as ‘decomposition before averaging’ (DBA). The difference
between these methods is visible in Figure 15.

Figure 15: Contribution of diffuse and direct horizontal irradiance to GHI in the PVP 2.0 climate
dataset. It is clear that the ABD method results in much higher diffuse fractions than the DBA
method.

The exact reason these two methods yield such different results can be attributed to the Reindl
decomposition method to calculate the diffuse irradiance from global irradiance. For each hour, the

1
The NOAA calculator can be accessed at https://www.esrl.noaa.gov/gmd/grad/solcalc/

43
clearness index kt is calculated, based on the ratio between the GHI and the expected maximum
sunlight (which is dependent on the Sun elevation, the solar constant and the eccentricity of the Sun
path). A high kt indicates that there is a clearer sky and therefore a high direct fraction of sunlight at
that time. The applied Reindl model is a step function with three levels, based on the value of k t. If the
GHI is averaged first, every day will have a GHI which matches that of a low clearness index. This
means the step function level is applied which almost always yields a high diffuse fraction. Instead, if
weather data is used, there will be a significant number of hours with a high GHI, leading to a high
clearness index and the higher occurrence of significant direct fractions.

When the decomposition before averaging method was applied, the PVP 2.0 climate diffuse and
direct fractions closely matched that of Meteonorm and of a TMY of the KNMI, as shown in Figure 16.

Figure 16: Annual sum of diffuse and direct irradiation for the Meteonorm TMY, the KNMI TMY, the
ABD PVP 2.0 dataset (referred to as original climate), and the DBA dataset (referred to as adjusted
climate).

A further evaluation of the Reindl decomposition method was made by applying it to every
measurement between 1991-2017 for Eelde Looking at the annual sum of global, diffuse, and direct
horizontal radiation displayed in Figure 17, it is noticeable that there is an inter-annual variation of
total irradiation, and that the Reindl decomposition model attributes this variation predominantly to
variation in the available direct horizontal irradiation, while the annual diffuse irradiation remains
fairly constant.

44
Figure 17: Total annual horizontal irradiation for the KNMI weather station of Eelde.

On visual inspection a light increasing trend seems to occur for annual GHI at the Eelde weather
2
station. However, basic fitting with a linear equation yielded a ~3 kWh/m increase per year, but with
2
a low R value, indicating that the observed trend is very weak. For irradiation, there does not seem to
be a long-term trend, but this may be different for temperature or rainfall, which could increase for
the Netherlands in the future due to climate change.

It is clear that the annual sum of the PVP 2.0 irradiation data is in line with the KNMI and
Meteonorm. However, it is also necessary to compare the diffuse and direct irradiance values of every
individual hour within these annual datasets. As expected, and as visualized in Figure 18, the climate
dataset contains a much lower variability of diffuse and direct fractions than the weather datasets.
While absent in the weather data, a clear positive correlation between diffuse and direct irradiance is
visible in the PVP 2.0 climate dataset. This is logical, as days close to 21 June will have a higher GHI
and therefore also a higher diffuse and direct irradiance.

45
Figure 18: Direct and diffuse irradiance for every hour in the annual KNMI TMY and climate dataset
for Eelde. The Meteonorm TMY (which is not shown in this figure) showed a similar variability in
diffuse and direct irradiance.

2.4.3 INTEGRATION IN WEBSITE

To be able to use the generated climate datasets in the online PVP 2.0 environment, the datasets
were exported from Matlab and uploaded to the MySQL climate database where they are stored in
tables. As mentioned before, the database system of the PVP 2.0 consists of two interconnected
databases. The weather database contains 70 tables storing real-time data for a period of 2 days. Of
these tables, 46 tables are dedicated to the individual weather stations, 12 to the province parameter
averages and 12 to the count values of the province parameter values. The climate database contains
116 tables, with 58 dedicated to the parameter values of the 46 weather stations and 12 provinces,
and 58 to the count values for each measurement average. The total climate database size is 115 MB,
taking up 23% of the total website storage space. This justifies the decision to adhere to a one-hour
time resolution as a 10-minute resolution would exceed the website storage significantly.

Real-time measurements of the KNMI are uploaded to the real-time database tables every 10
minutes. Every hour the 10-minute data of the past hour are retrieved with a PHP script, and
averaged with the historical data. The new weighed averages and counts are then uploaded to the
climate database. After becoming operational in May 2017, the two databases have functioned
without any issues and continue to store and update the weather and climate information.

As explained in the introduction, the primary use of the weather and climate databases is for
input in the PV system performance model. An additional functionality was implemented in the PVP
2.0 website to view the weather and climate data of all 46 KNMI stations and 12 provinces in
interactive charts in the browser, or to download the data in a comma-separated value (CSV) file,
which can be input into Excel, Matlab or any other modelling program of choice. Making the

46
information in the PVP 2.0 database publically available is in line with the goal of the Dutch PV Portal
to be an online resource that provides public access to useful information on solar energy. The
purpose of this view and download option is clear. The viewing option provides visitors with an
instantaneous grasp of the climate characteristics of the Netherlands. The download option is of great
use to any individual who wishes to use this data for solar energy modelling purposes. A number of
students in the PVMD group have already requested data from the PVP 2.0, and the data is already
being used in some Master theses within the PVMD group. Figure 19 displays a screenshot of the
page in which meteorological data can be viewed.

Figure 19: Screenshot of the climate data viewer in the PVP 2.0 is displayed. At the top of the page
other climate parameters can be selected.

2.5 DISCUSSION

One of the objectives for creating a new meteorological database for the PVP 2.0 was to improve
the spatial granularity over the PVP 1.0, in which the data was only available on a provincial level. By
switching to storing the data of 46 individual stations, the PVP 2.0 database has achieved a large
improvement in spatial granularity. A side note must be made: as shown in Appendix 9.2, in 26 of the
46 stations one or more parameters are not measured, which decreases the actual granularity of the
measurements stored in the database. The hope is that the KNMI will employ these 26 stations with
additional measurement equipment to record these parameters. If this will take place in the future,
these new measurements will automatically be included in the PVP 2.0 database. In other words, any
future expansion of the measurement capabilities of KNMI weather stations will result in an
automatic improvement of the spatial granularity of the PVP 2.0 database.

Through a visual check of the resulting climatological values and a comparison with climate
averages published by the KNMI, it was concluded that the PVP 2.0 climate database calculation
yielded realistic and consistent weather values for all weather stations. This enforced the confidence
in the correct method of averaging and thus the values stored in the database.

47
Several suggestions for future work on the PVP database can be made. Firstly, it may be possible
to implement an interpolation method in a new version of the PVP. In the PVP 2.0 website, weather
data of the KNMI station closest to the location input by a website user is selected, a simple method
of approximating the local weather conditions. Websites like PVGIS and Meteonorm apply
interpolation methods to calculate intermediate weather values for a location by averaging data of
multiple weather stations surrounding the location for a more accurate estimate. The user could be
presented with a choice between the data of the closest weather station or an interpolation of the
data of the surrounding weather stations in a future version of the Dutch PV Portal. A second
suggestion is an evaluation of the Reindl decomposition method. The Reindl et al. [23] approach to
diffuse fraction calculation has been criticized in some recent publications. Ridley et al. [28] found
that at for the most part, the Reindl model provided a reasonable prediction of the diffuse fraction
when compared with measurements. However, at high clearness indexes, the diffuse fraction was
overestimated. The logistic model proposed by [28] takes clearness index, apparent solar time, solar
altitude and correction factors for instantaneous clearness index as input. The logistic model was
shown to provide a better fit for the measured data at multiple locations around the world [28]. Due
to time constraints, no attempt was made to replace the Reindl model by the logistic model. In a new
version of the PVP it would be worth the effort to implement an updated diffuse fraction calculation
based on the research of Ridley et al..

It was concluded that the most significant difference between the datasets provided by
Meteonorm and the climate data stored in the PVP 2.0 database is the absence of large parameter
fluctuations in the PVP 2.0 climate database. It cannot be determined in advance whether the PVP 2.0
climate datasets will provide significantly different results compared to the conventional modelling
dataset when used as input in a PV system performance model. The slight departure of a climate
dataset from ‘real’ weather occurrences as a downside may be compensated or outweighed by the
advantages of its dynamic nature and its representation of the ‘true’ mean weather values for the
period under consideration. Both the Meteonorm and PVP 2.0 Climate datasets will be used in the
annual energy calculation model of the PVP 2.0, to assess whether a significant difference in PV
energy production between the two is found. The results are treated in Chapter 3.

2.6 CONCLUSION

In the start of the thesis, the following thesis objective was mentioned: the creation of a
meteorological database from which data can be retrieved that is applicable to any location within the
Netherlands, i.e. with a high geographical resolution. It can be stated with confidence that this
objective has been achieved, as the meteorological data of 46 weather stations and 12 provincial
averages are stored in the PVP 2.0. An evaluation of this data showed that all datasets together
contain realistic and representative values for the whole of the Netherlands, which is an excellent
result.

In addition, this chapter has presented a unique and innovative feature of the PVP 2.0 database:
the climate datasets are updated every hour with real-time weather measurements, making the
database dynamic instead of static. This distinguishes it from the other meteorological databases
which have static datasets that need to be updated manually every number of years. The PVP 2.0
database does this automatically which has as an additional advantage that it can incorporate the
effect of climate change on the meteorological conditions in the Netherlands.

The final important conclusion to be made is that this chapter has proposed a new, alternative
type of annual dataset for PV system performance modelling. The standard practice of solar radiation
databases and PV system modelling research is to use a typical meteorological year dataset that

48
consists of a selection of weather measurements that together match the average climate at a
location. The PVP 2.0 instead has created annual climate datasets created by averaging all weather
measurements from 1991-2017. The hypothesis is that these climate datasets will provide a valid
estimate of the PV system performance in a location despite the lower variability of the
meteorological data throughout the year. This hypothesis will be evaluated Chapter 3.

49
3 PV SYSTEM PERFORMANCE
MODEL
3.1 INTRODUCTION

In the previous chapter the creation of the meteorological database for PVP 2.0 was described.
The main use of the values stored in this database is to provide input for the estimation of the real-
time and annual power and energy production of PV systems. As stated in the introduction, the
second objective of this thesis is the creation of a model to calculate the annual energy output of a PV
system, based on meteorological data and a user-input PV system design.

In the PVP 1.0 a comprehensive PV system performance model was developed which used real-
time weather data together with user-input values customizing the design of a modeled PV system.
The PVP 1.0 model provided the PV system power output at the moment of calculation, but, in the
absence of annual meteorological datasets, no annual performance estimation could be made. With
the PVP 2.0 climate and weather database in place, the following step is to adjust the PVP 1.0 model
such that an annual PV system performance estimate is possible.

This chapter will describe the PVP 2.0 PV system performance model and how it builds upon the
PVP 1.0 model. The chapter will discuss both the adjustments and the additions made to the PVP 1.0
model and will therefore contain a critical assessment of the methodology employed in the PVP 1.0.

In the conclusion of the previous chapter, the hypothesis was posed that the PVP 2.0 climate
datasets would be a valid input for a PV system performance. In this chapter, this hypothesis will be
put to the test by comparing the effects of the PVP 2.0 climate dataset, the KNMI TMY dataset, and
the Meteonorm TMY dataset on the performance of a modeled PV system.

The chapter structure is as follows. The next section will show a schematic of the PVP 2.0 PV
system performance model which is based on the PVP 1.0 model. After this, the inputs necessary for
the model calculation are described including the options from which the website visitor can choose.
Subsequently, the calculation methodology behind each model component is discussed. The results
section will then evaluate the model as a whole and highlight several key components in further
detail. The results section will also contain a comparison of the meteorological datasets as model
inputs. In the discussion several important suggestions for future expansion and improvement of the
model are given. The conclusion will also give a brief prospect of the application of the PV system
performance model in the realization of the final three thesis objectives.

3.2 MODEL DESCRIPTION

The applied PV system performance calculation consists of two main steps. The first is the
determination of the irradiance incident on the PV array, also referred to as the plane-of-array
irradiance (GPoA). The second step is, given the incident irradiance, to model the practical operating
efficiency of the PV system (η PV), i.e. the percentage of GPoA that can be converted to electricity. When
both the values are known, the system output power PPV can be easily calculated as follows:

PPV  APV PV  GPoA (3.1)

50
Where APV denotes the total surface area of the solar panels in the PV system [1, Ch. 20]. The PVP
2.0 model therefore needs to determine the values of GPoA and ηPV for every time step in the annual
dataset to arrive at an annual energy production value for the PV system.

To calculate GPoA, the meteorological information on global, direct, and diffuse horizontal
irradiance needs to be input into a tilted surface radiation model, which translates the intensities of
the sunlight in a horizontal plane to the plane-of-array of the module [29]. This will result in a value
for the primary plane-of-array irradiance (GPoA,prim) which is the light intensity excluding any light
capture losses that may occur before the sunlight energy enters the panel. The main light capture
losses considered in the PVP 2.0 are panel shading by surrounding objects, light absorption and
reflection by dust accumulated on the panel (referred to as soiling losses), and angular reflection
losses due to a non-perpendicular angle of the sunlight to the tilted plane of the solar panel [1, Ch.
20]. When these loss factors are expressed as an efficiency term, GPoA can then be determined via
equation (3.2):

GPoA  GPoA ,prim shading soiling reflectance (3.2)

Similarly, ηPV can also be determined via a so-called 'chain of efficiencies' that quantify the energy
loss occurring at every step between the incidence of GPoA on the panel and the output of AC
electricity by the solar inverter. The starting point of this efficiency chain is the panel efficiency under
standard test conditions (STC). The term STC refers to the conditions under which a panel
manufacturer determines the conversion efficiency of a solar panel, which are a module temperature
2
of 25 °C, a GPoA of 1000 W/m , and a light spectrum comparable to the AM1.5 spectrum of sunlight [1,
Ch. 5]. STC efficiency (ηSTC) is considered as a benchmark efficiency. Module efficiencies in real life are
usually lower because of the adverse effect of high module temperatures under illumination on the
panel efficiency. Therefore, the efficiency loss due to non-STC irradiance (i.e. differing from 1000
2
W/m ) and non-STC temperature (i.e. differing from 25 °C) must be taken into account in the PVP 2.0
model via the efficiency factors ηirr and ηT. Additional efficiency losses in the PV system after the
module conversion are resistive (Ohmic) losses in cabling (η cable), a mismatch between power
production of interconnected modules leading to energy dissipation (η MM), and losses in the
conversion from direct current to alternating current by the solar inverter (η inv), which usually
includes the efficiency loss of powering the maximum power point tracking (MPPT) algorithm to
optimize PV panel output (ηMPPT) [1, Ch. 20]. The chain of efficiencies for ηPV is summarized in
equation (3.3):

PV  STC irr T cable MM MPPT inv (3.3)

Figure 20 displays a detailed overview of the PVP 2.0 model to calculate the PV system
performance. Although this model may seem overwhelming at first, each step will be discussed in
detail in the next subchapter. The black texts with orange arrows describe calculation inputs, purple
boxes refer to the steps required to determine GPoA, blue boxes refer to steps to determine the η PV,
and the green box denotes the final result of the model, i.e. the system's AC power output for that
time step.

51
Figure 20: Overview of the PVP 2.0 PV system performance model, which will be discussed in detail
in the rest of this chapter.

3.3 MODEL INPUT PARAMETERS

Figure 20 showed a set of inputs required to determine the PV system performance. Some of
these inputs need to be defined manually in order to run the code. In the PVP 2.0, like the PVP 1.0, a
front design page is present in which users can specify the model inputs. In the PVP 2.0, boundaries
on the input values are set in the HTML code which ensures that the user cannot (purposely or
unconsciously) misuse the model. The PV system design options are summarized in Table 4.

Table 4: PV system design inputs in the PVP 2.0.


Input Description Limit/choices
Location The installation location Within the Netherlands.
Module technology The type of solar panel. Seven crystalline silicon and
thin-film technologies.
System type The installation type. Building-added PV (BAPV) or
free-standing (FS)
System size The desired system size. Installation capacity [Wp],
2
installation area [m ], number
of modules, or area coverage of
province [%]
Tilt The system tilt w.r.t the 0-90°
horizontal plane (0°)
Azimuth Module orientation w.r.t. north 0-359°
(0°)

The previous chapter showed that the PVP 2.0 database stores data for 46 weather stations in
the Netherlands. For any location chosen by the user the data of the closest weather station is
retrieved from the meteorological database and used as input in the model.

52
In the PVP 1.0, six different types of solar panel technologies were available as options:
monocrystalline silicon (mono-Si), polycrystalline silicon (poly-Si), cadmium telluride (CdTe), CIGS,
silicon-based rigid thin-film (rigid a-Si) and silicon-based flexible thin-film (flex a-Si). In the PVP 2.0, the
rigid a-Si panel was removed as it was discontinued by the manufacturer. Instead, a heterojunction
silicon panel was included, which is another technology of significance in the solar market. In Table 5
an overview of the PVP 2.0 modules can be found.

Table 5: Overview of the solar panels used in the PVP 2.0.


Technology Name Efficiency at Peak Short-circuit Open-circuit
STC [%] power current [A] voltage [V]
[Wp]
High-efficiency SunPower SPR- 20.1 327 6.46 64.9
mono-Si E20-327
Medium-efficiency Trina Solar TSM- 18.3 300 9.64 39.9
mono-Si 300DD05A.08
Poly-Si Trina Solar TSM- 16.5 270 9.18 38.4
270PD05
Heterojunction Panasonic HIT 19.8 331 6.07 69.7
N330
CdTe First Solar FS-4122- 17.0 122 1.85 88.7
3
CIGS Solar Frontier 14.3 175 2.20 114
SF175-S
Flex a-Si HyET Solar 8.6 165 6.5 38.1
Powerfoil 165

The PVP 2.0 code makes a division is made into two types of systems: building added
photovoltaics (BAPV), a.k.a. rooftop systems, and free-standing, a.k.a. field PV systems. This division
reflects a system type division made by the PVMD group between urban-integrated PV and
environment-integrated PV. BAPV systems are usually smaller than free-standing PV systems. The
main reason for this division is to incorporate the different environments in which these systems are
placed (i.e. the city vs. the countryside).

The user can specify how they want to determine the size of the system. For the BAPV, the
desired installation capacity, available roof area, or requested number of modules can be input. This
is the case as well for the free-standing systems, with an additional option that the percentage
coverage of the selected province can be input to simulate the large-scale adoption of PV.

In the PVP 1.0, the azimuth for field systems was always set at 36° tilt and 180° azimuth. This is an
optimal orientation. The PVP 2.0 removed this restriction, allowing users to choose any tilt and
azimuth they want.

3.4 IN-PLANE IRRADIANCE

3.4.1 TILTED SURFACE MODEL

The goal of a tilted surface irradiance model is to calculate the total irradiance falling on a PV
module, a.k.a. the irradiance in plane-of-array (GPoA). The plane-of-array is the geographic plane in
which the panel surface is positioned, described by its tilt angle from the ground and its azimuth
relative to north. A tilted surface model calculates the values of the three main components of G PoA:
direct (a.k.a. beam), diffuse and ground-reflected (a.k.a. albedo) irradiance:

53
GPoA  Gbeam  GdiffusePoA  Galbedo (3.4)

The model takes several parameters as input, most importantly the Sun position, the PV module
orientation and the values of the horizontal irradiance components, as calculated in the Reindl
decomposition model described in Chapter 2.

The PVP 1.0 applied the Reindl, Beckman and Duffie [23] tilted surface radiation model to
calculate the beam, diffuse-PoA and albedo fractions. As described in [6], the main motivation to opt
for the Reindl et al. tilted radiation model was to keep it in line with the diffuse fraction correlation
model also authored by Reindl et al. [23]. However, it may be possible that alternative tilted radiation
models have been found to provide better modeling results when compared to in-plane radiation
measurements. Therefore it is beneficial to shortly look at model comparisons in literature.

A comparison of seven tilted radiation models by Loutzenhiser et al. [30] for Switzerland found
the mean deviation from measurements of the Reindl et al. model to be slightly higher than that of
the Muneer model (accounting for the difference between shaded and unshaded surfaces), the Hay &
Davies model which excludes horizontal brightening, and the Perez et al. model, which, like the Reindl
et al. model accounts for isotropic, circumsolar and horizontal brightening components while
additionally considering the sky clearness and brightness. Other comparative research by Padovan
and Del Col [31] between four models for Padova, Italy in 2007-2009, found the Reindl tilted surface
model to perform far better than isotropic models and to even outperform the Perez model. A
comparison between 12 tilted surface models by Noorian et al. [29] for Karaj, Iran found the Reindl et
al. model the best predictor of PoA irradiance for south-facing surfaces, while the Perez model slightly
outperformed Reindl for a western orientation. The comparative research suggests that both Reindl
and Perez are appropriate tilted surface models. As no comparative research was found for the
Netherlands, it is difficult to determine which tilted surface model would be best applied in the
country. Results from literature clearly varied between different locations. Although Reindl does not
always perform the best, it is continuously cited as a good predictor of PoA irradiance. After re-
evaluation the decision was therefore made to continue using the Reindl et al. tilted surface model in
the PVP 2.0.

The Reindl et al. model contains a straightforward calculation of beam and albedo irradiance:

Gdirect
Gbeam  DNI  cos(AOI)   cos(AOI) (3.5)
cos( z )

Where DNI refers to the direct normal irradiance, which is the intensity of the direct component
of sunlight in a plane perpendicular to the sunlight rays, and AOI is the angle of incidence which is the
angle between the Sun's position in the sky and a vector pointing out of the plane in which the panel
is oriented [23]. Figure 21 provides a visual representation of the two translations (from G direct to DNI
and from DNI to Gbeam) which are both dependent on the Sun position.

54
Figure 21: Relation between the irradiance vectors G direct, DNI and Gbeam via the module tilt θM,
angle of incidence θAOI and zenith angle θz.

The PVP 1.0 accidentally took Gdirect to be equivalent to DNI in its calculations, which led to a
significant underestimation of the beam irradiance intensity by a factor of cos θ z. This error has been
corrected in the PVP 2.0.

Albedo refers to the percentage of sunlight that is reflected by the surroundings of a PV system.
The albedo component of the sunlight is calculated with equation (3.6):

1  cosM
Galbedo  GHI    (3.6)
2
Where GHI is the global horizontal irradiance, ρ is the ground reflectivity which is also referred to
as the albedo, and θM is the module tilt. For BAPV and free-standing PV systems, different albedo
values should be used. This is as BAPV panels are placed on rooftops in the built environment,
whereas free-standing PV systems will be placed in fields. For the city of Rotterdam, neighborhood
albedo ranges of 0.06-0.16 were found [32]. An overview of urban albedo values for cities around the
world by Santamouris ranged from 0.10-0.20, indicating that a standard albedo value of 0.15 can be
used as an approximation [33]. For grass, the albedo was found to be 0.24 [33]. As building-added PV
systems are mostly employed in cities and towns, whereas free-standing PV plants in the Netherlands
will mostly be placed in grass fields, the model differentiates between the BAPV and free-standing
system types by applying the appropriate albedo value.

The diffuse radiation incident on the plane-of-array of the module is dependent on the fractions
of circumsolar irradiance (from the direction of the Sun), horizontal brightening (an increase in diffuse

55
light at the horizon) and isotropic diffuse irradiance (from all directions of the sky) [6]. The Reindl et
al. tilted surface model applies the following formula for calculation of the diffuse component:

  1  cosM   Gdirect 3  M  Gbeam 


GdiffusePoA  Gdiffuse   1  Ai       1  GHI  sin  2    Ai  G  (3.7)
  2     
 direct 

In equation (3.7), Ai refers to the anisotropy index, the ratio of DNI to the normal irradiance at
the top of the atmosphere. The similarity to the clearness index k used in the Reindl decomposition
model can clearly be noted in equation (3.8).

Gbeam
Ai 
S0  E0 (3.8)

2
Where S0 is the solar constant of 1361 W/m and E0 is the eccentricity of the Earth's orbit around
the Sun. The higher Ai is, the stronger the contribution of the circumsolar component to diffuse
irradiance in the plane-of-array.

By summing together the direct, albedo and diffuse PoA components, the G PoA,primary can be
established. To find the effective GPoA, light capture losses must be quantified as well.

3.4.2 LIGHT CAPTURE LOSSES

3.4.2.1 SHADING

The degree of shading on a PV system is location specific and depends on the presence of
obstacles blocking sunlight from falling on the panel. Multiple options for shading analysis exist, but in
general detailed information on the surroundings of the PV systems is needed. This can either be
achieved by an on-site visit or by using a LiDAR dataset that contains height data for the Netherlands
2
at a 0.5x0.5 m resolution. Both options were discarded for this thesis as the website should provide a
fast energy calculation, in which there is no place for on-site visits or computationally-intensive LiDAR
data analysis.

As it is difficult for website visitors themselves to estimate shading, instead the decision was
made to implement a hard-coded, i.e. fixed shading loss factor that would be representative for the
average shading losses in BAPV systems (it was assumed that field systems are placed in locations
with little to no surrounding shading).

To find the shading loss factor, a dataset of energy losses due to shading for 6446 simulated BAPV
systems was used, as provided by Solar Monkey. Solar Monkey is a Delft-based modelling company
that writes software to design PV systems and estimate their energy yield in any location in the
Netherlands. Solar Monkey uses Lidar data together with a Sun position calculator to simulate the
amount of shading that will occur on a rooftop. Based on this analysis they can provide a shading
factor, which is the percentage of incident sunlight on a panel or system that is lost due to shading.
ηshading is therefore the inverse of the shading factor.

Some of the systems in the dataset had such high shading that they would never be designed in
real life. According to Solar Monkey, this could either be attributed to a large amount of surrounding
obstacles shading the PV system, or to gaps in the Lidar data that are interpreted as a 100% shading
factor on specific panels, thereby increasing the interpreted shading on the entire PV system. After

56
consultation with Solar Monkey, the decision was made to ignore all systems with a shading factor
above 25.7%. This shading factor was based on the Solar Monkey rule-of-thumb that if the expected
annual yield of systems is lower than 650 kWh/kWp, the system will most likely not be built. The
Dutch government uses 875 kWh/kWp as the average annual PV system energy yield in the
Netherlands [34]. A system with an energy yield of 650 kWh/kW p due to shading would require a
shading factor of 1-650/875= 25.7%. Based on this, 1048 systems were excluded from the analysis,
leaving 5398 systems. Figure 22 shows the distribution of the shading factor of the Solar Monkey
systems. The distribution has a positive skew, with most systems having a limited shading factor
below 5%.

Figure 22: Shading factor distribution of the 5398 simulated PV systems of Solar Monkey.

The average shading factor of systems below 25.7% shading is 6.88%. This average shading factor
was taken as a representative factor for BAPV systems in the PVP 2.0 model. The total G PoA for BAPV
was derated by this factor, as the average shading factor represents the annual energy loss that
occurs due to shading. The GPoA of free-standing systems was left untouched, assuming they are built
in fields without surrounding obstacles.

The model takes several quite simplifying assumptions. Most strikingly, the GHI in all hours is
derated and uniform shading of the entire module, instead of partial shading, is assumed. In reality,
shading will occur at hours in which the specific Sun position causes its light to be blocked by an
object. This shading will move over parts of the module as the Sun progresses through the sky and will
therefore cause only partial shading. Although the modelled physical processes are different between
the PVP 2.0 and what happens in reality, it is proposed that the final result will be the same: a shading
loss of 6.88% on module output as quantified by Solar Monkey.

57
3.4.2.2 SOILING

The second factor that influences the amount of in-plane-irradiance that actually enters the solar
panel is soiling. Soiling refers to the accumulation of dirt on the panel surface that can absorb and
reflect light which would otherwise have been used by the panel itself to produce electricity. Rule-of-
thumb estimates of soiling suggest that, on average, 1% of the G PoA is lost due to soiling in the
Netherlands [1, Ch. 20].

A model to more accurately calculate the soiling loss at a certain location without taking on-site
measurements of soiling on the panels itself uses the concepts of the rain-free period (RFP, the
duration between rainfall occurrences) and the soiling factor (SF) [35]. In this model, a daily efficiency
loss attributed to soiling (i.e. the SF) is multiplied with the number of days it has not rained (i.e. the
RFP). Ergo, on days on which no rainfall occurs, dust will be accumulated, thereby heightening the
efficiency loss due to soiling. On days on which rain does occur, the RFP is ‘reset’ to zero as the
accumulated dust is washed off the panel. According to research of panel soiling in climates with long
dry spells, rainfall of 4-5 mm would be required to completely clean a soiled panel, the so-called
'cleaning threshold' [35], [36]. However, in locations with more frequent rainfall, the buildup of dust
on the panel is less severe, and by effect the cleaning threshold is lower. For the Netherlands, a
threshold of 2 mm/day of rainfall was chosen, in accordance with the Master thesis research of
2
Pramod Nepal [37]. In experiments by Nepal, it was found that 2 liters of water on a 1 m surface area
was sufficient to completely clean a module of dust.

The SF, having a unit of %/day, itself is again dependent on a number of variables that influence
the deposition of dirt on the panel, which are the dust settling velocity (DSV) (in m/s) and the
3
particulate matter concentration (CPM) (in μg/m ). The multiplication of the dust settling velocity and
the particulate matter concentration leads to a factor with a unit of:

 m   g   g 
DF  DSV  C PM    .  3    2  (3.9)
 s  m  m s

This factor, which for convenience will be referred to as the deposition factor (DF), indicates how
much dust accumulates on the panel surface each second [37]. The DF then can be translated to the
SF. The Dutch National Institute for Public Health and the Environment (the RIVM) has published data
on the daily average PM10 concentration, i.e. the concentration of particulate mass with a diameter
below 10 μm, as this concentration is a proxy for air pollution affecting the health of citizens. To this
extent, good data is available to determine the CPM for multiple locations in the Netherlands.
However, complications arise for the DSV calculation and the translation of DF to the actual light
transmission loss represented by the SF. For this to occur accurately, the distribution of particle size
within the PM10 concentration is needed. To determine the DSV, a constant particle size should be
assumed, but in reality the particle size will vary (amongst others due to weather conditions) [38]. The
uncertainty regarding this data makes it difficult to calculate an SF without large simplifying
assumptions.

An alternate approach was chosen for implementation of a soiling factor in the PVP 2.0. Instead
of calculating the SF indirectly via measurements of PM10 and particle size distribution, again the
Master thesis research of Pramod Nepal was used [37]. In this research, a comparison was made
between a rooftop PV system that was cleaned daily, thus preventing dust from accumulating upon
the panels, and a PV system that was not cleaned. Both PV systems were located on the same rooftop
in Delft, the Netherlands. The energy production data were compared and a soiling factor of
0.083%/day was found.

58
With the SF for the Netherlands as a whole assumed equal to that Delft, the next step is to
determine the RFP. For this thesis, a similar method was employed as for determination of all other
climate parameters in Chapter 2. KNMI data on hourly rainfall were downloaded for the 46 Dutch
weather stations over the same period as all other weather data used in the database. For each day
between 1991 and 2017, the RFP of that day was calculated. If on the previous day the sum of rainfall
was less than 2 mm, the RFP of the following day was assumed equivalent to the RFP of the previous
day plus one day. If larger than 2 mm, the RFP value of the following day was set to zero as the panel
had been cleaned the day before. A climate-averaged RFP for every of the 366 days in the year was
calculated for the 46 weather stations and 12 Dutch provinces and uploaded to uploaded to tables in
the PVP 2.0 climate database. For every time step in the annual energy calculation, the RFP of the
date of that time step and that location is fetched from the climate database and multiplied with the
soiling loss factor to find ηsoiling:

soiling  1   SF  RFP  (3.10)

As the PVP 2.0 stores both weather and climate data, the PVP 2.0 also distinguishes between the
real-time and climatic value of RFP. Real-time rainfall measurements are used to calculate today's RFP
for every station and province. At midnight, two actions are performed: a weighted average of the
daily/weather RFP in the weather database with the RFP for that same day of the year is made, and
the daily RFP is adjusted based on the amount of rainfall that occurred during the day before
midnight, as shown in Figure 23.

Figure 23: Determination and long-term storage of the daily and climate rain-free period in the PVP
2.0 database. RT stands for real-time values (as opposed to climate values).

3.4.2.3 REFLECTANCE

The final light capture loss is angular reflectance of sunlight from the panel surface. Compared to
the STC performance, sunlight falls on the panel in non-normal angles and therefore the reflectance
from the panel surface is larger than at STC. It is estimated that for commercial modules about 4% of
GPoA is lost annually due to reflectance, which is the fixed reflectance loss applied in the PVP 2.0
model [1, Ch. 20].

3.5 PV SYSTEM CONVERSION EFFICIENCY

To calculate the real conversion efficiency of a PV system, the efficiency losses are separated into
three sections for the purpose of this subchapter: the conversion efficiency of sunlight to DC
electricity of the module itself, efficiency losses in the additional system components excluding the
inverter, and the conversion efficiency of the inverter.

59
3.5.1 MODULE CONVERSION EFFICIENCY

In the PVP 2.0 model, the DC conversion efficiency of the module is the efficiency under non-STC
irradiance and temperature. Before the effects of these two factors are quantified, it is first necessary
to calculate the module temperature. In the PVP 1.0, the thermal model developed by Fuentes in
1987 was applied to calculate the real-time module temperature given the local weather. The thermal
model is based on a heat balance between incoming in-plane irradiance and radiation and convection
from the module to the atmosphere. A full coverage of the thermal model can be found in [6, pp. 33–
47]. Here a brief overview is given of the adjustments made to the thermal model for this thesis.

Firstly, the on-site atmospheric pressure was included in the thermal model calculations, whereas
in the PVP 1.0 a standard value of 101149.6 Pa for pressure was used.

Secondly, in the PVP 1.0 model, sky temperature (Tsky) was calculated from ambient temperature
(Tamb) via the relation in equation (3.11) [6]. Sky temperature determines the amount of radiation
from the top of a module. Equation (3.11) holds true for clear-sky conditions (i.e. with little to no
cloud coverage) [7].

3
Tsky  0.0552  Tamb 2 (3.11)

For overcast conditions, i.e. a cloud coverage of 6 okta or higher, sky temperature can be
assumed equal to ambient temperature [7]. However, in the PVP 1.0 model, equation (3.11) was
applied for both cloudy and non-cloudy conditions. In the PVP 2.0, a linear relation between Tsky and
cloud coverage C is applied, as adapted from [7]:

1  1 
3
Tsky    C  Tamb      8  C   0.0552  Tamb 
2
(3.12)
8  8 

Here, C refers to the cloud coverage in okta, i.e. as fraction of the sky dome in eights. At 0 okta,
the sky is completely cloudless and (3.12) reduces to (3.11). For a completely overcast sky of 8 okta,
Tsky  Tamb holds true.

Thirdly, a mistake in the formula for the ratio of actual to ideal back side convection was
amended in which Tsky instead of Tamb was used. In the PVP 1.0, ideal back side convection was
assumed equal to front side convection, for which radiation depends on T sky. However, Fuentes
applied Tamb as the ideal environment temperature, since back side radiation occurs to the ground,
instead of the sky [7]. Therefore, an ideal situation would be a ground temperature which is equal to
ambient temperature, not sky temperature.

In the model, a distinction between BAPV and free-standing modules was made regarding
radiation from the module backside. In the PVP 1.0 model, backside convection was neglected due to
the close proximity of the module to the rooftop (a building-integrated (BIPV) instead of BAPV system
was modelled in the PVP 1.0 model). In the PVP 2.0, a BAPV module is modelled. For a BAPV module,
backside convection is not neglected but commonly is quite low. Throughout the year, the ratio of
backside to front side convection varies between 1 and 10% as calculated in the PVP 2.0 model. In the
model, Tground is taken as equivalent to Troof, which is calculated as being between module and
ambient temperature, lying closest to module temperature. Conversely, a free-standing PV system
has a significant distance between the module and the ground. Here, the ground temperature as
measured by the KNMI is used. Consequently, radiative heat transfer from the back of the module is

60
not neglected and is equivalent to  back    TM4  Tground
4
 where εback refers to the backside emissivity
[7].

Finally, a different calculation of wind speed at the height of the module was implemented.
Fuentes applied the power law to infer wind speed at the module from the measured wind speed at
reference height (for the KNMI this is 10 meters) [7]:


h 
U(hM )  U(href )   M 
h (3.13)
 ref 

In equation (3.13), hM is the module height, href the reference height, and α a surface roughness
factor which is assumed to be 1/5 for open landscapes. No differentiation between urban
environments and open landscapes were made in the PVP 1.0 model. In meteorological simulations,
the logarithmic wind profile law, instead of the power law, is the most applied relation to estimate
wind speeds [39]. The logarithmic wind profile is better applicable at low heights (<60 m above
ground level) than the power law [40]. The logarithmic wind profile law is defined as:

h 
ln M 
z
U(hM )  U(href )   0  (3.14)
h 
ln ref 
 z0 

z0 refers to the terrain roughness, i.e the average obstacle height in the landscape. For open
landscapes, in which free-standing modules would be placed, the terrain roughness is ~0.03 m. In
urban environments with a high number of buildings, z 0 is assumed to be 0.4 [40]. The value of
reference height href is 10 m, which is the height of all published real-time KNMI measurements for
wind speed. A module height of 1 m is assumed for rack-mounted free-standing modules which would
be placed in a field. For BAPV, a height of 7 m is assumed, based on the average height of a 2-storey
building in the Netherlands.

3.5.2 TEMPERATURE AND IRRADIANCE EFFECT ON DC MODULE EFFICIENCY

With the module temperature calculated by the fluid-dynamic model, and the GPoA already
established, it is possible to quantify the effect on module STC efficiency.

For temperature effects, the same method was used as in the PVP 1.0 [6]. This means that the
temperature coefficients stated by the manufacturers in their datasheets were used by multiplying
the temperature difference of the module from STC (in K, as calculated with the fluid-dynamic model)
with the temperature coefficient of power (in %/K).

For irradiance effects, in the PVP 1.0 a linear relation between short-circuit current and irradiance
was assumed, while the fill factor was assumed to be independent of irradiance. Equation (3.15) was
applied to calculate the open-circuit voltage (VOC) of a module under low irradiance conditions [6]:

ln(GPoA )
VOC (GPoA )  VOC (GSTC )  (3.15)
ln(GSTC )

61
Where VOC (GSTC) refers to the VOC at STC irradiance. Application of this formula in the PVP 2.0
model led to an annual energy loss of >17% due to low irradiance effects, which is unrealistically high.
Further evaluation of the literature cited for equation (3.15) in [6] resulted in the conclusion that the
equation was based on a misinterpretation of the research of [41]. In place of equation (3.15),
equation (3.16) was applied as suggested by [1, Ch. 20]:

n  kB  T  GPoA 
VOC (cell , GPoA )  VOC (cell , GSTC )  ln  (3.16)
q  GSTC 
The equation above relates the change in VOC of a single solar cell to diode ideality factor n, the
Boltzmann constant kB, the cell temperature T (as only irradiance effects are considered here the
value of T is assumed to be 298 K or 25 °C), the elementary charge q, and the fraction of G PoA to STC
irradiance. Equation (3.16) quantifies VOC dependence on irradiance for a single cell, but for a module
as a whole the equation must be multiplied with the number of cells in series to find the change in V OC
for the entire module, as shown in equation (3.17):

 n  kB  T  GPoA 
VOC (M , GPoA )  VOC (M , GSTC )  N   ln  
 q (3.17)
  GSTC 

Where VOC(M,GPoA) refers to the VOC at module level, and N is the number of solar cells connected
in series in the module.

The ideality factor is a factor used to describe the behavior of a solar cell in a single diode model
[1, Ch. 20]. The ideality factor commonly has a value between 1 and 2. The lower the value of the
ideality factor, the higher the quality of the solar cell and the lower its series resistance [1, Ch. 20]. To
identify the ideality factor of a specific solar cell, it is necessary to measure its characteristics. As
measuring the ideality factor of the commercial modules used in the PVP 2.0 was outside the scope of
this thesis, a consultation with PhDs in the PVMD group was made on this topic. They advised, in
absence of actual measurement data of the module cells themselves, to use ideality factors based on
the module material. For mono-Si panels, which commonly consist of high-quality cells, an ideality
factor of 1 was suggested, 1.5 for poly-Si, and 2 for the thin-film technologies of CdTe, CIGS, and a-Si
(Gianluca Limodio & Johan Blanker, personal communication, 26 February 2018).

The specific ideality factor selected for a solar module determines the sensitivity of its V OC to
irradiance changes. Higher ideality factors result in lower VOC values and therefore lower efficiencies
at low light conditions. This is shown in Figure 24, together with the erroneous VOC dependence on
irradiance of the PVP 1.0 model. Figure 25 (and Figure 26 in more detail) shows the irradiance
dependence of all seven PVP 2.0 modules in the PVP 2.0 model. Here it can be seen that a high
ideality factor does not necessarily result in a high relative sensitivity of VOC to irradiance. The a-Si
module for example has an ideality factor of 2, but because of the higher cell voltage compared to a
poly-Si cell (with an n of 1.5), the percentage change in VOC is actually lower. The modules least
susceptible to non-STC irradiance losses are the mono-Si, heterojunction Si, a-Si panels, while the
CdTe and poly-Si panel have a medium sensitivity. The CIGS panel with both a high ideality factor and
a relatively low cell VOC is most susceptible to low-light losses in the PVP 2.0 model.

62
Figure 24: Dependence of the module VOC on the ideality factor. The voltage dependence used in
the PVP 1.0 is displayed as well. The module characteristics of the Panasonic HIT module were used.

Figure 25: Relative sensitivity of the PVP 2.0 modules to low-light conditions. The top cluster
contains four modules, i.e. mono-Si, HIT and a-Si. The middle cluster is the CdTe and poly-Si, while
CIGS is most sensitive.

63
Figure 26: VOC dependence on irradiance as zoomed-in from the previous figure to distinguish the
different module lines.

3.5.3 ADDITIONAL COMPONENT DC LOSSES

After the efficiency under non-STC irradiance and temperature was calculated, an additional loss
factor of 2% was employed for module mismatch losses and for Ohmic losses in cabling. Module
mismatch losses occur when modules in a string produce different currents which leads to power
dissipation in the module string, while Ohmic cabling losses are due to the electrical resistance of the
cable wires Module mismatch losses are estimated at 1.5% and Ohmic cabling losses at 0.5% [1, Ch.
20]. Together with the temperature and irradiance effects this results in the DC-side efficiency of the
PV system.

3.5.4 DC TO AC CONVERSION EFFICIENCY

3.5.4.1 INVERTER EFFICIENCY MODEL

After the DC efficiency of the modules has been calculated, the DC to AC conversion efficiency is
calculated with the inverter model. The choice was made to continue the approach applied in the PVP
1.0, but to expand it with research performed by another master student, Jessica Hernández Castro
Barreto [42].

First the PVP 1.0 inverter method is briefly described. , which is an application of the Sandia
National Laboratory (SNL) inverter model that.

In the PVP 1.0, the Sandia National Laboratory (SNL) inverter model was applied. This model is
based upon experimental measurements of inverter efficiency. The SNL model found that the input

64
power (PDC) and input voltage (VDC) into the inverter, together with eight electronic characteristics of
the inverter, determine the inverter output power and thereby its efficiency [43]. The great
advantage of the SNL model is that there is a clear time dependence of inverter efficiency that also
occurs in reality, in contrast to a constant efficiency model which would be the most simplistic model
application. The equations used in the SNL model are as follows:

 P 
PAC   AC 0  C   A  B     PDC  B   C   PDC  B 
2

 AB 
A  PDC 0  1  C1 VDC  VDC 0   (3.18)
B  PS 0  1  C2 VDC  VDC 0  
C  C 0  1  C3 VDC  VDC 0  

In equation (3.18), PAC is the inverter output power, PAC0 the nominal inverter output power, PDC
the instantaneous inverter input power, PDC0 the nominal inverter input power, and PS0 is the nominal
self-consumption of the inverter (the amount of energy needed to power the process). C 0 to C3 are
curvature coefficients, determining the dependence of the nominal powers on the difference
between the instantaneous input voltage VDC and the output voltage VDC0 [43]. The SNL determined
the efficiency characteristics of a large database of commercial inverters via this method and has
made the values of the parameters described in equation (3.18) publically available. This information
has been incorporated in the PVP 2.0 database, as was done in the PVP 1.0, which means that after
selection of a suitable inverter the inverter efficiency can be calculated dynamically for every
modelling time step. The range of inverter input powers in this dataset lies between 185W and 1.73
MW.

3.5.4.2 INVERTER SIZING FACTOR

Before the efficiency of an inverter can be calculated, one must be selected first. In the PVP 1.0,
the following selection criteria for the inverter were taken:

 The nominal input power must be close to the nominal installed capacity of the module
string/system.
 The nominal DC voltage should be equal to the open-circuit voltage of the string/system.
The PVP 2.0 adheres to these same criteria, but with the additional step to reduce the required
size of the inverter depending on the azimuth and tilt of the system. The concept of an ISF was
developed by Jessica Hernández as part of her Master thesis at the PVMD group [42]. The research
performed by Jessica Hernandez showed that, during the vast majority of the time, the PV system
does not operate at its nominal output power due to lower-than-STC irradiance and/or higher-than-
STC module temperature. The Sandia model already shows that when the input power is lower than
20% of the nominal inverter power, the inverter efficiency drops significantly and part of the
generated DC power is lost [6]. Between 30-100% of the nominal power, inverter efficiency is
relatively constant and commonly very high. [42] found that for non-optimal module orientations and
tilts, it is beneficial to select an inverter with a nominal power that is lower than the installed capacity.
The loss in generation power due to peak shaving (i.e. when the generated DC power exceeds the
nominal inverter power and the PV system operating point is moved away from the MPP) is more
than offset by the increased inverter efficiencies at lower irradiances, since for the same DC power a
smaller inverter will operate closer to its nominal power. The final result of the thesis of Hernandez
was an inverter sizing factor (ISF) matrix for a range of module tilt and azimuth angles, shown in Table
6. The ISF is defined as [42]:

65
PSTC ,system
ISF   100% (3.19)
Pinv ,nom

In equation (3.19), PPV refers to the installed STC PV capacity, and Pinv,nom is the nominal inverter
power installed in the system. The optimal size of the inverter can be found by dividing the known P PV
by the ISF found in Table 6. If the ISF is larger than 100%, this means the inverter can be undersized
with respect to the PV system given the likelihood that the module will mostly generate DC power far
below its installed capacity. The ISF values were calculated by implementing a range of inverter sizes
into an annual energy calculation model and finding for which inverter size the optimal AC energy
yield was generated by the PV system [42]. It was found that the optimal inverter size w.r.t. the PV
system capacity was highly dependent on the module tilt and orientation, which is unsurprising given
the strong dependence of incident irradiance on the module positioning. The results of [42] were
calculated for Dutch meteorological conditions and are therefore assumed to be applicable to the
annual energy model of the PVP 2.0. The optimal ISF allows for at most a 1% energy loss due to peak
shaving.

Table 6: Optimal inverter sizing factors in %, as calculated by [42]. The installed PV capacity should
be divided by the ISF to find the best nominal inverter power.
Azimuth\Tilt 0 10 15 30 45 60 75 90
0 111 119 124 156 204 270 322 377
22.5 111 117 124 152 192 232 270 322
45 111 115 121 139 152 168 181 204
67.5 111 115 117 124 126 131 142 159
90 111 108 109 109 109 115 126 145
112.5 111 104 104 101 103 108 119 139
135 111 103 101 98 98 104 117 136
157.5 111 101 100 97 98 103 115 134
180 111 100 97 97 98 104 113 129
202.5 111 101 98 97 98 103 115 131
225 111 103 101 97 100 104 117 136
247.5 111 106 104 101 103 108 119 139
270 111 109 109 108 111 115 124 142
292.5 111 111 117 119 126 131 142 159
315 111 117 119 136 152 163 181 197
337.5 111 117 124 152 186 224 259 307

The ISF model, used in conjunction with the SNL model is referred to as the orientation-
dependent inverter model in this report. The advantage of applying the orientation-dependent
inverter model is that downsizing the inverter both reduces the modelled inverter conversion losses
and the economic costs of the PV system. The retail price of an inverter depends in part on its size, i.e.
its nominal input capacity. If a smaller inverter can be used, the inverter costs can be reduced. This
issue will be discussed further in Chapter 6.

3.6 OVERVIEW OF INPUTS IN THE PERFORMANCE MODEL

The previous sections have mentioned the calculation methods to establish the PV system
performance. These are summarized in Table 7. Table 8 emphasizes the different inputs used for
BAPV and free-standing systems.

66
Table 7: Summary of the loss determination method and source applied to each step in the PV
system performance model.
Step Method or loss factor
GPoA,primary Reindl tilted surface model [23]
ηshading Deration of GPoA,primary by 6.88% for BAPV, 0% for free-standing (Solar Monkey dataset)
ηsoiling Product of rain-free period and soiling factor of 0.083%/day [37]
ηreflectance Fixed: 4% [1, Ch. 20]
ηSTC Manufacturer datasheets
ηirr Logistic dependence of VOC on irradiance together with technology-specific ideality factor
[1, Ch. 20]
ηT Determination of TM with fluid-dynamic model [7], multiplication of temperature
difference from 25 °C with manufacturer datasheet thermal coefficients
ηcable Fixed: 0.5% [1, Ch. 20]
ηMM Fixed: 1.5% [1, Ch. 20]
ηMPPT Fixed: 1% [1, Ch. 20]
ηinv Orientation-dependent Sandia National Laboratory inverter model [42], [43]

Table 8: Difference between BAPV and free-standing performance calculation in the PVP 2.0 model.
Parameter BAPV Free-standing
Albedo 0.15 (typical urban 0.24 (grass)
environment)
Shading factor 0.0688 (Solar Monkey) 0 (no environment or inter-
panel shading)
Mounting structure Direct mount (0 cm standoff) Rack mount
INOCT NOCT + 18 °C NOCT - 3 °C
Module height [m] 7 (rooftop of standard house) 1 (rack mount in field)
Surface roughness z0 [m] 0.4 0.03
Ground/roof temperature Tgr Between TM and Tamb (but Equal to KNMI ground
closest to TM) temperature.

3.7 RESULTS AND MODEL EVALUATION

The final PV system performance model in the PVP 2.0 consists of three functions that can be run
on the website server for any specified system design. Each function has its own time frame: the
HistoryRun() function fetches a climate dataset for the specified location and uses it as input to
calculate the annual energy yield of a system. The function RealRun() fetches all the weather data for
today and calculates the daily system performance. Finally, the function InstantRun() calculates the
real-time, instantaneous performance of the system by fetching only the latest weather
measurements from the database. When the website user designs a PV system on the website, all
three functions are run and the results are displayed in the browser to be examined by the user. An
option was also implemented to download a data file containing the simulation results for anyone
who would like to use the results for further data analysis.

As the purpose of this chapter was to describe the realization of a model to calculate the annual
system performance, only the results of the annual energy calculation function will be discussed here.
The results consist of three sections: an evaluation of the overall annual performance output for a PV
system, a more in-depth look at several components of the performance model, and a comparison of
the model outputs for the PVP 2.0 climate datasets to that of the Meteonorm and KNMI TMYs.

67
3.7.1 OVERALL MODEL PERFORMANCE

The overall model performance is determined by looking at a breakdown of the performance


ratio (PR) of a BAPV and free-standing system. The PR is a factor that indicates how close a system is
operating to its maximum annual efficiency, or alternatively how much losses occur in the system as a
whole. The PR was chosen as the main indicator as it allows for comparison between PV systems of
different size, different module STC efficiency, and different annual in-plane irradiation [44]. The PR of
a system can be calculated according to equation (3.20) [45]:

 E AC 
 
Yf  PSTC  EY  AC
PR    
Yr  HPoA   HPoA   STC (3.20)
   
 GSTC   GSTC 

Where Yf refers to the final system yield, a.k.a. the energy yield (EY) in kWh/kW p of installed
capacity, Yr is the reference yield, or the number of peak sun hours (PSH) for the location in kWh/kW.
HPoA is the annual in-plane radiation in kWh and GSTC is the irradiance under standard test conditions
2
(STC) (1 kW/m ). ηAC is the AC-side, or final, system efficiency, where ηSTC denotes the maximum
system efficiency which would be equal to the STC panel efficiency in case no losses occur in the PV
system. PR can be interpreted as the number of hours of peak power production divided by the
number of hours of peak sunlight that is available. In real-life conditions, a PR of 100% is unattainable
as there will always be irradiance and temperature losses leading to sub-optimal efficiencies [44].

The performance ratio has been a very useful indicator in tracking improvements in PV system
installations. In earlier decades, component failures and module degradation over time resulted in
performance ratios of 60-70%. Analyses of the PR of 6868 PV systems in France and 993 systems in
Belgium installed between 2007-2010 found average PR values of 76% and 78%, respectively [46],
[47]. Most utility-scale PV has the ability to achieve PR values significantly above 80%, and in well-
installed utility-scale PV systems PRs of more than 90% have been measured [48].

In the PVP 2.0 performance model, the PR varies slightly between system designs when the size,
panel choice, and location are changed. The differences are small though, and a PR of ~75-79% is
found for BAPV systems versus a higher PR of ~84-88% for free-standing systems. The lowest PR
values for both systems were found when a CIGS module was chosen, whereas the highest values
occurred for heterojunction panels, which can be attributed to the different sensitivity of the
respective panels to low irradiance conditions.

To examine where the losses occur in the model and to determine why the PR of the BAPV and
free-standing systems differ, Table 9 shows a breakdown of the performance ratio efficiency.

68
Table 9: PR for a BAPV and FS system in Eelde with a tilt of 30° and an azimuth of 180°.
Parameter BAPV Free-standing
ηSTC [%] 20.1 20.1
Installed capacity [kWp] 4.905 4905
3
EAC [kWh/year] 4410 4947·10
2
HPoA [kWh/(m · year)] 1142 1148
ηshading 0.931 1
ηsoiling 0.997 0.997
ηreflectance 0.960 0.960
ηirr 0.955 0.958
ηT 0.996 1.012
ηcable,MM 0.980 0.980
ηinv,MPPT 0.949 0.967
PR 0.787 0.878

It is clear that the primary differences between BAPV and the free-standing (FS) system are due
to the absence of shading in the FS system, the higher temperature efficiency and the higher inverter
efficiency. These outcomes are consistent with the applied performance model. The negligible
temperature effect on PR is surprising, but can be understood when looking at the fluid-dynamic
model together with Dutch meteorological conditions. The important inputs in the FD model are the
ambient temperature, wind speed, and irradiance, where a low ambient temperature and high wind
speed reduce the module temperature under the same irradiance level [7]. The Netherlands has
moderate ambient temperatures, with an average annual daytime temperature of 12.4 °C in the PVP
2.0 climate dataset for Eelde. In addition, wind speeds are high in the Netherlands, with an average
wind speed in Eelde of 4.6 m/s, far exceeding the 1 m/s speed at NOCT level. A third factor is that the
2
plane-of-array irradiance in the Netherlands does not exceed 706 W/m for the evaluated systems, so
2
it is below the NOCT irradiance of 800 W/m . All these factors contribute to a small temperature
effect. Because of the higher ability to lose thermal energy, the FS system even has a positive effect of
temperature as the average module temperature is lower than 25 °C during the year.

The PVP 2.0 model was checked against the output of a commercial PV system modelling
1
software of PV*Sol , to see if the models assume any differences. The evaluated system was a BAPV
system for Herwijnen, using poly-Si panels, a 37° tilt and a 270° azimuth. The results are summarized
in Table 10.

Table 10: PR as calculated by the PVP 2.0 and PV*Sol for a BAPV system in Herwijnen.
Parameter PVP 2.0 PV*Sol
ηSTC [%] 16.5 (Trina module) 15.5 (BenQ module)
Installed capacity [kWp] 2.967 3
EAC [kWh/year] 2171 2515
2
HPoA [kWh/(m · year)] 981 992
ηshading 0.931 1
ηsoiling 0.996 1
ηreflectance 0.960 0.993
ηirr 0.921 0.93
ηT 0.996 0.984
ηcable,MM 0.980 0.980
ηinv,MPPT 0.934 0.960
PR 0.746 0.855

1
https://www.valentin-software.com/en/products/photovoltaics/57/pvsol-premium

69
The PVP 2.0 PR is significantly lower, but this can be explained by the fact that PV*Sol assumes no
rooftop shading. In addition, soiling losses are ignored, and the value for angular reflectance is much
lower in PV*Sol. As research has found that reflectance losses are commonly 2-4% in a PV system, the
loss applied by PV*Sol seems too low [49]. If soiling, shading, and reflectance are ignored, the PR
values of the PVP 2.0 and PV*Sol are comparable. In term of temperature and irradiance effect, the
PV*Sol and PVP 2.0 differ slightly with higher temperature and lower irradiance losses for PV*Sol. As
will be explained in the comparison of the PVP 2.0 climate and Meteonorm and KNMI TMY datasets,
this is most likely due to the fact that PV*Sol uses a TMY dataset with higher irradiance variability.

3.7.2 INDIVIDUAL COMPONENT EVALUATION

The specific behaviour of three components of the performance model is studied in more detail
here: the FD model, the soiling loss model, and the orientation-dependent inverter efficiency model.

3.7.2.1 FLUID-DYNAMIC MODEL

The figures below display the module temperature for every hour in the year as calculated by the
FD model. A comparison of the temperature of a BAPV and FS system is made to emphasize the effect
of the differences listed in Table 8. It can be seen that, as expected, the BAPV module temperature
increases faster with irradiance than the FS temperature, as shown in Figure 27.

Figure 27: Difference in module temperature dependence on irradiance for the BAPV and FS
systems.

Under the same meteorological conditions of the PVP 2.0 climate dataset for Eelde, the module
temperature of the BAPV system exceeds that of the FS system, as illustrated by Figure 28. Figure 29
shows that the BAPV TM exceeds that of an FS at most by 10 °C, and consequently the BAPV system
has higher temperature losses.

70
Figure 28: Hourly calculated module temperature for the BAPV and FS systems using the Eelde
climate dataset.

Figure 29: Difference between TM for the BAPV and FS system under the same meteorological
conditions.

71
The module temperature is also compared to the temperature that would be estimated if an
INOCT model (i.e. a linear relation between module temperature and irradiance) or the Duffie-
Beckman (DB) model (which takes TM as dependent on ambient temperature, irradiance level and
wind speed) would be applied [1, Ch. 20]. The INOCT model expected the highest module
temperature as there is no cooling effect of wind in this model. The DB model found the lowest
module temperatures as the cooling effect of wind is more influential than in the FD model that
depends on a number of other factors. This was found to be true both for BAPV (Figure 30) and for
free-standing PV systems (Figure 31). The scatter in the DB and FD relation between T M and Tamb is
because both depend on other parameters besides irradiance level.

Figure 30: Comparison of TM as calculated with the INOCT, DB, and FD model for a BAPV system.

72
Figure 31: Comparison of TM as calculated with the INOCT, DB, and FD model for a FS system.

In the NOCT model, a constant wind speed of 1 m/s is assumed [1, Ch. 20]. In the climate dataset
for Eelde however, the wind speed never drops below 2 m/s and usually is far above 1 m/s, as shown
in Figure 32. The high wind speeds cool down the module for the DB and FD model more than in the
NOCT case. The NOCT model overestimates the module temperature for all hours of the climate
dataset.

73
Figure 32: Average wind speed for the Eelde climate dataset of the PVP 2.0.

3.7.2.2 SOILING

Figure 33 compares the rainfall patterns of one KNMI year to that of the PVP 2.0 climate dataset.
From Figure 33 it is clear that during a normal year there are frequent occurrences of heavy rainfall
and of dry spells. However, in terms of local climate, rainfall is pretty consistent throughout the year,
with perhaps a slightly drier spring compared to summer.

74
Figure 33: Annual rainfall measured at the weather station of Eelde in 2016 and the average of all
measurements between 1991-2017.

Figure 34: Rain-free period for 2016 and the average daily RFP for 1991-2017. An increase in RFP for
a set of days means that during this period daily rainfall does not exceed 2 mm.

75
Figure 34 reinforces the difference between a weather year and a climate year in terms of
rainfall. During 2016, there were frequent dry spells with the spells of longer duration occurring in the
latter half of the year. The climate RFP stays fairly consistent throughout the year, with the longest
RFPs, in line with the lower rainfall in Figure 33, occurring in the springtime. Figure 34 shows that,
over the lifetime of the panel (i.e. ~25 years), soiling losses are expected to be consistent for all
seasons and that a climate RFP dataset is valid for calculation of soiling losses. As discussed in the
previous section, the annual light capture loss due to soiling are small, in the order of 0.35%. The
range of annual mean RFP observed for the weather stations in the Netherlands was 3.6-4.5 days,
with an overall average of 4.2 days indicating that there is not much geographical difference in soiling
loss in the Netherlands.

3.7.2.3 INVERTER SIZING FACTOR

The practical advantage of using the ISF in the PVP 2.0 is illustrated in the following section. PV
systems with a non-optimal tilt and orientation will have a much smaller inverter capacity than would
be the case if only the installed PV capacity was looked at. The question might arise that in the
occurrence of high PV production for such a non-optimally placed system this will lead to significant
losses when the PV production exceeds the nominal inverter capacity. In the PVP 2.0 model, if the PV
production exceeds the PDC0, i.e. the nominal inverter capacity, the DC input is set equal to P DC0 and
the AC power output is therefore limited as well. For example, a system with an installed capacity of
1635 Wp but an ISF of 204% (tilt of 45° and 0° azimuth) would have a minimal DC inverter capacity of
800 W. The closest match for the inverter in the MySQL database is an inverter with a 958 W p
capacity. If, on one occasion the system should output 1200 Wp, the AC power output is calculated by
using equations (3.18). In the model, this yields an AC power output of 893 W. The effective inverter
efficiency then becomes:

PAC 893
inv    100%  74.4% (3.21)
PDC ,PV 1200

However, due to the positioning of the PV system, the GPoA for all hours throughout the year is so
low that system power never exceeds the downsized inverter capacity, as illustrated in Figure 35
which also includes the efficiency of an inverter with a capacity equal to that of the PV system.

76
Figure 35: Comparison of the PV system power output and the respective efficiencies of a small
inverter selected with the ISF and a large inverter with an input capacity equal to the installed PV
capacity. The module tilt of the system is 40° and the azimuth is 0° (north).

As can be seen, the inverter efficiency of the large system only starts to exceed the small inverter
efficiency above the PV power production values that occur during a year. The results were the same
when the PVP 2.0 climate dataset was used as when a TMY with weather data was used. This showed
that the implementation of the ISF has led to higher PV system efficiencies.

3.7.3 METEOROLOGICAL DATASET COMPARISON

To evaluate whether the results of the PVP 2.0 performance model changes significantly if a TMY
instead of the climate dataset was used, the model was run three times for a south-facing system,
using the PVP 2.0, KNMI TMY, and Meteonorm TMY, respectively. The sections highlighted here are
overall performance, module temperature, and estimation of tilted surface irradiation.

3.7.3.1 PERFORMANCE RATIO

A comparison of the final PR of the three runs is shown in Figure 36 and Figure 37. The main
conclusion is that the final differences in PR are very low.

77
Figure 36: PR breakdown for a BAPV system using the three meteorological datasets.

78
Figure 37: PR breakdown for a FS PV system using the three meteorological datasets.

The above figures show how small the differences are between the weather datasets and the PVP
2.0 climate dataset in terms of PR. The null hypothesis that the climate dataset is a good substitute
for a TMY dataset when calculating the performance of a PV system should therefore not be rejected.

Considering that the non-STC irradiance and temperature losses are the factors most directly
influenced by the meteorological conditions, these losses are looked at in closer detail in Table 11.

Table 11: Non-STC irradiance and temperature PR losses. A negative loss indicates a performance
ratio gain, i.e. a positive effect on the PV system performance.
System type Loss factor [% of PR] PVP 2.0 climate Meteonorm TMY KNMI TMY
BAPV Irradiance 9.4 8.1 7.8
Temperature 0.6 2.7 2.3
Free-standing Irradiance 8.7 7.5 7.2
Temperature -0.1 0.8 0.2

Interpreting the numbers in Table 11, it appears that the use of PVP 2.0 climate data results in
slightly higher irradiance losses and lower temperature losses than the TMY datasets. The fact that
the Meteonorm and KNMI TMYs provide lower irradiance losses and higher temperature losses can
be explained by the higher irradiance variability in the TMY datasets. TMY datasets have a higher
occurrence of both low and high irradiance measurements throughout the year. The calculation of the

79
annual temperature and irradiance effects in essence involves taking an average of the temperature
and irradiance losses in every hour, weighted by the amount of power production of each specific
hour. Hours with a high incident irradiance, i.e. hours with a high system power production, will also
be those hours that have higher temperature and lower irradiance losses than medium-irradiance
hours. As the contribution of the high-irradiance hours to annual energy production is the most
significant, it follows that an annual weighted average of a dataset with high irradiance variability (i.e.
a TMY dataset) will have higher temperature and lower irradiance losses than a climate dataset.

It must be noted that, in any case, the difference in annual performance between the TMY and
climate datasets is relatively small, and that, if it is justified to use Meteonorm or other TMY data for
the calculation of PV system performance, the same justification should be made for the PVP 2.0
climate datasets.

3.7.3.2 MODULE TEMPERATURE

To further analyze the meteorological datasets, an evaluation of the module temperature as


calculated by the FD model for the three meteorological datasets was made. Table 12 shows an
overview of the results for one BAPV system located in Eelde.

Table 12: Annual average values for module and ambient temperature of the three meteorological
datasets.
Parameter Climate KNMI TMY Meteonorm TMY
Mean module temperature (TM) 19.0 20.6 21.7
Standard deviation 11.2 13.9 14.1
Mean ambient temperature (Ta) 12.0 12.3 13.6
Mean TM-Ta 7.0 8.3 9.1
St. dev. TM-Ta 6.9 9.7 10.0

The module temperature during PV system operation is higher for the TMY than for the climate
data. Comparing the module temperature for every hour in the year between the KNMI TMY and PVP
2.0 climate dataset showed that the relation between module temperature and irradiance is more
clear for the climate than the TMY datasets because the influence of other parameters such as wind
speed is subdued, and that in the TMY dataset higher module temperature maxima occur as there are
more hours with a high irradiance. These differences are illustrated in Figure 38. The lines at the high
irradiances of the TMY dataset can be attributed to the fact that the KNMI TMY values only include
integer wind speeds whereas the PVP 2.0 dataset contains wind speeds with several decimals. Low
wind speeds lead to higher module temperatures for the same wind speed.

80
Figure 38: Relation of module temperature deviation from ambient temperature and incident
irradiance.

As expected, the difference in hourly irradiance and other weather factors for the TMY and
climate dataset leads to a difference in the distribution of module temperature throughout the year,
as shown in Figure 39 to Figure 41. Only the values for hours with daylight (GPoA>0) were taken into
account. The lower meteorological parameter variability also causes the module temperature to have
less spread for the PVP 2.0 climate than for the TMY datasets.

81
Figure 39: Distribution of module temperature for every hour in the PVP 2.0 climate dataset for
Eelde.

Figure 40: Distribution of module temperature for every hour in the KNMI TMY dataset for Eelde.

82
Figure 41: Distribution of module temperature for every hour in the Meteonorm TMY dataset for
Eelde.

3.7.3.3 TILTED SURFACE IRRADIATION

The optimal orientation for all three datasets was 35-37° and a south-facing azimuth. It was
noticed however that when the orientations of PV panels were changed from south, the annual tilted
surface irradiation outcomes differed significantly between the datasets. The main point of difference
was that the PVP 2.0 model found there to be a geographical relation between the difference in H PoA
for east- and west-facing modules. A bias to west means that these datasets showed that west-facing
modules received more sunlight on a yearly basis than east-facing modules, which occurred
predominantly in the west of the Netherlands. This is shown in Figure 42.

83
Figure 42: Percentage difference between HPoA for modules with a 270° and 90° tilt for the 46
stations.

Checking this relation for TMY datasets of the location of Eelde (with a slight climate bias toward
west), Ell (with the largest bias to east of all 46 stations), and Vlieland (with the largest bias to west)
resulted in the conclusion that the TMY datasets do not show a bias or show an opposite bias for the
three selected locations. The source of this discrepancy was traced to the DNI component of sunlight,
which varied between datasets while the diffuse and albedo component was relatively comparable.

The annual bias for west over east in DNI is shown in Figure 43 for Vlieland, which indicates that
in the climate dataset, DNI is higher when the Sun is in the western hemisphere of the sky. This
explains why west-oriented modules receive more irradiation in this location.

84
Figure 43: DNI vs Sun elevation for the PVP 2.0 climate dataset of Vlieland, differing between values
for when the Sun is in the east vs. when it is in the west of the sky.

A further evaluation of GPOA for Eelde found that there is a significant inter-annual variation in
the bias between the east and west HPoA for the KNMI data, and that only the long-term average
represented by the PVP 2.0 climate data reveals this geographical trend. The variation in east-west
bias compared to the PVP 2.0 average is summarized in Table 13. It is clear that, while the GHI of each
individual year may be close to the climate average, there can still be a large inter-annual difference
between GPoA in the East and West orientation.

Table 13: GPoA calculated for Eelde during 7 normal weather years (leap years) of the KNMI.
Leap year PVP 2.0
1992 1996 2000 2004 2008 2012 2016 average Climate Meteonorm
GPOA (90°) 539 536 508 624 646 635 655 592 607 625
GPOA (270°) 666 646 611 606 594 618 624 624 642 613
Difference 127 110 103 -18 -52 -17 -31 32 35 -12
West-East
GHI 1006 945 919 1015 1013 1005 1026 990 996 986

The first hypothesis for why this bias occurred was that it was due to an error in the direct
fraction calculation in the PVP 2.0. After evaluation of the clearness index calculation and comparison
with TMY data, no evidence could be found that the bias was due to an error in the data
manipulation. The alternate explanation was that it was a real, measurable climate characteristic for
the Netherlands. This characteristic would be that there is a higher occurrence of clear-sky conditions
at sunset than at sunrise for most locations, or more specifically that there is a lower cloud coverage
over sea than over land for Dutch climate. This was indeed found to be the case, as concluded by an
analysis of KNMI station measurements and NASA satellite imaging performed by PVMD group
Master student Sandeep Mishra [50]. The results of his analysis are shown in Figure 44 and Figure 45.

85
Figure 44: Annual cloud coverage, taken from NASA satellite data measured between 2011-2016.
Figure created by [50].

Figure 45: Annual cloud coverage, taken from KNMI weather station data measured between 2011-
2016. Figure created by [50].

It is clear that the data from Figure 44 and Figure 45, taken from two independent datasets,
support the hypothesis that the West of the Netherlands has lower cloud coverage than the East, and
therefore coastal regions will receive higher irradiance during sunset than during sunrise. Although
not evaluated in the research of [50], it is likely that the cloud fraction decrease from East to West will
continue over the North Sea. Contrary to global satellite observations mentioned in [51], where the
cloud fraction values above the sea are higher than above land, it is nonetheless possible that regional

86
differences may occur worldwide, and that the climate of the Netherlands has higher cloud fractions
above land as a characteristic.

The main conclusion is that there is to be a measurable difference in direct normal irradiance
between the HPoA of east-facing and of west-facing modules in the Dutch climate, which is not
accurately represented in the TMY datasets of KNMI or Meteonorm. The PVP 2.0 climate dataset does
take this long-term trend into account which provides an advantage of the use of the climate dataset
of the TMY in the accurate calculation of tilted surface irradiance.

3.8 DISCUSSION

In this section, some shortcomings of the PV system performance model are treated and
suggestions for further development of the model are given.

An important determinant of the PR for BAPV modules is the shading factor, which is taken from
the data collected by Solar Monkey. It must be kept in mind that the Solar Monkey dataset may be
subject to a self-selection of more shaded locations. The company is hired by external solar panel
installers, so it is possible that these installers may only consult Solar Monkey when the rooftop has
non-optimal attributes, such as an azimuth toward North or trees or buildings nearby that cause
shading. It follows that the dataset of 5398 systems could have a higher average shading factor than
the average shading of the entire rooftop capacity in the Netherlands and therefore not be
representative of Dutch BAPV systems as a whole. Future research could be performed to compare
this dataset to other sources quantifying rooftop shading.

Some of the model efficiency losses were calculated dynamically, but for part of the losses only a
fixed factor was applied. For angular reflectance the fixed loss of 4% could be replaced by a dynamic
calculation based on the model proposed by Martin & Ruiz, which is also integrated into the PVGIS
website [52]. An additional light capture loss which is not modelled in the PVP 2.0 is the effect of
spectral changes in sunlight on the light absorption by PV modules. Future research could look into
modelling this effect based on the research of [49]. In terms of PV efficiency conversion, the factors
that are not yet calculated dynamically are the cabling loss, module mismatch loss, and MPPT tracking
loss. Replacing the fixed losses by a dynamic calculation would bring the PVP 2.0 model to a higher
level.

Several issues that should be addressed in the losses that are calculated dynamically are as
follows. The evaluation of the FD model showed that PV systems in the Netherlands have very a low
non-STC temperature loss which is partly due to the high wind speeds. An issue is that the wind
speeds as measured by the KNMI are taken, which are not the wind speeds in an urban environment.
In cities the wind speed at the panel surface are likely to be lower than currently modeled, meaning
the temperature loss is higher in reality. How to model this issue could be discussed with researchers
at the wind energy research group in the TU Delft. A second issue, relating to the soiling loss is that a
while the RFP is dynamic, a fixed soiling factor is applied for every location in the Netherlands. Further
research needed into how PM10 concentrations translate into soiling factors. After this it would be
possible to use the real-time and historic PM10 measurements of the National Institute for Public
1
Health and the Environment (RIVM) to calculate a dynamic SF that differs between locations in the
Netherlands. A final issue relates to the sensitivity of module efficiency to irradiance. The effect of
shading or low irradiance may affect each panel technology differently. A recent publication by Ziar et
al. addressed this issue and presented a method to determine a "shading tolerability" (ST) of modules,
i.e. their performance under conditions limiting the incidence of light on the panel. The ST includes

1
All information on PM10 is gathered at https://www.rivm.nl/Onderwerpen/F/Fijn_stof#model

87
both uniform (e.g. large clouds or little sunlight) and non-uniform (leaves or partial shading by
surrounding objects) shading [53]. As determination of the ST requires actual testing of individual
modules, the findings of [53] cannot yet be directly applied to the PVP 2.0. In the future, the modules
shown in Table 5 could be tested under the method described by [53], and the ST results could be
integrated into the Dutch PV Portal calculation.

In the ideal case, the production data of a real PV system will be used to validate the real-time
system performance estimate of the PVP 2.0 model. A possible candidate could be the electric bicycle
charging station located close to the EEMCS building in the TU Delft campus. The check with real-time
performance could be used to improve the PVP 2.0 model when it deviates from reality.

3.9 CONCLUSION

This chapter described the realization of the second objective of the PVP 2.0 thesis: the creation
of a model to calculate the annual energy output of a PV system, based on meteorological data and a
user-input PV system design. It can be concluded that this objective has been achieved through
adjustment of the real-time performance model of the PVP 1.0. Several improvements have been
made to the PVP 1.0 model. Two important errors in the PVP 1.0 have been removed in the PVP 2.0
model. In the tilted surface irradiance model, the direct horizontal irradiance (Gdirect) is now translated
to a direct normal irradiance (DNI) before calculating the beam irradiance component, whereas in the
PVP 1.0 Gdirect was seen as equivalent to DNI. The second error was removed by applying a new
formula that decreased the sensitivity of the open-circuit voltage of a module to low irradiance
intensity levels. The removal of both these errors has led to a considerable improvement of the PV
system performance compared to the PVP 1.0.

Next to the removal of errors, additions have been made to the PVP 2.0 performance model.
Most importantly these are: the inclusion of a fixed reflectance and shading loss, the inclusion of a
dynamic soiling loss model based on real-time and historic rainfall measurements, and the
incorporation of an orientation-dependent inverter sizing factor in the Sandia National Laboratory
inverter model which leads to an improved annual inverter performance.

The results of the PVP 2.0 performance model were checked against literature and compared to
the modelling output of PVSol, a commercial modelling program for PV systems. It was concluded that
the PVP 2.0 model yielded results that were within values mentioned in literature. In addition, the
PVP 2.0 includes several real-life performance losses that are ignored or underestimated in PVSol,
most notably angular reflectance, soiling losses, and cabling and module mismatch losses. For better
validation of the PVP 2.0 performance model it is useful to check the performance results against
measurements of an operational PV system.

A final important conclusion concerns the hypothesis posed Chapter 2. The hypothesis stated
that the constructed PVP 2.0 annual climate datasets would be valid for annual PV system
performance modelling. As the PVP 2.0 climate data yielded very similar results to those when using
the KNMI and Meteonorm typical meteorological year (TMY) datasets, this null hypothesis is not
rejected. Additionally, it was found that the PVP 2.0 datasets show a distinct geographical difference
in annual in-plane irradiation between west-oriented and east-oriented PV systems, where in the
western coastal part of the Netherlands it is more beneficial to opt for a west orientation than an east
orientation of PV modules. Such a geographical difference was not present in the Meteonorm TMY,
and the KNMI data showed a large inter-annual variation of this difference. A thorough examination
did not yield any evidence that this difference can be attributed to an error in the meteorological data
or in the manipulation of this data. The most valid explanation, backed up by KNMI and NASA
measurements of cloud coverage, is that, despite the large year-to-year variation, the west parts of

88
the Netherlands and the North Sea have a lower cloud coverage than the eastern regions on a long-
term climate scale. The fact that this information is present in the PVP 2.0 data but absent in the TMY
datasets is a large advantage of the PVP 2.0 climate database for PV system modelling as it will
provide a better prediction of long-term in-plane irradiance than the Meteonorm datasets.

As shown in Figure 3 of the introduction, the PV system performance model can now be applied
to realize the final three objectives of this thesis: a figure displaying the real-time efficiency losses in a
PV system, the calculation of the Dutch national solar energy production, and an economic
profitability analysis that uses the annual energy output of a designed system. Each of these topics
will be discussed in the coming three chapters.

89
4 SYSTEM EFFICIENCY BREAKDOWN
4.1 INTRODUCTION

The PV system performance model of the PVP 2.0, explained in detail in Chapter 3, is based upon
a two-step method: first, the establishment of the irradiance incident on a module in each time step,
and second, the determination of the operating efficiency of each element of the PV system in these
time steps. As put forward in the thesis introduction, the third objective of this thesis is to create a
figure that displays the real-time efficiency losses occurring within a PV system, a so-called system
efficiency breakdown figure. The motivation for this objective is primarily educational, i.e. to provide
website visitors with scientific background on the factors that affect the performance of a PV system.
An additional motivation for creating a system efficiency loss breakdown is to provide the visitors
with additional insight into the PVP 2.0 performance model results by showing them how the final
system efficiency is arrived at.

The creation of electricity from sunlight consists of a sequence of steps in which part of the
energy contained in the incident sunlight is lost. The final AC electricity output of the PV system is the
energy that remains after all losses have been subtracted. The purpose of this chapter is to list the
losses that occur in photovoltaic energy conversion, with specific focus on those losses that have not
been explained in Chapter 3, and to explain how these losses have been integrated into an interactive
chart of the real-time and daily system efficiency breakdown.

The chapter structure is as follows. First a brief overview of the efficiency loss factors will be
given, followed by an explanation of the calculation method for each of these losses. The chapter will
discuss the implementation of the system efficiency breakdown figure in the PVP 2.0 website.

4.2 EFFICIENCY LOSSES OVERVIEW

Figure 46 displays the types of losses considered in the efficiency breakdown model. A distinction
is made between light capture losses (i.e. before the light enters the panel), photovoltaic conversion
losses (i.e. the thermal and electrical losses that occur in the PV module), and balance-of-system
losses (i.e. the losses in the system components after the module). The result is the AC or final system
efficiency.

90
Figure 46: Efficiency losses considered in the system efficiency breakdown. The different box colors
categorize the losses.

4.3 EFFICIENCY LOSS DETERMINATION

4.3.1 LIGHT CAPTURE LOSSES

The values of shading, panel soiling, and primary reflectance losses are output by the PV system
performance model. When looking at the efficiency losses that determine a panel's maximum (STC)
efficiency, one light capture should be considered as well: the area coverage ratio of photovoltaically
active to total module area. Only part of the panel surface consists of photoactive solar cells that can
generate power. All the additional light that falls on the module is considered as the input power, but
the light falling on inactive area cannot be converted to electricity [1, Ch. 10]. The effective efficiency
loss is determined by the area coverage factor (CF), the ratio of total cell to total module area, where
1-CF is the efficiency loss. The CFs for the seven panels in the PVP 2.0 are listed in Table 14:

Table 14: Coverage factor for the PVP 2.0 panels. The values for cell and module area were taken
from the manufacturer datasheets.
2 2
Panel Cell area [m ] Module area [m ] Coverage factor [-]
SunPower mono-Si 1.47 1.63 0.904
Trina mono-Si 1.47 1.64 0.901
Trina poly-Si 1.47 1.64 0.901
Panasonic heterojunction 1.55 1.67 0.925
Si
First Solar CdTe 0.67 0.72 0.926
Solar Frontier CIGS 1.12 1.23 0.909
HyET Solar a-Si 1.65 1.93 0.859

4.3.2 PHOTOVOLTAIC CONVERSION LOSSES

When a solar panel is illuminated, it is able to generate an electrical current. The output power is
always significantly less than the irradiance power of the sunlight itself. This power loss that occurs in
the conversion from sunlight to electricity can be separated in two parts:

91
 Theoretical efficiency losses. Each solar panel has a theoretical limit for conversion
efficiency due to the spectrum of the incident sunlight and the material properties of the
cells.
 Opto-electrical losses. These are losses attributed to the non-ideality of the solar cell,
namely that some optical (light absorption) and electrical (charge carrier collection)
losses are inherent in the design of commercial solar cells [1, Ch. 10]. The better the
design of the cell, the lower these losses and the closer the efficiency of the module can
be to its theoretical limit.

4.3.2.1 THEORETICAL EFFICIENCY LOSSES

The upper efficiency limit for the solar cell materials used in the PVP 2.0 can be determined in
several steps. A distinction between three technology categories is made: the direct bandgap single-
junction devices (CdTe, CIGS), indirect bandgap single-junction devices (mono-, poly- and
heterojunction c-Si), and tandem (dual-junction) devices (micromorph i.e. a-Si/mc-Si). For direct
bandgap materials, three steps are needed, whilst for the indirect bandgap (c-Si) modules an
additional step is needed to factor in the Auger recombination path dominant in c-Si (whereas the
only recombination loss in a direct bandgap material is considered to be radiative recombination) [1,
Ch. 10]. For the tandem a-Si/mc-Si (amorphous/microcrystalline silicon layers stacked on top of each
other) another approach is needed to the single-junction cells as the tandem device theoretically
utilizes the sunlight spectrum in a more efficient manner.

To understand why there is a theoretical upper limit on the conversion efficiency of solar panels,
a discussion on the spectrum of sunlight is warranted. Sunlight consists of a large number of photons,
i.e. energy packets [1, Ch. 5]. Photons have an energy Eph dependent on their frequency ν as follows:

E ph  hP  (4.1)

-34
Here hP is the Planck constant with a value of 6.626 ·10 J·s. As the photon wavelength is
inversely related to its frequency, equation (4.1) indicates that the shorter the photon wavelength,
the higher its energy. Solar cells cannot convert all of the photon energy because firstly the energy of
some photons is too low to be absorbed by the material. This loss is referred to as non-absorption of
photon energy. Secondly, the solar cell can only convert a fixed amount of photon energy into
electrical energy as defined by the bandgap energy of the solar cell, which is a material property. The
difference between the absorbed photon energy and bandgap energy is lost as heat which is why this
loss is referred to as thermalization [1, Ch. 10]. Figure 47 shows the distribution of sunlight energy
over the range of wavelengths for a sunlight spectrum referred to as AM1.5G.

92
Figure 47: AM1.5 solar spectrum. The largest amount of power in sunlight occurs at wavelengths
below 1500 nm.

The non-absorption and thermalization efficiency, i.e. the η abs and ηuse of a solar cell material is
dependent on the material bandgap energy. By multiplying η abs and ηuse the ultimate conversion
efficiency ηult can be determined, which is the conversion efficiency of solar cells at a temperature of
zero K [1, Ch. 10]. The ultimate efficiency as the product of η abs and ηuse is shown in Figure 48 for a
range of bandgap energies.

93
Figure 48: Ultimate conversion efficiency under AM1.5. The maximum conversion efficiency of
49.1% occurs at a bandgap of 1.12 eV, exactly the bandgap of silicon. Note that the x-axis has
changed from wavelength to bandgap energy.

In solar cells operating above zero K, the module is in thermal equilibrium with the environment
and therefore radiative recombination processes take place, which decrease the charge carrier open-
circuit voltage VOC and the fill factor (FF) of the cell. For direct bandgap materials, the actual
theoretical conversion efficiency is called the Shockley-Queisser limit (SQL) [54]. To find the ηSQL, the
ηult must be multiplied with the bandgap utilization efficiency η V and the FF:

 SQL  ult V  FF (4.2)

The ηV and the FF (the bandgap utilization losses) were not calculated directly in this thesis but
inferred from the SQL found in literature and the calculation of non-absorption and thermalization
losses and can be viewed in Table 15.

As stated before, crystalline silicon is an indirect bandgap material and therefore radiative
recombination is not the dominant recombination process in the material. The SQL does not apply,
therefore. An additional step is needed to determine the efficiency losses occurring from Auger
recombination. The required calculations are outside of the scope of this thesis. Instead, an estimate
of the theoretical efficiency limit for crystalline silicon by [55] of 29.43% was used.

For tandem devices, all photons with a wavelength below the mc-Si bandgap wavelength are
absorbed. The thermalization efficiency was calculated by assuming that in an ideal scenario the
higher-bandgap material (a-Si) absorbs all high-energy photons and the lower bandgap material
absorbs the lower-energy photons thereby reducing the thermalization loss. For the tandem devices
as well the SQL does not apply. Next to the fact that mc-Si is an indirect bandgap material, there is an

94
additional consideration to be accounted for in the theoretical efficiency calculation, namely that the
photocurrent in each layer, i.e. the number of photons absorbed in both layers, should be equal as
the layers are connected in series [56]. A study performed by [56] found an upper efficiency limit for
tandem devices of 40%.

An overview of the non-absorption, thermalization, bandgap utilization and theoretical


efficiencies of the various technologies is displayed in Table 15.

Table 15: Non-absorption, thermalization, and bandgap utilization efficiencies for the PVP 2.0
module technologies are displayed. ηv*FF was inferred from the values for ηtheor, ηabsand ηuse.
Material Bandgap [eV] ηabs ηuse ηv*FF ηtheor
c-Si (mono-, poly-, and HIT) 1.12 [1, Ch. 12] 0.81 0.61 0.60 0.29 [55]
CdTe 1.44 [1, Ch. 13] 0.65 0.69 0.72 0.32 [57]
CIS 1.15 [1, Ch. 13] 0.79 0.62 0.67 0.32 [57]
Tandem (a-Si/mc-Si) 1.75/1.1 [56], [58] 0.81 0.77 0.64 0.40 [56]

4.3.2.2 OPTO-ELECTRICAL LOSSES

Now the theoretical limits have been established, there still remains a gap between the
theoretical efficiency and the STC efficiency of a module. The difference between the theoretical
efficiency limit and the efficiency under standard testing conditions (η STC) are caused by the imperfect
photon use in the solar cell. The imperfect photon use refers to optical losses, i.e. that not all photons
with above-bandgap energies are absorbed, and electrical losses, that not all the electrical energy
generated from absorbed photons is collected at the electrodes of the solar panel [1, Ch. 10]. A
precise characterization of the opto-electrical losses in the various module technologies is complex,
and this level of detail was deemed unnecessary for the efficiency breakdown model. The opto-
electrical losses are not calculated directly but determined as the difference between theoretical and
STC efficiency.

4.3.3 BALANCE-OF-SYSTEM LOSSES

The determination of balance of system losses (cabling, module mismatch, MPPT tracking and DC
to AC conversion) have already been discussed in Chapter 3. The PV system performance output
values for these losses are used in the figure.

4.4 RESULTS

The efficiency breakdown figure in the PVP 2.0 was created by hard-coding the conversion
efficiencies discussed in this chapter (i.e. area coverage, non-absorption, thermalization, bandgap
utilization, and opto-electrical losses) and combining them with the losses calculated in the PVP 2.0
PV system performance function. The values for these efficiency losses, calculated in the PHP code,
were passed to the JavaScript code of an interactive Google Chart where the efficiencies could be
displayed on-screen in the browser.

An efficiency breakdown plot was created in two versions: an area plot to show the development
of system efficiency throughout the day, and a bar chart showing the real-time efficiency losses at the
moment of consultation of the website. A screenshot of the area plot is displayed in Figure 49.

95
Figure 49: Efficiency breakdown for a CIGS PV system. The blue area displays the final system
efficiency. When the Sun is below the horizon, system efficiency goes to zero and the figure displays
only one value for “No sunlight”.

The efficiency breakdown figure is placed at two locations on the website: on the case study page
where it shows the efficiency of the 6.9 MWP simulated PV system near Delft, and as an output of the
PV system design model. When a user designs a PV system in the website design page, the efficiency
breakdown for today is automatically displayed. The chart is interactive and responsive to the design
choices made by the user. For example, if a CIGS panel is chosen instead of a mono-Si panel, the
temperature and irradiance loss goes down visibly. At low irradiance conditions, the inverter
efficiency is low as well, in accordance with the applied inverter model. The module technology
determines the cell area coverage and theoretical losses before STC performance, as well as the
irradiance and temperature losses. The location, and thereby the selected weather data, influences
the irradiance and temperature losses as well. Finally, the system size is a factor of importance in the
inverter conversion efficiency. Cabling, mismatch, and MPPT losses are constant for all designs.

4.5 DISCUSSION

This thesis has attempted to list the most important efficiency losses in a scientifically exact
manner. However, some additional level of detail could be incorporated in future development of this
feature.

The efficiency model does not attempt to quantify the optical losses due to the parasitic layer
absorption of an ARC (anti-reflective coating) or a TCO (transparent conductive oxide) but instead
takes all optical and electrical losses together for simplicity's sake. A further development of the
efficiency model could attempt to quantify the optical losses due to standard layers in the various
solar cell technologies.

An important assumption made for this model is that the solar spectrum falling on the module is
always equivalent to the AM1.5 spectrum. However, in reality the solar spectrum is dependent on
atmospheric properties such as cloud coverage and cloud characteristics, humidity, and air

96
temperature. This causes the solar spectrum to differ from AM1.5 throughout the time of the day and
the seasons of the year. In addition, the albedo component of sunlight has a spectrum dependent on
the properties of the material from which it is reflected (among many other factors) [59, Ch. 6].
Further research could characterize the solar spectrum based on the real-time weather values and
calculate the thermalization and non-absorption efficiencies dynamically on the PVP server. This
would allow the model to become even more dynamic.

Next to these scientific considerations, there is a small programming issue that could not be
resolved. A bug in the Google Chart area plot is that negative loss values cause a visual disturbance in
the charts. This means that if the combined temperature and irradiance effect results in a higher-than
STC performance, the temperature and irradiance loss has to be set to 0 instead of displaying the
actual power gain from this effect. This issue could likely be solved with some adjustments to the
chart code but due to time constraints this has not yet been attempted.

4.6 CONCLUSION

The third objective of the PVP 2.0 thesis was the creation of a system efficiency breakdown figure
to display real-time efficiency losses occurring within a PV system. This objective has been
successfully executed and the Dutch PV Portal 2.0 now includes a dynamic system efficiency
breakdown figure both for the real-time system performance as well as the daily system
performance. The implementation of the figure on the website contributes to the goal of the Dutch
PV Portal to communicate the scientific background of photovoltaic energy conversion to the general
public.

The creation of the breakdown figure has been made possible through the PV system
performance model described in the previous chapter. However, this figure is one of three features
that have been enabled by the performance model. The next chapter will discuss the application of
the PV system performance model to a method to calculate the real-time and annual solar energy
production for the Netherlands as a whole.

97
5 NATIONAL SOLAR ENERGY
PRODUCTION MODEL
5.1 INTRODUCTION

Chapter 2 and Chapter 3 have discussed the accurate representation of short-term and long-term
meteorological conditions within the Netherlands and the subsequent determination of the
performance of PV systems under these conditions. Together, the database and performance model
form the backbone of the Dutch PV Portal 2.0 upon which a variety of features can be created. The
previous chapter discussed one such feature, a figure displaying a comprehensive, real-time overview
of PV system efficiency losses.

This chapter will look to an additional application of the PV system performance model to make
an attempt at estimating the total national solar energy production (NSEP, i.e. the amount of
electrical power or energy produced by photovoltaic systems within the Netherlands), both in real-
time and over an entire year, which would realize the fourth objective of the PVP 2.0 thesis. The
reason for installing this objective is quite straightforward: as one of the overall goals of the Dutch PV
Portal project is to provide information on the potential for solar energy generation in the
Netherlands, it follows that displaying the realized national potential of photovoltaic generation is a
valuable attribute for the website. An informed estimate of real-time solar power production can help
in the education of website visitors by indicating the place of solar energy generation in the overall
national electricity production, which would help visitors to consider photovoltaics as part of the
bigger picture.

Besides the educational motivation for the creation of an NSEP model, the current state of affairs
in the Netherlands emphasizes the importance of including an NSEP in the PVP 2.0. Firstly, the
European Union (EU) has stipulated that the Netherlands should achieve a 14% contribution of
renewable energy to the total energy production mix by 2020 [60]. Growth of the solar energy
production in the Netherlands in recent and the coming years will contribute to the achievement of
this goal. The NSEP in the Netherlands has been rapidly increasing in recent years and this production
increase is expected to continue in the coming decades [61], [62]. If the PVP 2.0 manages to include
an estimate of the real-time NSEP and store this data, the website will be able to track the growth of
solar energy in the Netherlands in the coming years, which will be a highly interesting source of data
both to laymen and people that are professionally active in the Dutch solar energy sector.

A final, scientific, motivation for the objective is that the application of the PVP 2.0 performance
model to the topic of NSEP provides the opportunity to compare the results of the PVP 2.0 to that of
official or alternative estimates of the NSEP. This allows for an additional validation or evaluation of
the PVP 2.0 model.

This chapter is structured in the following manner. First, background on the installed capacity and
NSEP in the Netherlands is given. Then a review of existing NSEP models is presented and the
methodology behind these models is evaluated. This is followed by a proposal for a new NSEP
estimation method using the PVP 2.0 meteorological database and PV system performance model.
The required inputs in the PVP 2.0 NSEP model are discussed together with an explanation how
representative values for these inputs were found. In the results section the output of the PVP 2.0
model will be compared to that of other NSEP models, and it will be highlighted how the NSEP results

98
were integrated into the PVP 2.0 website. The discussion will critically assess the applied method, and
finally the conclusion will include an outlook for further development and improvement of the PVP
2.0 NSEP model.

5.2 INSTALLED CAPACITY AND NSEP

At the start of this thesis project, two sources of information were consulted to get an idea of the
state of solar energy and the determination of NSEP in the Netherlands. The first source consisted of
official institution reports on solar energy over the past years. In addition to this, personal
communication was instigated with an employee of the transmission system operator (TSO) TenneT
and an employee of the regional distribution service operator (DSO) Stedin. The summary of the
personal communication can be found in Appendix 9.3. The main conclusions of this in-depth
evaluation are as follows. The primary indicator for the contribution of solar energy to the national
energy production is the national installed capacity of solar panels, as this is the maximum (peak) PV
electricity that can be produced at any given moment. The current size of installed capacity is not
easily determined. As there is no central or legally obligated registration of installed PV systems, an
estimate of the total capacity must be constructed from multiple data sources, of which the most
important registration facilities are the Central Production and Installation Register (C-PIR) in which
households can register there system to qualify for a tax break, and the CertiQ register in which
commercial PV systems are registered that receive subsidy on the produced solar energy. The Central
Bureau of Statistics, the authority responsible for the estimation of renewable energy capacity,
employs seven other data sources to determine the estimate of national installed capacity [63]. This
causes the estimate to have a significant margin of error, and the collection from different sources
means that the end-of-year installed capacity is published multiple months later, i.e. an official
estimate of real-time installed capacity is not publically available.

For the determination of the NSEP a similar uncertainty is the case. There is no legal obligation
for households to share their solar power production and there is no central database that collects
such production values for all PV systems in the Netherlands. TenneT and Stedin also emphasized that
the penetration of solar energy is still so low in the Netherlands that it does not pose a threat to the
stability of the electricity grid. This means that there is no incentive for the TSO and DSOs to spend
effort on accurate real-time modelling of the NSEP. Nevertheless, TenneT is required by the Dutch
government to provide a forecast of solar energy production for the Netherlands to comply with
European regulations.

5.3 EXISTING NSEP MODELS

Given the uncertainties in installed capacity and actual power output of PV systems, creating a
NSEP estimate is therefore challenging. Nevertheless, four NSEP models for the Netherlands of
different organizations were found during this thesis project. Two of them are official estimates
instigated on behalf of the Dutch government: an estimate of the annual NSEP by the CBS and a
forecast of the daily NSEP for the next day by TenneT. Of the two alternate or unofficial NSEP models,
one is created by iCarus, a project within the regional DSO Alliander, and another is created by a
research institute of the Hanzehogeschool Groningen named the Energy Transition Centre
(EnTranCe). The methodology, as far as it is disclosed to the public, of each NSEP is discussed.

5.3.1 CBS

The CBS estimates the NSEP via a simple multiplication of the estimated installed power in kWp
with a 'multiplication factor' for PV system energy yield (EY) in the Netherlands of 875 kWh/kW p. This

99
EY multiplication factor was determined in a study of more than 1500 PV systems in the Netherlands
by Utrecht University [64]. The study found an average annual EY of the systems of ~875 kWh/kW p in
the period of 2012 and 2013, but the large standard deviation of 137 kWh/kW p indicated that large
differences in performance occur between PV systems [64]. The report indicated that the
multiplication factor may need to be adjusted upwards over time as the quality of newly installed PV
systems increases over time as well. The CBS does not provide a real-time estimate, and the annual
NSEP is published after the total installed capacity over the previous year has been established.

5.3.2 TENNET

As mentioned, TenneT is legally obliged to provide an estimate of daily NSEP at a 15-minute


1
resolution. The data are published on the website of ENTSO-E , an umbrella organization for
European TSOs. In personal communication with TenneT, the employee mentioned that TenneT
outsources the creation of this estimate to a commercial partner, but did not wish to disclose the
name of this partner. The only information on the method TenneT provided is that it takes the
meteorological forecast and puts it into a model with the installed capacity to determine the power
output for the next day.

5.3.3 ICARUS/ALLIANDER

Although the impact of solar energy on grid stability is still low, this is likely to change in the
future as capacity grows. The regional DSO Alliander has instigated a project called iCarus to predict
the solar energy production hours or minutes in advance to maintain grid stability in residential areas
2
[65]. The website of iCarus gives a brief description of the applied method. PV system owners are
invited to join the program after which the PV production data of their system is sent to the iCarus
database by internet. This production data is combined with a weather model to forecast short-term
PV production capacity [66]. The website also insinuates that the current official method to calculate
PV production (i.e. that of TenneT) is done by using weather station measurements together with a
large number of assumptions which leads to a too high inaccuracy for DSO requirements. The use of
individual PV system production data throughout the country provides a much better estimate,
according to iCarus.

5.3.4 ENTRANCE

The final NSEP model is that of EnTranCe. In November 2016, EnTranCe launched a website called
3
Energieopwek that shows the real-time energy production from biomass, wind energy, and solar
energy. EnTranCe worked together with TenneT, Gasunie and Netbeheer Nederland to achieve this
[67]. Some information is provided by Energieopwek on the method through which they calculate
solar energy production. Data on known installed capacity is used together with KNMI-data on real-
time irradiance. A model calculates how much energy is produced per W p given a certain irradiance
level [68]. The website itself does not elaborate on how is arrived at the installed capacity. In an
interview with the Dutch Renewable Energy Association NVDE, Martien Visser of EnTranCe stated that
for installed capacity they consult experts in the solar energy market once every 6 months to update
the value [69]. In addition, documents for the monthly energy production from renewables on the
EnTranCe website sometimes give mention of an update of the installed capacity number [70]. No
further details could be found on the model to calculate power output given real-time irradiance. A

1
transparency.entsoe.eu/generation/r2/dayAheadGenerationForecastWindAndSolar/show
2
https://icarus.energy/
3
http://energieopwek.nl/

100
comparison of the EnTranCe estimate with the predictions of TenneT found them to be a close match.
EnTranCe was contacted for this thesis project to check whether TenneT and EnTranCe employ the
same model, but according to EnTranCe the NSEP model of Energieopwek is developed by EnTranCe
itself.

5.4 THE PVP 2.0 NSEP MODEL

5.4.1 PROPOSED METHOD

The information found on the NSEP models employed led to the inception of the idea to create
an NSEP model based on the PVP 2.0 PV system performance model instead of using the estimate of
another organization. The basis of this idea came from the case study present in the PVP 1.0, in which
the real-time performance of a 6.9 MW p field PV system has been tracked since 2014. The power
production data of this system was stored in the database which meant that the long-term energy
production of the system was registered.

The premise of the idea is as follows: if such a real-time simulation is possible for one PV system,
then it must also be possible for a whole set of PV systems that together would be representative of
the total Dutch installed capacity. The main goal is then to determine a representative set of system
characteristics, and input this, together with real-time weather or historic climate data, into the PVP
2.0 energy calculation model. The result of this calculation, a system yield estimate (either in W/W p or
Wh/Wp, dependent on whether the real-time or annual energy model is run), is multiplied with the
respective contribution of this type of system to the total installed capacity. The method is
summarized in Figure 50.

Figure 50: PVP 2.0 NSEP real-time estimation model. EY refers to electricity yield in W/Wp.

The goal is therefore to create a representative design portfolio, based on data from literature
and accessed datasets on PV systems. The EY of all system designs in the portfolio is calculated
independently. The same portfolio is run for the weather data of all twelve provinces, thereby
incorporating the effect of weather differences between regions on production. By knowing which
percentage of the total capacity each design represents, a weighted average of EY can be calculated
for every province. Multiplying this by the installed capacity in every province, the power production
in Watt for the Netherlands can be estimated in near-real-time.

With the PVP 2.0 model, the annual EY will change when the system design portfolio changes,
and when the climate data in the PVP 2.0 database changes over time. By changing the portfolio it is
easy to make adjustments to the final NSEP estimate.

101
As explained in the PV system performance model chapter, the following set of system
characteristics need to be determined to calculate the EY: location, system size, panel technology,
system type (BAPV or free-standing), and orientation (module tilt and azimuth). The next sections
describe how data for each of these inputs were found.

5.4.2 LOCATION

The decision was made to subdivide the total installed capacity into subsets per province, with
each province having the same system characteristics. The motivation for this is that there is some
public information available on the geographic distribution of PV systems on a province level, but that
this information is not available on a more detailed level (only in part in the C-PIR which could not be
accessed in full for this thesis). Therefore the weather and climate data of the provinces, averaged
from the individual weather stations, in the PVP 2.0 database were used. In absence of more recent
data, the distribution key of the Klimaatmonitor for 2015 was used, as shown in Table 16 [71].

Table 16: The installed capacity in each of the 12 Dutch provinces, as well as the installed capacity
normalized to the number of inhabitants per province in the Netherlands in 2015 [71].
Province Installed capacity Installed capacity per person Distribution key
[MWp] [Wp] [%]
Drenthe 83.0 169.9 5.6
Flevoland 67.5 168.1 4.5
Friesland 92.0 142.4 6.2
Gelderland 173.5 85.6 11.6
Groningen 90.0 154.2 6.0
Limburg 143.8 128.6 9.6
Noord- 238.0 95.6 16.0
Brabant
Noord-Holland 142.3 51.5 9.5
Overijssel 178.6 156.6 12.0
Utrecht 76.6 60.6 5.1
Zeeland 50.2 131.9 3.4
Zuid-Holland 152.1 42.3 10.2
Total 1491.2 88.2 100.0

Although it is likely that the distribution of capacity between provinces has changed as the total
capacity grew in 2016 and 2017, the data of the Klimaatmonitor is still the most recent publicly
available data on the geographic distribution. On the PVP 2.0 website the distribution is shown as a
pie chart, represented in Figure 51.

102
Figure 51: Klimaatmonitor distribution key of installed solar capacity over the twelve Dutch
provinces as shown on the PVP 2.0 website.

As noted, the assumption is made that there is little difference in the other system characteristics
(such as size, orientation, panel technology, and system type) between provinces. With no data on
the actual characteristics of PV systems below the national geographic level, it cannot be determined
whether this assumption is valid, but at this moment it is the most logical approach.

5.4.3 SYSTEM SIZE

The system size is the second input characteristic in the PVP 2.0 model. To determine the
distribution of system size, two datasets were used. The datasets each represent one of two types of
PV systems: residential, small-scale PV systems, and commercial PV systems ranging from a small to a
large scale. The TSO Stedin was contacted to find more information on the characteristics of Dutch PV
systems. The notes of an interview with Marko Kruithof of Stedin can be found in Appendix 9.3.1. For
this thesis, Stedin sent an excerpt of the C-PIR with all the residential PV systems located in their grid
network area that were registered in the C-PIR up until February 2017, with a total number of 74312
systems and a total installed capacity of 266.6 MW p (~13% of the national installed capacity at that
time, as estimated by EnTranCe) [70]. Apart from the year of installation and the city in which it was
installed, the system size was the only parameter provided. A histogram of the system size of this
dataset is shown in Figure 52.

103
Figure 52: System size of residential PV systems in the Stedin territory. The 10 kW p bin contains
systems exceeding 10 kWp as well. They are grouped in this bin to reduce the number of system
sizes input into the PVP 2.0 design portfolio.

The information source used to characterise the size of commercial systems is CertiQ, which
registers the production of PV systems that qualify for certificates of origin, i.e. the systems that aim
to sell their energy on the electricity market [72]. This does not include residential household systems,
which save money through net metering, i.e. negating consumed energy with produced energy by
their systems, thereby lowering their electricity bill [1]. CertiQ subdivides the systems by their size, in
eight categories. For simplicity’s sake, the lowest six categories (1-15 kWp) are agglomerated into two
categories of 1-4 kWp and 5-15 kWp.

In the PVP model, the size of the system is not that relevant for its performance in terms of
electricity yield. The size is only used to determine which inverter model and configuration should be
selected, and to determine the total system output in kWh. Therefore, to limit the number of system
design options in the portfolio and thus the number of runs that the model should make, the decision
was made to choose only one average system size for each of these four categories. The size is shown
in Table 17.

Table 17: System size for the four portfolio subsets.


System type Average size [kWp] Contribution to national capacity [%]
Residential 3.1 (C-PIR Stedin) 76.9 [61], [72]
Commercial (1-15 kWp) 4.0 [72] 1.5 [61], [72]
Commercial (16-100 kWp) 51.6 [72] 3.9 [61], [72]
Commercial (>100 kWp) 337.1 [72] 17.7 [61], [72]
Total 70.7 100

104
5.4.4 PANEL TECHNOLOGY

The PVP 2.0 offers seven different module options. In reality, a far larger number of module
brands and types are installed in the Netherlands. As the total national capacity in W p is taken as a
reference point, instead of the installation area, the specific module efficiency is not that relevant for
determining the NSEP. Instead, it is deemed more important to characterise the panels in terms of
the material of which they consist, as each panel technology will respond differently to the
meteorological conditions. However there are multiple sources that list the technology of modules
produced worldwide in 2015, and one recent source for the Netherlands in 2017. These are listed in
Table 18 below.

Table 18: Division of module capacity by technology. The data represents the produced or installed
capacity in a certain year.
Source Production/ Geographic scale Technology Contribution to
installation total annual
year capacity [%]
IEA-PVPS [73] 2015 Worldwide c-Si 6.0
Thin-film 94.0
ITRPV [74] 2015 Worldwide c-Si 92.0
Thin-film 8.0
Fraunhofer ISE [75] 2015 Worldwide Poly c-Si 69.5
Mono c-Si 23.9
CdTe 4.0
CIGS 1.7
a-Si 0.9
Solar Solutions [61] 2017 Netherlands Mono c-Si 59.7
Poly c-Si 36.7
Thin-film 3.6

From Table 18, it can be inferred that the Dutch PV market seems to differ from the worldwide
market as mono-Si, not poly-Si is the most important panel technology. This indicates that the
Netherlands has a preference for higher-efficiency panels. This is not due to the fact that the data are
for the year of 2017. In 2015 as well, there was a 51-44% division between mono- and poly-Si [76].
However, more recent reports on the projections of the PV market between 2015-2020 note that in
the coming years, the bulk production comes from poly-Si, with thin-film silicon destined to remain a
niche technology, at least for the moment [74]. The division of Solar Solutions between the mono-Si,
poly-Si and thin-film shares for 2017 was used in the PVP 2.0 model given the fact that the date are
both applied to the Netherlands and the most recent (and the difference with the mono-Si/poly-Si
division in 2015 is small), together with the division between thin-film technology shares as this
information is absent in [61]. Together, five technologies were selected for the PVP 2.0.

For mono-Si, the run-of-the-mill Trina Solar panel was selected over the SunPower mono-Si and
Panasonic HIT panel, which are both a higher efficiency but also more expensive and more of a niche
technology than is likely to be used on a widespread scale in the Netherlands. The specific modules
used in the PVP 2.0 simulations are shown in Table 19.

105
Table 19: Modules selected per technology and their associated efficiency.
Technology Name Efficiency [%] Share of PVP2.0 capacity
Mono-Si Trina Solar TSM-300DD05A.08 18.3 0.597
Poly-Si Trina Solar TSM-270PD05 16.5 0.367
CdTe First Solar FS-4122-3 17.0 0.021
CIGS Solar Frontier SF175-S 14.3 0.01
a-Si HyET Solar Powerfoil 165 8.6 0.005

5.4.5 SYSTEM TYPE

The PVP 2.0 model differentiates between two types of systems: BAPV systems and free-standing
(FS) systems. The main differences between these two types are the fact that for the BAPV systems, a
shading factor of ~7% is incorporated due to surrounding objects, and that the forced convection
from BAPV modules is much lower, leading to higher cell temperatures and reduced performance
under illumination. FS systems are imagined to be PV systems placed in rows in a field or on a flat
rooftop without significant surrounding object or inter-row shading. The effect of diminished forced
convection in BAPV modules was found to be much lower than the effect of panel shading on power
production in the PVP 2.0 model. Therefore, BAPV systems with little to no shading could be assumed
to be performing almost as well as FS systems.

No publicly available dataset providing such a division between BAPV and FS systems was
available. As an assumption, attention was turned to the previously mentioned datasets concerning
system size. All systems present in the CertiQ dataset are commercial PV systems, and as such are
assumed to be placed with concern for minimized surrounding shading. In addition, the bulk of the
CertiQ dataset capacity, (94%) consists of PV systems with a size exceeding 15 kW p (the other 6% are
systems between 1 and 15 kWp), which would translate to more than 55 poly-Si modules. For systems
of such a size, the location would be subject to a shading analysis in which shading would be
minimized. It is nonetheless true that this division is relatively arbitrary and that in reality there is
likely not such a strict division between residential-BAPV and commercial-FS systems, as for example
there are large farm roofs that can host >15 kW p BAPV systems. However, as a starting assumption
this division is made, which is summarized in Table 20. This assumption can be refined in a future
version of the PVP.

Table 20: Division between residential/BAPV and commercial/FS PV systems in the Netherlands at
the end of 2017.
System type Capacity at end of 2017 [MWp] Contribution to national capacity [%]
Residential/BAPV 2242 [61], [72] 76.9
Commercial/FS 672 [72] 23.1
Total 2914 [61] 100

5.4.6 ORIENTATION

5.4.6.1 RESIDENTIAL SYSTEMS

Residential systems, which make up 76.9% of the total Dutch capacity, have a range of different
orientations. Three datasets have been consulted in order to get a picture of the tilt and azimuth of
Dutch PV systems. The first dataset is provided by the start-up company Solar Monkey. Solar Monkey
has created software to design rooftop PV systems remotely, i.e. without needing to visit the location.
The company provided a dataset of 7648 PV system designs, of which 651 were actually built. For
these systems, the tilt and azimuth of the modules had been determined with the software.

106
The second dataset, consisting of 1542 installed PV systems, is provided by iCarus, a project of
DNO Alliander to predict the solar energy production of households in order to diminish stress on the
electricity grid due to feed-in of PV energy. The prediction is based on real-time production data sent
from household PV system inverters. iCarus calculated the tilt and azimuth of systems based on their
production data during the day. Systems with an eastward azimuth showed higher production values
in the morning than in the evening. The calculated azimuth values of a number of systems were
checked against the real azimuth recorded by the PV system installer, and were found to be in close
agreement. However, the tilt angles could not be calculated accurately with the model. The reason for
this is that a non-optimal tilt can be noticed by a lower-than optimal performance. This does not
indicate whether the tilt is higher or lower than the optimal tilt angle. In addition, a lower
performance could also be due to other losses in the system. Therefore the tilt angles in the iCarus
dataset are not deemed reliable.

The third dataset was obtained through contact with researchers R. Loonen and A. Bognár of
Eindhoven University of Technology (TU/e). Through a joint project of Utrecht University (UU) and TU
/e named the Nationaal Solar Onderzoek (National Solar Investigation), data on the characteristics of
328 PV systems was obtained by requesting system owners to provide details on PV production in
return for an estimate of their performance ratio. Through a remote sensing analysis with LIDAR
height data (similar to the method employed by Solar Monkey), performed by company Autarco, the
tilt and azimuth of the household systems was estimated. The researchers at TU/e are also
performing an analysis of tilt and azimuth based on real-time production data of 500 PV systems. Like
iCarus they found that the azimuth calculation provides accurate results but that several additional
steps may be needed to calculate tilt angles correctly. The results of the production data analysis are
not used in the PVP 2.0 thesis.

The decision was made not to combine the three data sources, as they were obtained with
different methods and in general it is not good practice to aggregate sets created with a different
methodology. In the future the validity of the tilt and azimuth distribution should be tested with
additional data sources.

Figure 53 shows the tilt distribution of all the datasets. The tilt distribution of iCarus is markedly
different from that of the other datasets. This, as mentioned, is because the method for calculation
cannot detect whether a system has a higher than optimal or lower than optimal tilt, which can
explain the peak at 60°. The iCarus data was excluded and instead the Solar Monkey and TU/e data
were studied further, as shown in Figure 54.

107
Figure 53: Tilt distribution of the three datasets employed for the PVP 2.0.

Figure 54: Closer look at the tilt distribution of the Solar Monkey and TU/e datasets.

108
The peak at 10° is most likely for PV systems designed on a flat roof which are placed under a
slight angle. There are almost no PV systems outside the range of 10-55°. In addition, only 0.8 and
0.7% of systems, respectively, have a tilt of 20 or 25°. To reduce the number of bins used in the PVP,
the systems of 20° are assumed to have a 15° tilt, and the 25° systems are assumed to have a 30° tilt.
Given their small share of total capacity, this is expected to have little influence on the final NSEP
result. The final choice is for 8 bins: 10, 15 and 30-55° (in 5° steps).

One might expect that the systems with a 10° or 15° tilt would be the ones placed on a flat roof,
and consequently would have a 180° azimuth. This was not reflected in the data, however, with a
wide range of combinations of tilts and azimuths. This strengthens the approach of forming all
possible combinations of tilts and azimuths and assigning them a weight equivalent to their respective
share in total capacity. The distributions found in both datasets are comparable, with Solar Monkey
having a higher peak at 35°, which is spread out over 25-30° in the TU/e dataset. It must be noted that
the difference in frequency is only 2-3% for the 20 and 25° bins. The dataset of Solar Monkey was
opted for, since it was assumed unlikely that a self-selection in terms of tilt has taken place, and the
advantage of the Solar Monkey data is that the sample is largest.

After the tilt analysis, the azimuth distribution of the datasets was studied.

Figure 55: Azimuth distribution of the three datasets employed for the PVP 2.0.

Figure 55 suggests that a self-selection in terms of azimuth is the case for the Solar Monkey
dataset. PV system installers, unsure whether a certain rooftop in a non-optimal position (i.e. not
facing south) would yield enough energy, might consult Solar Monkey for an analysis, whereas for
south-facing systems they are confident enough to install the system without consultation. This
seems to be the case as the azimuth distribution is much larger than that of iCarus and the azimuths,
determined via a similar method as Solar Monkey, for the dataset of the TU/e. The TU/e and iCarus
data are compared further.

109
Figure 56: Closer look at the azimuth distribution of the TU/e and iCarus datasets.

The TU/e dataset has a larger distribution range than the iCarus data, and a more pronounced
peak at 180°. The iCarus dataset has a smaller range, although some systems with a 110° or 240°
azimuth exist. Because of the larger dataset size, the iCarus sample was opted for. All systems below
140° and above 220° were aggregated in the 140° and 220° bin, respectively because of the negligible
amount of systems outside this range.

In Figure 57 and Figure 58 the selected tilt and azimuth ranges for residential systems are shown.

110
Figure 57: Final tilt distribution, taken from the Solar Monkey dataset, is limited between 10 and
55°, resulting in 10 different options.

Figure 58: Final azimuth distribution, taken from the iCarus dataset, is limited between 140 and
220°, resulting in 9 different options.

111
5.4.6.2 COMMERCIAL SYSTEMS

An east-west configuration has become a common option for the largest solar parks in the
Netherlands. While this may not provide as much energy as a South-facing system at optimal tilt, the
advantage is that the spike in electricity production throughout the day is much lower, leading to a
more uniform electricity production during the day and therefore less stress on the electricity grid the
energy is fed into. An additional advantage cited by solar park installers is the fact that more panels
can be placed on a certain area, as South-facing modules would need to have significant row spacing
to avoid inter-row shading in the winter months [77], [78]. The ground coverage ratio is potentially
92% versus 54% for a South-facing solar park [78]. In East-West configuration, the panels are placed at
a ~15° angle, which reduces horizon pollution and inter-row shading in morning or evening. Some of
the currently largest solar parks in the Netherlands have opted for an East-West configuration, such
as the Sunport Delfzijl, the largest solar park in the Netherlands with an installed capacity of 30 MWp
[79]. Another example is the 7 MWp park in Garyp, Friesland [80]. On the other hand, the 12 MWp
solar park in Woldjerspoor, Groningen was built in a South-facing configuration with modules at a
~20° tilt angle [81]. The previous largest park in the Netherlands is located in Ameland with a 6 MW p
capacity and south-facing modules at a tilt angle of 25° [82]. This lower-than-optimal tilt angle is often
opted for to reduce the required ground area for the capacity by reducing shading lengths. The
relative loss in PoA irradiance, according to the PVP 2.0 climate data, for tilt reduction from 36 to 25°
is low (~1%) as the panels are still positioned in the optimal azimuth angle. Next to these already
active solar parks, the Solarpark Scaldia, a 50 MW p solar park in Vlissingen, Zeeland will become active
in September 2018 and become the largest Dutch solar park. The park will have an East-West
configuration as well and a 10° tilt [83].

Therefore, to take into account these two configuration types, the assumption is made that half
of the largest commercial systems (i.e. above 100 kWp) in the Netherlands use an East-West
configuration, with the other half as South-facing systems at 25° tilt to represent the choice for lower-
than-optimal tilt angles as a trade-off with the area used by a solar park. The exact division is
unknown, but it is clear that both configurations are the most significant for large solar parks. A
further analysis of all solar parks in the Netherlands may yield a more sophisticated estimate of the
orientation distribution of these systems. The small and medium-sized commercial systems are
assumed to have an optimal orientation of 36° tilt and a 180° azimuth, i.e. having orientations closer
to that of the most common residential system.

5.4.7 INSTALLED CAPACITY

Because the CBS only publishes final estimates of installed capacity several months after the year
under scrutiny has ended, it is necessary to rely on alternative estimates of installed capacity. Two
sources that provide publicly available data are EnTranCe and the Nationaal Solar Trendrapport of
Solar Solutions Int.. Every month EnTranCe publishes an overview of energy production in the
Netherlands, subdivided by energy technology. For solar energy, a monthly estimate of the national
installed capacity is provided, based on data of CertiQ, PolderPV (a one-man organisation that tracks
the solar energy market and installed capacity through surveys and research), and the CBS. Data is
available since January 2015. For 2015 and 2016, the installed capacity estimates closely match the
‘official’ annual capacity as published by the CBS afterwards. However, for 2017 there are not yet
official CBS data available. EnTranCe found an end-of-year capacity of 2640 MWp [70].

The Nationaal Solar Trendrapport is a report of the solar market representative Solar Solutions.
The report is published every year at the end of January and provides both an estimate of the
previous year’s installed capacity and a prediction of the market developments expected in the

112
coming year. As was the case, the initial estimate of end-of-year capacity for 2015 and 2016 by Solar
Solutions closely matched the official CBS capacity that was published later on [61], [62], [76].
However, the estimate of installed capacity at the end of 2017 of Solar Solutions was much larger
than that of EnTranCe. In 2017, an estimated 853 MW p was installed, resulting in an end-of-year
capacity of 2914 MWp. This would mean that ~70 MWp was installed every month, assuming a linear
growth. Solar Solutions arrived at this estimate through consultation of the CBS, PolderPV, the TU
Delft, and the ECN (Energieonderzoek Centrum Nederland), through a survey of 149 individuals active
in the solar market. Solar Solutions also contacted installation structure manufacturers for PV systems
to ask for how many solar panels they provided installation structures in 2017.

Comparing the two sources, the Solar Solutions source is deemed more reliable. The EnTranCe
monthly capacity estimates have on multiple occasions been adjusted upwards. For example,
between January 2017 and May 2017, the national capacity increased by 40 MW p per month. In June
2017, the national capacity increased from 2115 MWp to 2290 MWp, an increase of 175 MWp. The
monthly increment in the following months was set at 50 MW p. The Nationaal Solar Trendrapport is
also more open about the methodology used to estimate the national capacity. Therefore the PVP 2.0
used an initial installed capacity of 2914 MWp in the NSEP model as the starting value on 1 January
2018. Every month, 70 MWp is added to the initial capacity to simulate the growth of the national
capacity throughout 2018.

Figure 59: Installed capacity as displayed on the PVP 2.0 website. Data until 2016 is taken from the
CBS, for 2017 and 2018 it is based on the Solar Solutions data.

5.4.8 FINAL MODEL

As mentioned in the previous subchapters, four different subsets are identified: small-scale
residential/BAPV systems with a large combination of tilts and azimuths, small-scale commercial/FS
systems with an assumed optimal positioning, medium-scale commercial systems and large-scale (>15
kWp) FS systems with an equal division between East-West systems at low tilt angles (10°) and South-
facing systems at a 25° tilt angle. Within these four subsets, the five technologies identified in section
0 will be used. Table 21 summarizes the characteristics of the PVP 2.0 NSEP portfolio.

113
Table 21: Portfolio of the 475 PV systems representative of national installed capacity.
Parameter Residential BAPV Small commercial Medium Large commercial
options commercial
Type BAPV FS FS FS
Size [kWp] 3.1 4.0 51.6 337.1
Panel technology 5 options 5 options 5 options 5 options
Tilt and azimuth 10 tilt and 9 1 option 1 option 3 options
azimuth options
Total number of 450 5 5 15
systems
Portfolio share 76.9 1.5 3.9 17.7
[%]

5.5 RESULTS

5.5.1 ESTIMATE OF THE NATI ONAL SOLAR ENERGY PRODUCTION

The national energy production for the current installed capacity portfolio was calculated with
the PVP 2.0 model, using the climate datasets for each province together with the province
distribution key in Table 16. For the total installed capacity of 2984 MWp (as of February 2018), a total
12
annual NSEP of 2.7 TWh (2.67 ·10 Wh) was found. This is equivalent to 2.28% of the total national
electricity production in 2016 [84].

The annual energy yield that was found by the PVP 2.0 simulation is 913 kWh/kW p. This exceeds
the annual yield value estimate of the CBS of 875 kWh/kWp. This could either be because the CBS uses
an outdated value and current PV systems perform better than the research in 2012 and 2013 found,
or because the PVP 2.0 simulation represents an idealized operation of PV systems, whereas in real
life operation some malfunctioning systems may be included in the total capacity. The CBS research in
2014 did indicate that the top-performing systems evaluated had EYs of 900-950 kWh/kWp [64]. The
NSEP estimate of the PVP 2.0 is within that range.

The combined performance ratio of the national installed capacity is 79.4%. Research by
Tsafarakis et al. among 8000 Dutch households with PV systems found a combined PR of 74% for
these systems during the evaluation period [85]. As that research excluded the performance of large-
scale PV plants which commonly have higher PRs, it is likely that the real performance of the Dutch
installed capacity lies higher than that communicated by Tsafarakis et al. and therefore closer to the
PVP 2.0 PR estimate.

5.5.2 COMPARISON WITH ENERGIEOPWEK AND ENTSO-E

The power production estimates of the PVP 2.0 and Energieopwek.nl, the website of EnTranCe,
were compared on several days. The two main noticeable differences are that the PVP 2.0 model
gives a higher estimate of solar power production, and that the national solar power production in
the PVP 2.0 model is more erratic than the Energieopwek model. The higher estimate can most likely
be attributed in part to the higher installed capacity (2914 MW p in January 2018) used in the PVP 2.0
model than the EnTranCe capacity estimate of 2700 MW p. There are undoubtedly more differences
between the EnTranCe and PVP 2.0 energy calculation model, but due to the lack of insight into the
EnTranCe model these differences cannot be pinpointed. The more erratic power production in the
PVP 2.0 is due to irradiance fluctuations in the Netherlands, which cause total production to rapidly
drop or rise. It is unclear why such fluctuations are not shown on the Energieopwek.nl website, but
this could be due to a smoothing function applied to the figures.

114
An important difference between the ENTSO-E/TenneT estimate and that of the PVP 2.0 is that
the TenneT estimate is actually a prediction, based on the weather forecast for the next day. It is
therefore arrived at through a simulation of the performance of a PV system under occurrence of the
predicted weather for tomorrow. According to TenneT, the published forecast is equivalent to the
final solar energy production published on ENTSO-E because of the absence of actual solar production
measurements. The PVP 2.0 and TenneT estimates seem to compare relatively well, as shown in
Figure 60, Figure 61, and Figure 62. Differences may occur due to a different installed capacity value,
differences in the model assumptions, and deviations of the real-time irradiance from the forecasted
irradiance. The PVP 2.0 power output goes to zero before the TenneT estimate, likely due to the
inclusion of the dynamic inverter efficiency, which means that at low sunlight conditions the AC
power output can be zero due to inverter self-consumption. This could also, in part, explain the lower
power prediction of the PVP 2.0 on 4 February, 2018 in Figure 60.

Figure 60: Comparison of the NSEP estimate of the PVP 2.0 and TenneT for 4 February, 2018.

115
Figure 61: Comparison of the NSEP estimate of the PVP 2.0 and TenneT for 6 February, 2018.

Figure 62: Comparison of the NSEP estimate of the PVP 2.0 and TenneT for 8 February, 2018.

116
5.5.3 WEBSITE OUTPUT

On the PVP 2.0 website, visitors are shown an estimate of the NSEP for the current date, as
displayed in Figure 63.

Figure 63: NSEP as calculated by the PVP 2.0 model for 8 February, 2018.

In addition, the database has stored all power production values of each province and the
national total solar power production since 31 January, 2018, with a 10-minute time resolution. On
the website, the user can select any date between 31/01/2018 and today, and view the power output
on that date, as shown in Figure 64. Users can also download the data for their own analysis
purposes.

117
Figure 64: Breakdown by province of the NSEP for 6 February, 2018. The interactive date selector
allows a user to find any date of choice.

5.6 DISCUSSION

The PVP 2.0 model assumes a linear increase in capacity every month. It is very likely that the
growth in the first months of the year is lower than at the end of the year, given the increase in
annual installed capacity between 2015 and 2017. In addition, it is possible that during any given year
there is a difference in when people decide to install a PV system, i.e. more systems may be installed
in the summer months than in the spring or autumn. In absence of reliable data on these two issues,
linear growth is assumed, as an exponential growth factor may yield large errors over time if the
exponent values deviate from reality.

The PVP 2.0 NSEP model as described in this chapter is a proof-of-concept, i.e. the most
important goal was to prove that the PVP 2.0 PV system performance model can be applied to a
nation-wide scope of PV systems. One of the main flaws of the PVP 2.0 NSEP model is that it requires
specific PV system input values for which representative data for the Netherlands is sometimes
absent or outdated. For some input data it is difficult to check the validity as this would require more
in-depth field research which was outside the scope of this thesis.

The best approach to improve upon the PVP 2.0's proof of concept would be twofold. Firstly, it is
advisable that future researchers seek further contact with the most important institutions and
organizations that are actively involved in the collection of data related to installed PV capacity and
real-time solar power production, such as the CBS, TenneT, and regional distribution service operators
(DSOs). A second advice would be for the PVMD group to start a scientific collaboration or exchange
of ideas with the iCarus project of DSO Alliander. Because the iCarus project relies primarily on real-

118
time PV system performance measurements and forecasts of weather conditions, whereas the PVP
2.0 uses a modelled simulation of PV system performance together with real-time KNMI
measurements, the two projects take a different approach. An exchange of ideas would allow for
further comparison of the methods and the possibility for improvement of both NSEP models.

5.7 CONCLUSION

The fourth objective of the PVP 2.0 thesis was the creation of a model to estimate the national
solar energy production (NSEP) for the Netherlands. This objective has been achieved through three
steps: first the collection of information on the PV system characteristics of the installed capacity in
the Netherlands, secondly the creation of a set of 475 PV systems that represent the characteristics of
the Dutch installed capacity, and thirdly, the modelling of the combined performance (specifically the
energy yield) of these 475 PV systems using meteorological data. The power production of all twelve
Dutch provinces and of the Netherlands as a whole is stored permanently in the PVP 2.0 database.
The data can be downloaded by any interested website visitor, thereby contributing to the overall
PVP 2.0 goal of providing publically available information on solar energy.

An important conclusion to be drawn from the research presented chapter is that even in official
numbers on national installed capacity and NSEP significant uncertainties still exist as there is no
comprehensive central registration of installed capacity and no nation-wide collection of PV system
production data. The governmental NSEP estimate uses an energy yield multiplication factor of 875
kWh/kWp established in 2013 which is likely outdated. The PVP 2.0 model suggests that an annual
energy yield of 913 kWh/kW p may be more representative for the current installed capacity in the
Netherlands. Considering that the performance ratios of newly installed PV systems are expected to
be higher than the currently installed capacity, and the fact that the share of free-standing
commercial systems with a high performance ratio in the total Dutch capacity has increased
significantly in recent years, it is likely that the CBS multiplication factor will shift toward the PVP 2.0
NSEP annual energy yield estimate.

Another conclusion that can be drawn from the realization of a PVP 2.0 real-time NSEP estimate
is that combining the meteorological database with the PV system performance model has created a
powerful tool for PV system modelling both that can be applied to individual systems as well as the
country as a whole.

The next and final chapter will look at a third use of the PVP 2.0 performance model in an
economic profitability analysis of user-designed PV systems.

119
6 ECONOMIC PROFITABILITY MODEL
6.1 INTRODUCTION

The previous chapter has highlighted the methodology of PV system performance modelling.
From a scientific viewpoint, modelling the performance and energy production of PV systems is
important. For people and organizations that are not directly involved in the scientific field of
photovoltaics, a more primary concern is the economic profitability of an investment in a PV system.
In the past decades, the price of PV systems has plummeted which has made the technology
commercially available to many [75], [86]. Research of the International Energy Agency (IEA) suggests
that solar energy is set to rival fossil fuels as a low-cost source of solar energy generation in the
coming years [87]. In light of these developments and the public emphasis on the importance of
financial returns on investment, many solar energy websites include an economic profitability
1
analysis . Such a feature was not yet implemented in the Dutch PV Portal 1.0, which has led to the
fifth and final thesis objective to create an economic profitability estimate for user-designed PV
systems.

The goal of this chapter is to describe the model to calculate the profitability of a household or
commercial PV system, using up-to-date economic input parameters that are applicable to the Dutch
solar energy and the national electricity market. The analysis will include the most important
government subsidies and tax breaks for solar energy, as these financial benefits contribute
significantly to the profitability of PV systems in the Netherlands.

The chapter setup is as follows. First the set of profitability indicators that have been selected for
the PVP 2.0 are introduced and the required input variables to calculate these indicators are listed.
After this, an explanation is given on how values for each of the model input variables were found.
The results section will present the outcome of the economic analysis and describe its sensitivity to
user design inputs in the PVP 2.0 website. The discussion will then assess the model validity and
suggest options for future improvement, followed finally by a conclusion.

6.2 ECONOMIC PROFITABILITY INDICATORS

In a profitability analysis for solar energy, multiple indicators are usually calculated to provide
insight into whether an investment is worthwhile. Based upon literature research and Dutch websites
discussing solar energy profitability, the following four economic indicators were selected for the PVP
2.0:

1. The net present value (NPV) of the system. This value represents the total amount of
money that could be earned (or lost, if NPV is negative) if the system is built. To calculate
the NPV, all future costs and earnings are translated into their current monetary value,
i.e. the effects of inflation, interest rate, etc. are removed.
2. The payback period (PBP). The PBP represents the amount of time it will take to
breakeven, i.e. after how much time the total profit will start to outweigh the initial
investment. It follows that the PBP should be smaller than the system lifetime for the
investment to be profitable.

1
For example: SunnyDesign, PVGIS, Project Sunroof by Google, and the Dutch website of De
Energiebespaarders.

120
3. The compound annual growth rate (CAGR). The CAGR describes the average interest
made over the total investment considering the total revenue made on the system over
its lifetime. The value of CAGR can be used for comparisons with interest rates of
alternative investments.
4. The levelized cost of electricity (LCoE). The LCoE is an important parameter when
comparing the production costs of different electricity sources. It represents the amount
of money it costs to produce one kWh of electricity, and therefore LCoE is expressed in
€/kWh. Comparing the LCoE of solar energy to that of conventional (fossil) electricity
sources can show whether photovoltaics are economically competitive with these
sources.
The calculation of each of the four indicators is discussed in the sections below.

6.2.1 NET PRESENT VALUE

The NPV indicates the current monetary value (i.e. the value in € 2018) of the profit made over the
investment lifetime. The formula for NPV is as follows [88]:

Lt
Rt  Ct
NPV  C0  
t 1 (1  r )t (6.1)

In equation (6.1), C0 is the purchasing and installation cost, a.k.a. the initial investment, in year 0.
Lt is the lifetime of the project. Ergo, for every year of the system lifetime, the annual O&M costs (C t)
are subtracted from the annual revenue (Rt) to give the annual profit. Equation (6.1) includes the
discount rate r. The discount rate is a concept related to the time value of money. As money can
accumulate value over time through interest, an amount of money is worth more at the present
moment in time than the same amount would be worth in the future. By discounting future revenues
or costs, these cash flows are made less important in the investment analysis, i.e. short-term gains
become more important than profits further in the future. The discount rate can be seen as the
average interest rate that could be earned with an alternate investment of the same risk. The higher
the discount rate, the more profitable the considered investment must be to be competitive with the
default investment of the money. As an example: instead of purchasing a PV system, a person could
also leave the money on the bank and allow it to accumulate interest, thereby earning an alternative
return on investment.

From equation (6.1) it can be seen that five factors need to be determined to find the NPV: the
system lifetime Lt, the initial investment C 0, the annual revenue Rt, the annual operation and
maintenance costs Ct, and the discount rate r.

6.2.2 PAYBACK PERIOD

The economic payback period of a PV system is the time it takes to earn back the initial
investment. Two types of payback periods are distinguished in profitability studies: the non-
discounted payback period (PBP) which ignores the discount rate, and the discounted payback period
(DPBP) that includes the discount rate. The PBP can be calculated with equation (6.2) [88]:

PBP
C0   Rt  Ct  0 (6.2)
t 1

121
Now, including the discount rate, the DPBP can be calculated with equation (6.3) [88]:

DPBP
Rt  Ct
C 0  
t 1 (1  r )t
0 (6.3)

The connection to equation (6.1) is clear: when the sum of annual profit is equal to the initial
investment, the payback period has been reached. A second conclusion that can be made is that the
DPBP is longer than the PBP. This is logical because PV systems require a high initial investment which
is earned back over the system lifetime. If future profit is discounted, it will take longer to breakeven.
Again, here the same factors as for the NPV need to be found, except for system lifetime: annual
revenue, annual costs, discount rate, and the initial investment.

6.2.3 COMPOUND ANNUAL GROWTH RATE

The compound annual growth rate can be calculated with equation (6.4) [89]:

1
 Lt
 Lt
  Rt 
CAGR   t 1 Lt   1
 

(6.4)
 C 0  C t 
 t 1 
As Lt is the system lifetime, the CAGR represents the average annual return made on the
investment. The numerator in equation (6.4) is the final value of the investment in the system (the
total revenue accumulated over the system lifetime) whereas the denominator represents the total
value of the PV system investment, including the operation and maintenance (O&M) costs. The CAGR
explicitly does not discount the revenues and costs, as the CAGR is an interest rate itself and can
therefore be compared with the discount rate, the average interest rate of an investment portfolio, or
with the interest rate of a savings account. To calculate the CAGR one must first find the system
lifetime, the annual revenue, annual costs, and the initial investment.

6.2.4 LEVELISED COST OF ELECTRICITY

The LCoE, the total economic cost to produce one kWh of energy, can be calculated via equation
(6.5) [87]:

Lt
Ct
C0  
1  r 
t
t 1
LCoE  Lt EPV , t
 (6.5)

1  r 
t
t 1

An accurate estimate of the LCoE divides the discounted investment costs over the system
lifetime by the discounted annual energy production EPV,t. For the LCoE, it is therefore necessary to

122
know the annual energy production, the annual O&M costs, the initial investment, and the discount
rate.

6.2.5 ECONOMIC MODEL FLOWCHART

Figure 65 shows an overview of the PVP 2.0 economic model to calculate each of these four
indicators and the relation between the indicators. The numerical inputs and additional influencing
factors will be described in the following subchapter.

Figure 65: Overview of the general model to calculate four economic profitability indicators. Blue
boxes are numerical inputs and purple boxes are factors influencing these numerical inputs over
the system lifetime.

6.3 INPUT PARAMETER VALUES

In the previous section it was found that only six parameters need to be used as input to calculate
the values of the four economic indicators. These are the system lifetime, the initial investment cost,
the annual operation and maintenance (O&M) cost, the annual energy production, the annual
revenue, and the discount rate. The following section will describe which parameter values are used
as input in the PVP 2.0 economic model.

Before the parameters are discussed, an important disclaimer must be made. An accurate
economic profitability analysis is a complex matter, requiring in-depth analysis of multiple variables
that can affect the profitability of a PV system [90], [91]. For a comprehensive overview of the factors
influencing system profitability, the reader is referred to [91]. Especially utility-scale field PV systems
are subject to a rigorous cost-benefit analysis performed by financial experts [92]. In order to balance
the time dedicated to the objective of creating an economic model for the PVP 2.0 with the time
spent on the other objectives of the website, the decision was made to opt for a preliminary
profitability analysis, which makes a number of simplifying assumptions regarding the input

123
parameters in the model. While this comes at the cost of a less accurate profitability estimate, it will
also limit the complexity of the model and eliminate the need for the website user to provide
additional information as input into the model. Therefore the PVP 2.0 model will be a simplified
economic model that gives a ballpark profitability estimate. In the discussion recommendations are
made for future expansion of the model to increase its accuracy.

6.3.1 SYSTEM LIFETIME

The standard guaranteed performance lifetime promised by solar panel manufacturers is 25


years. This lifetime is cited in the most recent datasheets of Yingli, Trina Solar, Panasonic, SunPower,
and First Solar [93]–[97]. A 25 year lifetime is also commonly assumed in profitability analyses in
literature [91], [92], [98]. In line with the manufacturers and literature studies, the PVP 2.0 model also
takes a 25 year lifetime for the designed PV systems.

6.3.2 INITIAL INVESTMENT

A number of costs are involved in the acquisition and installation of PV system components. For
the purchase of components, commonly a distinction is made between the purchasing costs of the
solar panels, the inverter, and the other system components (referred to as the Balance of System
(BoS)). BoS includes the cabling and mounting structure of the PV system. On top of the component
cost, the installation costs need to be added, which are the man-hours required to install the system.
For field PV systems, two additional costs are usually taken into account: the cost of acquiring the
land upon which the PV system is built, and the financing cost. The financing cost refers to the
interest paid on a mixture of debt and equity financing. The total investment costs for purchasing and
installing the system are usually represented in one figure: the specific system cost expressed in €/Wp.
The specific system cost indicates how much money must be paid to install one Watt-peak of
production capacity.

To determine the initial investment cost for a rooftop PV system, two sources were employed
next to literature. Data of the total cost of a 1456 PV systems, with a capacity of 825-9900 Wp, were
provided by the Delft-based company Solar Monkey. Solar Monkey designs PV systems at the request
of installation companies. The PV system installation companies in return provide Solar Monkey with
their estimates of the installation costs for the designed system. The specific system cost of the Solar
Monkey systems are shown in Figure 2. As a second, alternative, source of information, contact was
made with a solar panel installation company in Delft, owned by Stephan van Berkel. The installation
company has practical experience of the installation costs for residential PV. This company supplied
information on the panel, inverter, and the combined BoS and installation costs for five residential
rooftop PV system designs, with a capacity of 1650-6600 Wp. The specific cost breakdown of the five
systems provided by the installation company is shown in Figure 67.

All the values provided by Solar Monkey and the installation company are excluding value-added
tax (VAT). As the Dutch government provides a tax break, in the form of a VAT exemption, for rooftop
PV system investment costs, it is justifiable to take the non-VAT cost as the actual system cost [99].
The benefit of this tax break is significant, as VAT in the Netherlands amounts to 21%. Commercial PV
systems are not exempted from VAT.

124
Figure 66: Correlation of installed capacity and specific system cost for the 1456 Solar Monkey
systems. No discernable relation between capacity and specific cost is observed. The displayed
values are excluding VAT.

In Figure 2, a large scatter of specific installation costs for the Solar Monkey data is noticeable.
Multiple possible explanations could be given for this. Firstly, as stated by Solar Monkey, the reported
price by the installers is not always accurate, as sometimes a rounded or random value is filled in
during a test. Secondly, as Figure 3 shows, the installation and BoS cost is an important part of the
total system cost. It is likely that some rooftops require a more complicated and therefore more
costly installation than others. As no further explanatory variables for the scatter of the Solar Monkey
systems could be extracted from the data, the decision was made to taken an average installation
value for all systems, regardless of the installed capacity. The average specific system cost for the
1456 systems is €1.51/Wp.

125
Figure 67: Breakdown of the total specific system cost, based on the information provided by a
solar installation company in Delft. The displayed values are excluding VAT.

When comparing the values of the two data sources, an interesting difference was observed. In
the data of Solar Monkey, there was little to no relation between the system capacity and the specific
system cost, as shown in Figure 66. The correlation between panel efficiency and system cost was
checked as well, as higher-efficiency panels commonly have a higher specific cost. However, again
there was no correlation to be found. In contrast with the data of Solar Monkey, the owner of the
installation company mentioned that for small residential systems, a significant economies of scale
effect occurs. When the size of a small PV system increases, the per-watt-peak cost of this system
decreases, as there are fixed installation costs necessary for any PV system, regardless of the size. A
similar economies of scale effect can be applied to the inverters that are installed in a PV system. All
inverters have a minimum fixed cost, regardless of their input capacity. While inverters with a larger
capacity do cost more, their per-watt-peak cost is lower than for smaller inverters (Stephan van
Berkel, personal communication, 6 July 2017). Economies of scale do not occur for the solar panels in
household systems as they have a fixed per-Watt-peak cost. Large commercial installers may be able
to purchase panels from retailers at a reduced price.

The comparison between the data of Solar Monkey and that of the installation company
immediately poses a complexity: while the values of the installation company lie within the cluster of
values of Solar Monkey, the question arises whether to employ economies of scale in the PVP 2.0
calculation. To solve this issue, the decision was made to look at economic profitability analyses in
literature. [98] performed a multi-country analysis of specific system costs for a 1 kW p and 5 kWp
system, and found that the 5 kWp system has a lower specific cost than the 1 kWp system, indicating
economies of scale. The U.S. Department of Energy reported that 8-10kWp systems have, on average,
a 15% lower specific cost than 2-4 kWp systems, observing a lower but still significant economies of
scale effect for commercial-scale PV systems [86]. Solar Bankability, a project of the European Union

126
to identify risks to the profitability of PV systems in the EU, indicates that the specific price of a PV
system goes down when the total system size increases (in the report a comparison is made between
residential (<5 kWp), commercial (5 kWp-1 MWp), and utility-scale (>1 MWp) PV) [92]. An analysis of
installation costs in the Dutch solar energy market in the Nationaal Solar Trendrapport 2018 also hints
to economies of scale [61]. Nonetheless, most reports only analyze two or three system sizes, and no
regression analysis of installed capacity and specific system costs was found in literature.

Considering the above together with the fact that hard-coding economies of scale in the PVP 2.0
will require a more in-depth regression analysis, the PVP 2.0 instead makes a division of designed
systems into three PV system categories, comparable to the approach applied by Solar Bankability,
the Nationaal Solar Trendrapport, and a LCoE study by the International Energy Agency (IEA) [61],
[87], [92]. BAPV systems with an installed capacity below 10 kW p are considered to be residential
systems. The 10 kWp threshold is based on the Central Production and Installation Register (C-PIR)
dataset provided by network operator Stedin, which consists of 73412 small-scale PV systems in the
Netherlands. 97.6% of these systems had an installed capacity below 10 kWp, and therefore the 10
kWp threshold is taken as a maximum size for a residential system. BAPV systems with a capacity
above 10 kWp are assumed to be commercial rooftop systems. Finally, all solar park systems (also
referred to as field PV systems or free-standing PV systems), regardless of the installed capacity, are
assumed to be commercial-scale field PV systems. The motivation for this is that it is unlikely for an
individual to install a free-standing PV system instead of a rooftop PV system. These three categories
are also important for the revenue generated by the PV systems, as will be explained in section 6.3.5.

In Table 22, an overview of specific system costs as found in literature is made for the three
categories. Unfortunately, research for the Netherlands was sparse compared to more mature
markets like Germany and Italy. Therefore, values for Germany are also included in the overview.
Although the data of the solar installation company in Delft show a decreasing price trend for
increasing size, a single value is taken to be representative for small-scale residential systems. Based
on the average size of the 73412 PV systems in the Stedin dataset which is 3082 W p, the specific
system cost for a 3300 Wp system was chosen as the representative value for the solar installation
company data. As mentioned above, the average of all 1456 systems was taken as the value for the
Solar Monkey dataset. The values found by [61] were established through a survey of PV system
installation companies in the Netherlands.

127
Table 22: Specific system cost values found in the datasets and in literature. All residential systems
exclude the 21% VAT for the Netherlands. As commercial systems are not eligible for the VAT tax
exemption under Dutch law, these systems include VAT.
Category System Specific Region Representative Source
size system year
[kWp] cost [€/Wp]
Residential 0.8-9.9 1.51 Netherlands 2017 Solar Monkey dataset
BAPV
3.3 1.45 Netherlands 2017 Solar installation
company dataset
- 1.10 Netherlands 2017 [61]
<5 1.40 EU 2016 [92]
1-5 1.50 Germany 2016 [98]
10 1.29 Germany 2015 [91]
<10 1.50 Germany 2015 [73]
Commercial >500 0.83 Netherlands 2017 [61]
BAPV
>25 0.80 Netherlands 2017 Solar installation
company, personal
communication
<1000 1.20 EU 2016 [92]
10-100 1.27 Germany 2015 [75]
- 1.15 Germany 2015 [73]
Commercial >500 0.92 Netherlands 2017 [61]
free-standing
>25 1.00 Netherlands 2017 Solar installation
company, personal
communication
>1000 1.00 EU 2016 [92]
- <1.00 Germany 2015 [73]

Table 22 shows that the estimates of installation costs vary significantly in literature. Naturally,
the final estimate depends on the methodology employed in the research which could explain some
of the variation, together with the different regions and representative years. No explanation could
be found for why the Nationaal Solar Trendrapport gives such a low price estimate for residential
systems. Considering the values presented in Table 22, the following specific system costs were
selected for the three PVP 2.0 categories. For residential systems, a cost of €1.45/W p was taken, in
line with the design of the solar installation company and because it lay close to most other literature
values. For the commercial BAPV systems, a specific cost of €1.03/Wp was taken. Given the disparity
between the four references for this category, an average was taken as a compromise. For
commercial free-standing systems, the literature values lie close together, and an average specific
cost of €0.97/Wp was taken. In this way, the input values in the PVP 2.0 were chosen to balance the
different estimates in literature.

As in literature the total specific system cost is normally presented instead of a breakdown of the
cost in panel, inverter, BoS and installation costs, the PVP 2.0 model used the total system cost as the
main input as well. To nonetheless incorporate the effect of a certain panel technology choice by the
user, and to include the effect of the inverter sizing factor, described in detail in Chapter 3, two
adjustments to the total system cost are made.

The first adjustment is the effect of panel technology on total system costs. To find information
on the price of the modules used in the PVP 2.0, contact was sought with a Dutch solar panel installer,

128
who provided an up-to-date retail price for six of the seven PV modules used in the PVP 2.0 model.
The only technology for which no price was supplied was the flexible a-Si panel. As an alternative, an
Italian and a Dutch a-Si panel manufacturer were approached with the request to share information
on the retail price of a-Si panels. Unfortunately, both companies were not willing to disclose this
information. Therefore, the price of the CdTe panel was also applied to the a-Si panel, based on the
hypothesis that the a-Si price is higher than that of conventional c-Si modules but comparable to
other thin-film technologies. Table 23 shows an overview of the panel costs.

Table 23: Specific cost of the solar panels used in the PVP 2.0, excluding VAT. *The price of the
flexible a-Si module is unknown, but assumed equal to the CdTe panel.
Panel technology Panel name Specific cost [€/Wp]
Mono-Si SunPower SPR-E20-327 0.885
Cheaper Mono-Si Trina Solar TSM-300DD05A.08 0.515
Poly-Si Trina Solar TSM-270PD05 0.405
Heterojunction Panasonic HIT N330 0.77
CdTe First Solar FS-4122-3 0.60
CIGS Solar Frontier SF175-S 0.52
Flex a-Si HyET Solar Powerfoil 165 0.60*

In the PVP 2.0 model, the standard module cost is assumed to be equal to the Trina Solar mono-Si
panel, i.e. €0.515/Wp, in line with the values of the solar installation company and the module costs in
[61]. If an alternative module is chosen, the difference between this module and the Trina mono-Si
panel is added to the standard total system cost applied for the PV system category. As an example, if
a Trina poly-Si panel is selected instead, the total specific system cost will be reduced by 0.515-
0.405=€0.11/Wp.

The second adjustment takes the inverter sizing factor (ISF) into account. As explained in Chapter
3, PV systems with a non-optimal orientation can opt for a smaller inverter capacity than the installed
panel capacity as the system will almost never reach its peak power at a non-optimal tilt or azimuth in
the Netherlands. The advantages of undersizing the inverter are twofold: the operational efficiency of
the inverter is higher, and the retail price of a smaller inverter is lower resulting in a lower specific
system cost. This second factor is incorporated in the PVP 2.0 model by dividing the standard specific
cost of an inverter by the ISF. As mentioned before, the specific inverter cost is subject to economies
of scale. Therefore the specific inverter cost of the commercial BAPV and the commercial field
systems will be lower than that of residential systems. For residential systems, the standard inverter
cost is taken from the solar installation company data. For a 3300 Wp system, the SMA Sunny Boy 3.0
is installed. At a retail price (excluding VAT) of €788, the specific cost is €0.24/W p. For commercial
BAPV and free-standing systems, the inverter values are taken from research of the Fraunhofer
Institute for Solar Energy (ISE). Assuming that, on average, commercial BAPV systems will be smaller
than solar parks, the specific inverter cost will be slightly higher. For commercial BAPV, inverter costs
are taken to be €0.15/Wp vs. €0.10/Wp for commercial free-standing PV [75].

6.3.3 OPERATION AND MAINTENANCE COSTS

In contrast with conventional energy systems such as gas turbines or coal plants, PV systems do
not have any fuel expenses. Consequently, the fixed annual operation costs for a PV system can be
assumed to be negligible. However, a small annual expense for component maintenance is
incorporated in economic profitability assessments for PV systems. [92] reports a fixed annual O&M
expenditure for residential BAPV of 0.5% of the initial investment, a 1% O&M cost for commercial
BAPV, and a 1.5% O&M expenditure for free-standing field systems. In the PVP 2.0 model, these
values are assumed for the fixed O&M costs of three categories.

129
Next to the fixed annual O&M costs, an expense for replacement of the inverter(s) must be taken
into account. While the panel lifetime is estimated to be 25 years, inverter lifetime is taken to be 12
years, based on the general warranty provided by SolarEdge, one of the most important
manufacturers of inverters for PV systems [100]. This would mean that, after installation in year 0, the
th th
inverter would have to be replaced in the 12 and 24 year of the system lifetime. Considering that
the total PV system lifetime is 25 years, it is unlikely that a PV system owner would replace the
inverter in year 24, especially as the inverter will likely not break down in the first year after the
warranty expires. The assumption of replacing the inverter once is also made in the research
methodology of [98]. Therefore in the PVP 2.0 model, the inverter is/inverters are replaced once at
the start of year 12. This expense is considered as a one-time expense on top of the annual, fixed
O&M costs. Again, the specific inverter cost is multiplied by the ISF to find the inverter replacement
cost in that year.

6.3.4 ANNUAL ENERGY PRODUCTION

The annual energy production of the system is calculated as described in Chapter 3. The
produced electricity can be used either for own consumption, thereby saving on electricity expenses,
or for selling on the electricity market. In addition, the LCoE can be calculated by dividing the total
investment costs over the lifetime by the total electricity produced over the lifetime.

PV panels are subject to performance degradation over their lifetime: when the panels age, their
net power output becomes lower due to material quality reduction. Therefore, the annual energy
output in year t can be calculated as:

t
EPV , t  EPV ,0  1  PD  (6.6)

Here, PD is the annual performance degradation in percent. An analysis of 2000 PV modules and
systems by Jordan and Kurtz found an average performance degradation of 0.7%/year, which leads to
a performance of 84% of the original rated power after 25 years of operation, i.e. above the 80%
performance commonly guaranteed by panel manufacturers [101].

6.3.5 ANNUAL REVENUE

The annual revenue is related to the electricity price in the Netherlands. For regular households,
the government ensures that net metering is allowed. Net metering means that the household is
allowed to subtract the PV production energy from the net electricity consumption, which in effect
means that each kWh of solar energy produced has a value equivalent to the electricity price paid by
consumers. The Dutch government aims to remove the net metering regulation as of 2020, and
replace it with a feed-in tariff in order to reduce the costs spend on solar energy [61], [102]. This is a
signal that the Dutch solar energy market is maturing. More mature markets like the German and
Italian markets already have set a feed-in tariff [90]. In a feed-in tariff system, the consumer receives
a fixed remuneration for the PV energy delivered back to the grid. This remuneration usually is lower
than the consumer electricity price in order to stimulate self-consumption of PV electricity [1, Ch. 21].
At the moment of writing, it is unclear which exact tariff regulation will be put in place by the
government. Therefore the assumption is made that the financial profitability of PV systems already
installed, or installed between 2018 and 2020, will not be harmed by the future changes.

The consumer electricity price fluctuates between years, as shown in Figure 68. In order to take a
more representative electricity price, the average of the consumer electricity price in the period of

130
2010-2016 was taken from the Central Bureau of Statistics (CBS), which is €0.181/kWh. The CBS value
is the average consumer electricity price of all Dutch electricity providers [103].

Figure 68: Average consumer electricity price in the Netherlands for the period of 2010-2016 [103].

For large-scale electricity consumers, such as companies, electricity prices are commonly lower
than for consumers. Data from the CBS shows that for an annual energy consumption of 20-2000
MWh, the electricity price is €0.10-0.11/kWh. A PV system with an annual yield of 800 kWh/kWp and
an installed capacity of 25-2500 kWp would be able to completely satisfy the annual energy demand
of a company within this consumption range. In reality most PV systems with a capacity exceeding 2.5
MWp will be solar parks, as a rooftop would have to be large enough to fit >9000 panels on it to
achieve this capacity. As households have an annual energy consumption of ~3 MWh, it is reasonable
to assume that even a small company building would have an energy demand exceeding that of seven
households [104]. This leads to the assumption that the energy consumption of most companies will
be between 20 and 2000 MWh. Every kWh of solar electricity consumed by the commercial building
will save the company ~€0.105. This leaves the question of how much revenue can be made on
electricity that is not self-consumed, but fed back into the grid. Unlike households, commercial PV
systems are not eligible for net metering [105]. However, all commercial systems with a capacity
exceeding 15 kWp can apply for the Stimulation of Sustainable Energy Production (SDE+) subsidy,
provided by the Dutch government. Whereas net metering and VAT exemption are the most
important support mechanisms for residential PV, SDE+ is the main subsidy for commercial PV, both
for BAPV and free-standing field PV [99]. If the SDE+ subsidy is allotted to an applicant, the owner will
receive a guaranteed, fixed per-kWh remuneration for the produced renewable electricity over a
period of 15 years [106]. The height of this fixed price depends on the height for which the PV system
owner applies. In general, the lower the requested subsidy is, the higher the chance it will be allotted
[106]. For 2018, people can apply for a subsidized electricity price between €0.09/kWh and
€0.112/kWh. An application for an electricity price of €0.105/kWh will have a reasonable chance of

131
being allotted (Stephan van Berkel, personal communication 6 July 2017). The PVP 2.0 model
therefore assigns a fixed €0.105/kWh revenue for electricity produced by commercial BAPV or
commercial solar parks. In reality, no solar park will be built with a lower capacity than 15 kW p as this
is the threshold for SDE+. As a simplification, however, field PV systems with a capacity below 15 kW p
receive a €0.105/kWh revenue as well in the PVP 2.0 model. An additional simplification is that for
both commercial systems, the duration of the SDE+ subsidy applied in the PVP 2.0 model is 25 years,
instead of the 15 years it is in reality. This is comparable to assuming that after 15 years an equally
profitable revenue system will be available for commercial systems to replace the SDE+ subsidy after
the term has expired.

Now an estimate of the electricity price have been established, it is necessary to determine how
this price will change over the coming 25 years. It is difficult to predict the development of the
electricity price in the future as this will depend on market forces as well as the change in the national
electricity mix. Although increases occurred in many European countries, nominal electricity prices in
the Netherlands over the period 2000-2017 have been more or less stable, both for households and
commercial users [103], [107], [108]. The most reliable prediction for the Netherlands comes from the
Energieonderzoek Centrum Nederland (ECN), the largest Dutch research institute in energy-related
topics. In the Nationale Energieverkenning 2017 report, the ECN gave a projection of the electricity
price development between 2017 and 2035 [109]. Until 2020, the market electricity price (excluding
energy taxes and levies) is expected to remain around €30/MWh because of low fuel prices and
production overcapacity (in part due to the growth of renewable energy production in the
Netherlands and Germany). From 2020 to 2035, overcapacity is expected to disappear and the fuel
prices are expected to increase, leading to a wholesale trade price of €50/MWh for electricity in 2035
[109]. This corresponds to a CAGR of 3%, which is taken as the input in the PVP 2.0 model. The ECN
mentions that any prediction of long-term electricity prices, theirs included, is subject to large
margins of error as it depends on a large number of factors of which some are volatile [109]. As the
profitability of a PV system depends for an important part on the electricity price, this introduces a
large but unavoidable uncertainty in the profitability model.

In conclusion, the electricity price, as well as the SDE+ remuneration is expected to increase by
3% per year in the PVP 2.0 model.

6.3.6 DISCOUNT RATE

Determining the value of the discount rate (r) for PV system investments is an important matter.
A difference of a few percent in r can have a large influence on the perceived profitability in the later
years of the system lifetime. Assuming for example that r has a value of 5%, the net profit in year 25
of the system lifetime will be discounted by a factor of 1  0.05 
25
i.e. 3.4. Instead, a higher value of
r, say 8%, will lead to a discount factor of 6.8 in year 25, a factor 2 difference! Given the fact that PV
systems require a high initial investment that is gradually earned back over the lifetime, high r values
will lead to a reduced profitability of the system, represented in lower NPV values [87], [91]. It must
be noted that the height of the discount rate is of less importance to households than to investors.
The alternative investment for the capital of households is usually a savings account on a bank, and
therefore the CAGR of the PV system should only exceed the bank's interest rate for savings accounts.
For commercial investors, however, the alternative investment usually has a high interest rate and
therefore the CAGR of the PV system should be high as well to be attractive to a commercial investor.

The International Energy Agency (IEA) included a lengthy discussion on the influence of r on
LCoEs in the 2015 Projected Costs of Generating Electricity report [87]. In the IEA report, three
scenarios were presented, each with a different r value (3, 7, and 10%), which allow readers to

132
compare the LCoE of energy technologies under different discount rates. According to the IEA, an r of
3% applies to government projects in countries that have trustworthy government bonds and
legislation to ensure consistent rate-of-returns. A 7% discount rate is applied to projects of non-
government investors in a well-regulated country. The 10% rate applies to higher-risk investments,
either due to the project characteristics (such as unproven technologies) or the investment
environment [87].
For the PVP 2.0 model, the decision was made to calculate two interest rate scenarios. The first
scenario applies a low interest rate of 3%, the second a medium interest rate of 7%, in line with the
IEA report. A 3% interest rate could be considered applicable for commercial PV systems supported by
a government-backed loan. The Dutch government supports green funds which provide low-interest
loans for renewable energy projects [99]. A 7% discount rate would be applicable to more
conventional commercial project financing.

The benefit of this multi-scenario approach is that it provides the opportunity to show the PVP
2.0 website visitors how decisive the interest rate can be on the modeled system profitability. The
website will include an explanation of the discount rate to increase visitor understanding of this
concept.

6.3.7 SUMMARY OF APPLIED VALUES

Table 24 shows the final values for the three PV system categories used in the PVP 2.0 economic
model. The PHP script for the model can be found in Appendix 9.1.2.

Table 24: Input values for the PVP 2.0 model, based on the values found in literature and datasets
supplied by solar panel companies. The inverter costs are divided by the inverter sizing factor to
include the economic benefit of inverter downsizing.
Parameter Residential BAPV Commercial BAPV Free-standing
System lifetime [year] 25 25 25
Initial investment 1.45 1.03 0.97
[€/Wp]
Inverter cost [€/Wp] 0.24/ISF 0.15/ISF 0.10/ISF
Annual O&M costs [% 0.5 1 1.5
of initial investment]
Panel performance -0.7 -0.7 -0.7
degradation [%/year]
Electricity revenue 0.181 0.105 0.105
[€/kWh]
Electricity price 3 3 3
increase [%/year]
Discount rate [%/year] 3 (low scenario), 7 3 (low scenario), 7 3 (low scenario), 7
(high scenario) (high scenario) (high scenario)

6.4 RESULTS

The economic analysis was integrated into the PVP 2.0 website by adding it as an output of the
PV system design feature. Whenever a user designs a system, the system design choices are used to
calculate the total system costs whereas the annual energy production is used to calculate the total
annual revenue.

Figure 69 shows the effect of the discount rate on the perceived system profitability. Whereas
the real profit increases steadily from year to year, the net present value is lower due to the
discounting of future profits. PV systems with a higher initial investment and/or a lower annual

133
energy production can have negative NPVs, even if the total profit after 25 years is positive. This
simply indicates that the investment would be less attractive than one with an interest rate equal to
the discount rate.

Figure 69: Screenshot of the economic profitability chart on the website. For every year, the total
profit is compared to the net present value at a 3% and 7% discount rate. The designed system is an
optimally-oriented residential BAPV system with Trina Solar mono-Si panels.

The profitability of the system is sensitive to the panel choice, the orientation, and the choice of
system type. In the PVP 2.0 model it does not pay of to opt for a higher-efficiency module because the
purchasing cost is not paid off through increased energy generation over the lifetime. The further the
module tilt and azimuth are away from the optimum orientation, the higher the LCoE and the lower
the profitability due to the decreased energy generation. Finally, residential systems were found to be
much more profitable than commercial systems because of the twice as high electricity price for
consumers. This electricity revenue is enough to offset the larger initial investment for residential PV
compared to commercial PV.

6.5 DISCUSSION

A point of concern for the validity of the economic analysis is that all the parameters used in the
calculations are subject to change over time as a result of developments in the solar market or
changes in legislation. A well-known example is the spectacular reduction in specific panel costs [87].
These changes will make the input values hard-coded in the PVP 2.0 less representative of the actual
situation on the Dutch solar energy as time goes on. A future overhaul of the PVP 2.0 could provide an
update of the used values. The ECN mentions that any prediction of long-term electricity prices, theirs
included, is subject to large margins of error as it depends on a large number of factors of which some
are volatile [109]. As the profitability of a PV system depends for an important part on the electricity
price, this introduces a large but unavoidable uncertainty in the profitability model. A third important
factor is the subsidy legislation of the Dutch government. The analysis showed that without the net

134
metering and SDE+ subsidy the profitability of Dutch PV systems would be seriously harmed. The fact
that the current Dutch government is considering changes to the revenue legislation for solar energy
could drastically affect the attractiveness of an investment in PV.

For the PVP 2.0 project, a dataset containing representative load profiles (i.e. profiles of power
consumption versus time) for Dutch households was provided by an employee of Eneco. Due to time
constraints, this load profile information could not be integrated in the PVP 2.0. Especially in the light
of the recent decision of the Dutch government to abandon the net metering scheme by 2020 and,
most likely, replace it with a feed-in tariff, the amount of self-consumption of PV system energy by
households will become relevant [102]. Overlapping the load profile for a household with the PV
system production profile for an entire year will make it possible to calculate the self-consumption
level and, from this information, calculate the total annual revenue for households in a feed-in tariff
system. Such a calculation could be included in a third version of the PV Portal.

6.6 CONCLUSION

The fifth objective of the PVP 2.0 thesis was the creation of an economic profitability analysis of a
user-designed PV system in the PVP 2.0 website. In this chapter a preliminary ballpark estimate of PV
system profitability has been described. It was shown that the profitability of a PV system is sensitive
to the choice of input parameter values. In the Netherlands, the profitability of residential PV systems
depends to a large extent on the electricity price and the future development of this electricity price,
together with the generous net metering policy of the Dutch government. Commercial PV systems in
the Netherlands are currently only profitable because they are eligible for the SDE+ renewable energy
subsidy. The current dependence of system profitability on these parameters means that any change
in the parameters will affect the economic analysis outcome. As both electricity price development
and government legislation are notoriously volatile, the only resolution for this issue will be to
periodically monitor their development and make adjustments to the PVP 2.0 model if warranted.
Despite these side notes, it is nevertheless important that an economic analysis is now present in the
website. The mention of the complexities of an adequate economic analysis in the website text will
additionally contribute to the educational purpose of the PVP 2.0 as well.

135
7 OVERALL CONCLUSION
7.1 EXTENT OF ACHIEVEMENT OF OBJECTIVES

This thesis report has discussed the Dutch Photovoltaic Portal 2.0 (PVP 2.0) as a continuation of
the original Dutch PV Portal (PVP 1.0) website on solar energy, developed in the TU Delft Photovoltaic
Materials and Devices (PVMD) group. The main goal of the thesis was to expand upon the PVP 1.0 by
creating a new website that incorporated additional scientific and informative features relevant to the
topic of solar energy, taking the Netherlands as the country of focus. The five objectives that were
realized in the PVP 2.0 are listed below, with a brief summation of the most important results.

1. The creation of a meteorological database from which data can be retrieved that is
applicable to any location within the Netherlands, i.e. with a high geographical resolution.

Two databases were implemented in the PVP 2.0, of which one stores real-time weather
measurements for today and yesterday, and the other stores climate averages for an entire year. A
connection between the real-time and climate database transfers weather values to the climate
datasets, thereby making the climate database dynamic and improving the quality of the climate
averages. The database information can be used for real-time and annual PV system performance
modelling. Data is present for 46 individual weather stations and twelve province averages. The
maximum distance of any location in the Netherlands to a weather station is 35.4 km, which is small
enough to provide a good representation of local meteorological conditions. The addition of the 46
weather stations is a large improvement in spatial granularity of the meteorological data over the PVP
1.0.

2. The creation of a model to calculate the annual energy output of a PV system, based on
meteorological data and a user-input PV system design.

The PVP 1.0 model to calculate the real-time power output of a PV system was adjusted to enable
the calculation of annual energy production using the annual climate datasets of the meteorological
database. Two additional dynamic elements were incorporated in the PV system performance model,
based on research of the PVMD group. The first is a soiling efficiency loss dependent on the amount
of daily rainfall measured at a location. The second is an inverter sizing factor (ISF) that selects a
smaller inverter based on the orientation of the PV system, thereby improving the DC to AC
conversion efficiency.

The model results were consistent with values in literature. In addition, the climate datasets
developed for the first objective were found to be a valid substitution for the 'typical meteorological
year' (TMY) datasets that are standard in PV system modelling. Moreover, for PV module azimuths
deviating from south, the PVP 2.0 climate datasets are argued to give a better estimate of annual in-
plane irradiation than TMY datasets by incorporating the long-term geographical differences in cloud
coverage fractions in the Netherlands.

3. The creation of a figure to display, in real-time, the efficiency losses occurring within a PV
system.

The PV system performance model, combined with knowledge on the theoretical efficiency limits
of different solar panel materials, has allowed for the display of a real-time PV system efficiency
breakdown. The integration of this figure in the website serves two useful roles: the education of

136
website visitors on the practical limits of photovoltaic energy conversion from sunlight, and the
tangible display of the output of the PV system performance model.

4. The creation of a model to estimate the national solar energy production for the
Netherlands, in near-real-time.

Through literature research and personal communication with the TSO TenneT and the DSO
Stedin it was established that the official estimates of the installed capacity and the amount of
national solar energy production (NSEP) contain significant uncertainties, both because data
collection of the geographically dispersed capacity is complex, and because the penetration of solar
energy production in the Netherlands is so low that a high accuracy estimate is not yet of importance
to the official institutions. As an alternative to the official estimates, a NSEP model was developed in
the PVP 2.0 by creating a set of 475 PV systems representative of the characteristics of the national
installed capacity. By modelling the performance of these systems using real-time weather data and
creating a weighted average energy yield, the power production of solar energy in the Netherlands is
estimated in 10-minute intervals. The model outputs realistic values of the NSEP that incorporate the
effect of irradiance changes on a province level which are not visible in the official TenneT estimate.

The PVP 2.0 NSEP model was created as a proof of concept, to show that the website
infrastructure can be applied to a national scale. Two shortcomings of the NSEP model are the fact
that it relies in part on outdated information of PV system characteristics and the geographic
distribution of installed capacity, and that the model may represent an idealized PV system
performance whereas in reality the installed capacity may include a share of malfunctioning PV
systems with a poor performance.

5. The creation of an economic profitability estimate for user-designed PV systems.

The PVP 2.0 website now incorporates a preliminary economic analysis for any user-designed PV
system. This analysis was created by collecting information from literature and through contact with a
residential PV system installation company in Delft. The PVP 2.0 model makes a number of simplifying
assumptions to limit the analysis complexity and the required amount of input data, and should
therefore be considered as a ballpark estimate. Despite these simplifications, the inclusion of the
profitability estimate is expected to contribute to the mainstream appeal of the website.

From the above it can be concluded that, by building upon the solid foundation of the Dutch PV
Portal 1.0, a meteorological database and PV system performance model have been created that
allow for the representative modelling of the solar energy potential at any location in the
Netherlands. Together, the database and the performance model have been used to create three
additional features that highlight aspects concerning the practical application of solar energy. The
informational features incorporated in the PVP 2.0 website are of interest to both laymen and
technical experts. By increasing the spatial granularity of the weather data and creating a real-time
estimate of the national solar energy production, the website now provides a better view of the
Dutch potential for photovoltaic energy generation. The resultant PVP 2.0 website is therefore in line
with the original goal of the PVP 1.0 to create an interactive space for communicating scientific
information on PV in the Netherlands in an accessible manner.

7.2 FUTURE RESEARCH

As the Dutch PV Portal 2.0 was built upon the Dutch PV Portal 1.0, the logical next step would be
to start the development of a third version, i.e. a Dutch PV Portal 3.0. Reiterating some of the
suggestions that were made in the individual chapter discussions, three main topics for future

137
research are identified here. The first relates to the PV system performance model, which calculates
the final PV system production by dynamically calculating the efficiency losses at each step from
incident sunlight to the final AC electricity output. While the majority of efficiency losses are
calculated dynamically in the website based on the user input and meteorological conditions, there
are still a number of losses for which a fixed value is taken. The PVP 3.0 could look at scientific
research that presents models for quantifying angular reflectance losses, cabling and module
mismatch losses, and maximum power point tracking (MPPT) efficiency losses. Additionally the effect
of changes in the spectral distribution of sunlight could be taken into account in the performance
model. The second suggested topic for further research is the NSEP model, for which the validity of
the input data should be evaluated. A promising course of action would also be to present the NSEP
model as it now exists in the PVP 2.0 to organisations such as the DSO Alliander and the Central
Bureau of Statistics (CBS) that may be willing to engage in knowledge exchange of their respective
NSEP models with the TU Delft. A third and final topic would be to increase the complexity of the
economic analysis. Naturally, any additional features related to solar energy that were unexplored in
this thesis could be of interest for a PVP 3.0. It is the hope that the work of this thesis may provide a
good starting point for a future development of the Dutch PV Portal.

138
8 REFERENCES
[1] A. Smets, K. Jäger, O. Isabella, R. Van Swaaij, and M. Zeman, Solar Energy: The physics and
engineering of photovoltaic conversion, technologies and systems, 1st ed. Cambridge, United
Kingdom: UIT, 2016.
[2] ITRPV, “International Technology Roadmap for Photovoltaic (ITRPV) - 2016 Results including
maturity report,” Frankfurt am Main, Germany, 2017.
[3] G. Jifan, “The next energy revolution is already here,” World Economic Forum, 2017. [Online].
Available: https://www.weforum.org/agenda/2017/09/next-energy-revolution-already-here/.
[Accessed: 27-Feb-2018].
[4] World Bank, “Solar Powers India’s Clean Energy Revolution,” 2017. [Online]. Available:
http://www.worldbank.org/en/news/immersive-story/2017/06/29/solar-powers-india-s-
clean-energy-revolution. [Accessed: 27-Feb-2018].
[5] Internet World Stats, “Internet Growth Statistics 1995 to 2017 - the Global Village Online,”
2017. [Online]. Available: http://www.internetworldstats.com/emarketing.htm. [Accessed:
11-Jul-2017].
[6] A. Tozzi, “The Dutch Photovoltaic Energy Portal,” Delft University of Technology, 2014.
[7] M. K. Fuentes, “A simplified thermal model for flat-plate photovoltaic arrays,” Sandia National
Labs., Albuquerque, NM (USA), 1987.
[8] Joint Research Centre, “Performance of Grid-connected PV.” [Online]. Available:
http://re.jrc.ec.europa.eu/pvgis/apps4/PVcalchelp_en.html. [Accessed: 07-May-2017].
[9] KNMI, “stationslijst_nederland,” 2016. [Online]. Available:
http://projects.knmi.nl/datacentrum/catalogus/catalogus/content/nl-obs-surf-
stationslijst.htm. [Accessed: 30-Apr-2016].
[10] World Meteorological Organization, “FAQs - Climate,” 2016. [Online]. Available:
https://public.wmo.int/en/about-us/FAQs/faqs-climate. [Accessed: 10-May-2017].
[11] KNMI, “Warmste jaren,” 2017. [Online]. Available: https://www.knmi.nl/kennis-en-
datacentrum/uitleg/warmste-jaren. [Accessed: 10-May-2017].
[12] Meteonorm, “Features.” [Online]. Available: http://www.meteonorm.com/en/features.
[Accessed: 10-May-2017].
[13] Joint Research Centre, “PVGIS radiation databases.” [Online]. Available:
http://re.jrc.ec.europa.eu/pvgis/apps4/databasehelp_en.html. [Accessed: 10-May-2017].
[14] M. Herrera, S. Natarajan, D. A. Coley, T. Kershaw, A. P. Ramallo-González, M. Eames, D. Fosas,
and M. Wood, “A review of current and future weather data for building simulation,” Build.
Serv. Eng. Res. Technol., p. 143624417705937, Apr. 2017.
[15] S. Pelland, C. Maalouf, R. Kenny, L. Leahy, B. Schneider, and G. Bender, “Solar energy
assessments: When is a typical meteorological year good enough?,” in SOLAR 2016 - American
Solar Energy Society National Solar Conference 2016 Proceedings, 2016, pp. 77–83.
[16] G. J. Levermore and J. B. Parkinson, “Analyses and algorithms for new Test Reference Years
and Design Summer Years for the UK,” Build. Serv. Eng. Res. Technol., vol. 27, no. 4,
pp. 311–325, Nov. 2006.
[17] National Renewable Energy Laboratory, “Weather Data,” 2014. .
[18] Meteonorm, “Handbook part II: Theory,” Bern, Switzerland, 2017.
[19] FreeMapTools, “Radius Around Point,” 2017. [Online]. Available:
https://www.freemaptools.com/radius-around-point.htm. [Accessed: 16-May-2017].
[20] D. Palmer, I. Cole, T. Betts, and R. Gottschalg, “Interpolating and Estimating Horizontal Diffuse
Solar Irradiation to Provide UK-Wide Coverage: Selection of the Best Performing Models,”

139
Energies , vol. 10, no. 2. 2017.
[21] KNMI, “Klimaatatlas - Langjarige gemiddelden 1981-2010,” 2017. [Online]. Available:
http://www.klimaatatlas.nl/. [Accessed: 20-May-2017].
[22] Compendium voor de Leefomgeving, “Jaarlijkse hoeveelheid neerslag in Nederland, 1910-
2015,” 2016. [Online]. Available: http://www.clo.nl/indicatoren/nl0508-jaarlijkse-
hoeveelheid-neerslag-in-nederland. [Accessed: 20-Jan-2018].
[23] D. T. Reindl, W. A. Beckman, and J. A. Duffie, “Diffuse fraction correlations,” Sol. Energy, vol.
45, no. 1, pp. 1–7, 1990.
[24] Sandia National Laboratories, “Global Horizontal Irradiance,” 2014. [Online]. Available:
https://pvpmc.sandia.gov/modeling-steps/1-weather-design-inputs/irradiance-and-
insolation-2/global-horizontal-irradiance/. [Accessed: 10-May-2017].
[25] A. Woyte, R. Belmans, and J. Nijs, “Fluctuations in instantaneous clearness index: Analysis and
statistics,” Sol. Energy, vol. 81, no. 2, pp. 195–206, Feb. 2007.
[26] M. Iqbal, An Introduction to Solar Radiation. Vancouver, Canada: Academic Press Canada,
1983.
[27] F. Kasten and A. T. Young, “Revised optical air mass tables and approximation formula,” Appl.
Opt., vol. 28, no. 22, pp. 4735–4738, 1989.
[28] B. Ridley, J. Boland, and P. Lauret, “Modelling of diffuse solar fraction with multiple
predictors,” Renew. Energy, vol. 35, no. 2, pp. 478–483, 2010.
[29] A. M. Noorian, I. Moradi, and G. A. Kamali, “Evaluation of 12 models to estimate hourly
diffuse irradiation on inclined surfaces,” 2008.
[30] P. G. Loutzenhiser, H. Manz, C. Felsmann, P. A. Strachan, T. Frank, and G. M. Maxwell,
“Empirical validation of models to compute solar irradiance on inclined surfaces for building
energy simulation,” Sol. Energy, vol. 81, no. 2, pp. 254–267, 2007.
[31] A. Padovan and D. Del Col, “Measurement and modeling of solar irradiance components on
horizontal and tilted planes,” Sol. Energy, vol. 84, no. 12, pp. 2068–2084, 2010.
[32] L. Klok, S. Zwart, H. Verhagen, and E. Mauri, “The surface heat island of Rotterdam and its
relationship with urban surface characteristics,” Resour. Conserv. Recycl., vol. 64, pp. 23–29,
2012.
[33] M. Santamouris, “Appropriate materials for the urban environment,” in Energy and climate in
the urban built environment, M. Santamouris, Ed. Abingdon, United Kingdom: Routledge,
2011, pp. 160–180.
[34] CBS, “Hernieuwbare Energie in Nederland 2015,” Den Haag, The Netherlands, 2016.
[35] A. Kimber, L. Mitchell, S. Nogradi, and H. Wenger, “The Effect of Soiling on Large Grid-
Connected Photovoltaic Systems in California and the Southwest Region of the United States,”
in 2006 IEEE 4th World Conference on Photovoltaic Energy Conference, 2006, vol. 2, pp. 2391–
2395.
[36] J. Cano, “Photovoltaic modules: Effect of Tilt Angle on Soiling,” Arizona State University, 2011.
[37] P. Nepal, “Effect of soiling on the PV panel kWh output,” Delft University of Technology, 2018.
[38] C. Hoffmann, R. Funk, M. Sommer, and Y. Li, “Temporal variations in PM10 and particle size
distribution during Asian dust storms in Inner Mongolia,” Atmos. Environ., vol. 42, no. 36, pp.
8422–8431, 2008.
[39] S. A. Hsu, E. A. Meindl, and D. B. Gilhousen, “Determining the power-law wind-profile
exponent under near-neutral stability conditions at sea,” J. Appl. Meteorol., vol. 33, no. 6, pp.
757–765, 1994.
[40] M. Zaaijer, “Introduction to Wind Energy - Wind climate and energy production
(presentation).” TU Delft, Delft, the Netherlands, pp. 1–66, 2015.
[41] J. F. Randall and J. Jacot, “Is AM1. 5 applicable in practice? Modelling eight photovoltaic

140
materials with respect to light intensity and two spectra,” Renew. Energy, vol. 28, no. 12, pp.
1851–1864, 2003.
[42] J. Hernández Castro Barreto, “Optimizing inverter sizing for PV systems according to their
installation characteristics,” Delft University of Technology, 2014.
[43] W. E. Boyson, G. M. Galbraith, D. L. King, and S. Gonzalez, “Performance model for grid-
connected photovoltaic inverters.,” Sandia National Laboratories, Albuquerque, New Mexico,
USA, 2007.
[44] N. H. Reich, B. Mueller, A. Armbruster, W. G. J. H. M. Van Sark, K. Kiefer, and C. Reise,
“Performance ratio revisited: Is PR > 90% realistic?,” Prog. Photovoltaics Res. Appl., vol. 20,
no. 6, pp. 717–726, 2012.
[45] P. Yadav, B. Tripathi, S. Rathod, and M. Kumar, “Real-time analysis of low-concentration
photovoltaic systems: A review towards development of sustainable energy technology,”
Renew. Sustain. Energy Rev., vol. 28, no. 0, pp. 812–823, 2013.
[46] J. Leloux, L. Narvarte, and D. Trebosc, “Review of the performance of residential PV systems in
France,” Renew. Sustain. Energy Rev., vol. 16, no. 2, pp. 1369–1376, 2012.
[47] J. Leloux, L. Narvarte, and D. Trebosc, “Review of the performance of residential PV systems in
Belgium,” Renew. Sustain. Energy Rev., vol. 16, no. 1, pp. 178–184, 2012.
[48] W. van Sark, N. H. Reich, B. Müller, A. Armbruster, K. Kiefer, and C. Reise, “Review of PV
performance ratio development,” in World Renewable Energy Congress, 2012, vol. 6, pp.
4795–4800.
[49] A. M. Gracia and T. Huld, “Performance comparison of different models for the estimation of
global irradiance on inclined surfaces,” Publications Office of the European Union,
Luxembourg, 2013.
[50] S. Mishra, “Simulation map for PV module installation and yield prediction based on Shading
Tolerability and Temperature Coefficient,” Delft University of Technology, 2018.
[51] C. J. Stubenrauch, W. B. Rossow, S. Kinne, S. Ackerman, G. Cesana, H. Chepfer, L. Di Girolamo,
B. Getzewich, A. Guignard, A. Heidinger, B. C. Maddux, W. P. Menzel, P. Minnis, C. Pearl, S.
Platnick, C. Poulsen, J. Riedi, S. Sun-Mack, A. Walther, D. Winker, S. Zeng, and G. Zhao,
“Assessment of Global Cloud Datasets from Satellites: Project and Database Initiated by the
GEWEX Radiation Panel,” Bull. Am. Meteorol. Soc., vol. 94, no. 7, pp. 1031–1049, Jan. 2013.
[52] N. Martin and J. M. Ruiz, “Calculation of the PV modules angular losses under field conditions
by means of an analytical model,” Sol. Energy Mater. Sol. Cells, vol. 70, no. 1, pp. 25–38, 2001.
[53] H. Ziar, B. Asaei, S. Farhangi, M. Korevaar, O. Isabella, and M. Zeman, “Quantification of
Shading Tolerability for Photovoltaic Modules,” IEEE J. Photovoltaics, vol. 7, no. 5, pp. 1390–
1399, 2017.
[54] W. Shockley and H. J. Queisser, “Detailed balance limit of efficiency of p‐n junction solar
cells,” J. Appl. Phys., vol. 32, no. 3, pp. 510–519, 1961.
[55] A. Richter, M. Hermle, and S. W. Glunz, “Reassessment of the Limiting Efficiency for
Crystalline Silicon Solar Cells,” IEEE J. Photovoltaics, vol. 3, no. 4, pp. 1184–1191, 2013.
[56] A. V Shah, M. Vaněček, J. Meier, F. Meillaud, J. Guillet, D. Fischer, C. Droz, X. Niquille, S. Faÿ, E.
Vallat-Sauvain, V. Terrazzoni-Daudrix, and J. Bailat, “Basic efficiency limits, recent
experimental results and novel light-trapping schemes in a-Si:H, μc-Si:H and `micromorph
tandem’ solar cells,” J. Non. Cryst. Solids, vol. 338–340, no. Supplement C, pp. 639–645, 2004.
[57] S. Rühle, “Tabulated values of the Shockley–Queisser limit for single junction solar cells,” Sol.
Energy, vol. 130, pp. 139–147, Jun. 2016.
[58] A. V Shah, J. Meier, E. Vallat-Sauvain, N. Wyrsch, U. Kroll, C. Droz, and U. Graf, “Material and
solar cell research in microcrystalline silicon,” Sol. Energy Mater. Sol. Cells, vol. 78, no. 1, pp.
469–491, 2003.
[59] F. T. Si, “Quadruple-junction thin-film silicon-based solar cells,” Delft University of Technology,

141
2017.
[60] European Commission, “Report from the Commission to the European Parliament, the
Council, the European Economic and Social Committee and the Committee of the Regions -
Renewable Energy Progress Report,” Brussels, Belgium, 2017.
[61] Solar Solutions Int., “Nationaal Solar Trendrapport 2018,” Haarlem, the Netherlands, 2018.
[62] CBS, “Hernieuwbare elektriciteit; productie en vermogen,” 2016. [Online]. Available:
http://statline.cbs.nl/Statweb/publication/?DM=SLNL&PA=82610ned. [Accessed: 09-Jan-
2017].
[63] CBS and RVO, “Protocol Monitoring Hernieuwbare Energie - herziening 2015,” Utrecht, the
Netherlands, 2015.
[64] W. van Sark, L. Bosselaar, P. Gerrissen, K. B. D. Esmeijer, P. Moraitis, M. van den Donker, and
G. Emsbroek, “Update of the Dutch PV specific yield for determination of PV contribution to
renewable energy production: 25% more energy!,” Proc. 29th EUR-PSEC, pp. 4095–4097,
2014.
[65] iCarus, “iCarus zonneopwekvoorspeller,” 2018. [Online]. Available: https://icarus.energy/.
[Accessed: 01-Mar-2018].
[66] iCarus, “iCarus zonneopwekvoorspeller,” 2018. .
[67] EnTranCe, “Realtime inzicht in duurzame energieproductie met nieuwe web-app,” 2016.
[Online]. Available: http://en-tran-ce.org/realtime-inzicht-in-duurzame-energieproductie-
met-nieuwe-web-app/. [Accessed: 16-Jan-2017].
[68] Energieopwek, “Over deze website,” 2016. [Online]. Available: http://energieopwek.nl/.
[Accessed: 16-Jan-2017].
[69] NVDE, “Inzicht in de dagelijkse productie duurzame energie,” 2016. [Online]. Available:
http://www.nvde.nl/756-2/. [Accessed: 16-Jan-2017].
[70] EnTranCe, “Renewable Energy in the Netherlands,” 2017. [Online]. Available: http://en-tran-
ce.org/newsletter-renewable-energy-in-the-netherlands/. [Accessed: 16-Jan-2017].
[71] Rijkswaterstaat, “Klimaatmonitor,” 2016. [Online]. Available:
https://klimaatmonitor.databank.nl/Jive. [Accessed: 09-Jan-2017].
[72] CertiQ, “Statistisch jaaroverzicht CertiQ 2017,” Arnhem, The Netherlands, 2018.
[73] IEA PVPS, Trends 2016 in Photovoltaic Applications, 2016th ed. Paris, France: International
Energy Agency, 2016.
[74] ITRPV, “International Technology Roadmap for Photovoltaic - 2015 Results including maturity
reports,” Frankfurt am Main, Germany, 2016.
[75] S. Philipps and W. Warmuth, “Photovoltaics report 2016,” Freiburg, Germany, 2016.
[76] Solar Solutions Int., “Nationaal Solar Trendrapport 2016,” Haarlem, the Netherlands, 2016.
[77] Zonneparken Nederland B.V., “Niet alleen op zuid, maar ook oost west.” [Online]. Available:
http://www.zonneparkennederland.nl/niet-alleen-op-zuid-maar-ook-oost-west.html.
[Accessed: 25-Jan-2018].
[78] Groenleven, “Oost-west opstelling,” 2018. [Online]. Available:
https://www.groenleven.nl/zonneparken/oost-west-opstelling. [Accessed: 25-Jan-2018].
[79] NOS, “Delfzijl heeft grootste zonnepark van Nederland,” 2017. [Online]. Available:
https://nos.nl/artikel/2153827-delfzijl-heeft-grootste-zonnepark-van-nederland.html.
[Accessed: 25-Jan-2018].
[80] Groenleven, “Zonnepark op vuilstort,” 2017. [Online]. Available:
https://www.groenleven.nl/projecten/zakelijk/zonnepark-op-vuilstort-groenleven-2.htm.
[Accessed: 25-Jan-2018].
[81] Groenleven, “Zonnepark op vuilstort Woldjerspoor,” 2017. [Online]. Available:
https://www.groenleven.nl/projecten/zakelijk/zonnepark-op-vuilstort-woldjerspoor-

142
groenleven.htm. [Accessed: 25-Jan-2018].
[82] D. Ameland, “Zonnepark Ameland,” 2017. [Online]. Available:
http://www.duurzaamameland.nl/projecten/. [Accessed: 26-Jan-2018].
[83] W. Fikken, “Solarpark Scaldia - Ruimtelijke onderbouwing,” Vlissingen, The Netherlands, 2016.
[84] CBS, “Elektriciteitsbalans; aanbod en verbruik,” 2018. [Online]. Available:
https://opendata.cbs.nl/statline/#/CBS/nl/dataset/00377/table?ts=1518107934998.
[Accessed: 25-Feb-2018].
[85] O. Tsafarakis, P. Moraitis, B. B. Kausika, H. Van Der Velde, S. Hart’T, A. de Vries, P. de Rijk, M.
M. De Jong, H. P. Van Leeuwen, and W. Van Sark, “Three years experience in a Dutch public
awareness campaign on photovoltaic system performance,” IET Renew. Power Gener., vol. 11,
no. 10, pp. 1229–1233, 2017.
[86] D. Feldman, G. Barbose, R. Margolis, M. Bolinger, D. Chung, R. Fu, J. Seel, C. Davidson, N.
Darghouth, and R. Wiser, “Photovoltaic System Pricing Trends: Historical, Recent, and Near-
Term Projections 2015 Edition,” 2015.
[87] IEA, “Projected Costs of Generating Electricity - 2015 Edition,” Paris, France, 2015.
[88] M. Bortolini, M. Gamberi, A. Graziani, C. Mora, and A. Regattieri, “Multi-parameter analysis
for the technical and economic assessment of photovoltaic systems in the main European
Union countries,” Energy Convers. Manag., vol. 74, pp. 117–128, 2013.
[89] Investopedia, “Annual return,” 2018. [Online]. Available:
https://www.investopedia.com/terms/a/annual-return.asp. [Accessed: 17-Feb-2018].
[90] F. Cucchiella, I. D’Adamo, and M. Gastaldi, “A profitability assessment of small-scale
photovoltaic systems in an electricity market without subsidies,” Energy Convers. Manag., vol.
129, pp. 62–74, 2016.
[91] T. Lang, E. Gloerfeld, and B. Girod, “Don‫׳‬t just follow the sun – A global assessment of
economic performance for residential building photovoltaics,” Renew. Sustain. Energy Rev.,
vol. 42, pp. 932–951, Feb. 2015.
[92] C. Tjengdrawira, M. Richter, and I. Theologitis, “Best Practice Guidelines for PV Cost
Calculation - Accounting for Technical Risks and Assumptions in PV LCOE,” Bolzano, Italy,
2016.
TM
[93] First Solar, “First Solar Series 4 PV Module.” First Solar, Tempe, Arizona, USA, 2017.
[94] Trina Solar, “The Allmax Plus+ framed 60-cell module.” Trina Solar, Changzhou, China, 2017.
[95] Yingli Solar, “YGE 60 Cell - Series 2.” Yingli Solar, Baoding, China, 2017.
[96] SunPower, “SunPower® E-Series Residential Solar Panels | E20-327.” SunPower, San Jose,
California, USA, 2017.
[97] Panasonic Ltd., “Photovoltaic module HIT® VBHN330SJ47/VBHN325SJ47.” Panasonic Eco
Solutions Europe, Ottobrunn, Germany, 2017.
[98] S. Rodrigues, R. Torabikalaki, F. Faria, N. Cafôfo, X. Chen, A. R. Ivaki, H. Mata-Lima, and F.
Morgado-Dias, “Economic feasibility analysis of small scale PV systems in different countries,”
Sol. Energy, vol. 131, pp. 81–95, 2016.
[99] RES LEGAL, “Promotion in Netherlands,” 2016. [Online]. Available: http://www.res-
legal.eu/search-by-country/netherlands/tools-list/c/netherlands/s/res-
e/t/promotion/sum/172/lpid/171/. [Accessed: 02-Jan-2017].
[100] SolarEdge, “SolarEdge Limited Product Warranty.” SolarEdge, Fremont, California, USA, 2016.
[101] D. C. Jordan and S. R. Kurtz, “Photovoltaic Degradation Rates-an Analytical Review,” Prog.
Photovoltaics Res. Appl., vol. 21, no. 1, pp. 12–29, Jan. 2013.
[102] NRC, “Dure regeling zonnepanelen gaat op de schop. Wat nu?,” 2018. .
[103] CBS, “Aardgas en elektriciteit, gemiddelde prijzen van eindverbruikers,” 2017. [Online].
Available:

143
http://statline.cbs.nl/Statweb/publication/?DM=SLNL&PA=81309NED&D1=8&D2=0&D3=0&D
4=19,24,29,34,39,44,49&HDR=T&STB=G2,G3,G1&CHARTTYPE=3&VW=T. [Accessed: 18-Feb-
2018].
[104] CBS, “Energieverbruik particuliere woningen; woningtype en regio’s,” 2016. [Online].
Available:
http://statline.cbs.nl/Statweb/publication/?DM=SLNL&PA=81528NED&D1=1&D2=a&D3=0,5-
16&D4=4-5&VW=T. [Accessed: 04-Jan-2017].
[105] RVO, “Saldering voor kleinverbruikers,” 2016. [Online]. Available:
http://www.rvo.nl/onderwerpen/duurzaam-ondernemen/duurzame-energie-
opwekken/duurzame-energie/saldering-en-zelflevering/saldering-voor-kleinverbruikers.
[Accessed: 02-Jan-2017].
[106] RVO, “Stimulering Duurzame Energieproductie (SDE+),” 2016. [Online]. Available:
http://www.rvo.nl/subsidies-regelingen/stimulering-duurzame-energieproductie-sde.
[Accessed: 02-Jan-2017].
[107] M. Verbič, S. Filipović, and M. Radovanović, “Electricity prices and energy intensity in Europe,”
Util. Policy, vol. 47, pp. 58–68, Aug. 2017.
[108] E. Trujillo-Baute, P. del Río, and P. Mir-Artigues, “Analysing the impact of renewable energy
regulation on retail electricity prices,” Energy Policy, vol. 114, pp. 153–164, Mar. 2018.
[109] K. Schoots, M. Hekkenberg, and P. Hammingh, “Nationale Energieverkenning 2017,” Petten,
the Netherlands, 2017.
[110] SolarPower Europe, “Global Market Outlook for Solar Power / 2016 - 2020,” Brussels,
Belgium, 2016.

144
9 APPENDIX
9.1 SELECTION OF PHP SCRIPTS

Two examples of the PHP scripts written for the PVP 2.0 website are shown here: a set of
functions to dynamically update the climate database with weather measurements and the economic
model.

9.1.1 WEATHER AND CLIMATE DATABASE SCRIPT


// In this function, the a weighted average of the real-time and
historic data is made, and the count values are increased by the
average amount of real-time measurements taken in the hour timeslot.
function
DataUpdate($real_parameter,$real_count,$history_parameter,$history_co
unt){
$parameter_array =
array('Irradiation','T_ambient','T_ground','Wind','Cloud','Pressure',
'Diffuse','Direct','Elevation','A_s');
$count_array =
array('IrradiationCount','T_ambientCount','T_groundCount','WindCount'
,'CloudCount','PressureCount','DiffuseCount','DirectCount','Elevation
Count','A_sCount');

for($i=0;$i<sizeof($parameter_array);$i++){ // We loop
through each value and assign a weighted average.
$history_parameter[$parameter_array[$i]] =
round((($history_parameter[$parameter_array[$i]]*$history_count[$coun
t_array[$i]]) +
($real_parameter[$parameter_array[$i]]*$real_count[$count_array[$i]])
)/($history_count[$count_array[$i]] +
$real_count[$count_array[$i]]),3); // The weighted weather
parameter value is calculated.
$history_count[$count_array[$i]] =
$history_count[$count_array[$i]] + $real_count[$count_array[$i]];
// The measurement counts are added to form the new count value.
}

$final_array = array($history_parameter,$history_count);
// The new parameter and count values are uploaded.
return $final_array;
}

function
DataUpload($end_time,$table_name,$final_upload,$data_type){
include '/data/www/vhosts/pvportal-
2.ewi.tudelft.nl/httpdocs/MySQL/Hisconnect.inc.php'; // A
connection is made to the historic database.

if($data_type == 'Weather'){
$fetch_array =
array('Irradiation','T_ambient','T_ground','Wind','Cloud','Pressure',
'Diffuse','Direct','Elevation','A_s'); // The MySQL table column
names.
}elseif($data_type == 'Count'){
$fetch_array =
array('IrradiationCount','T_ambientCount','T_groundCount','WindCount'

145
,'CloudCount','PressureCount','DiffuseCount','DirectCount','Elevation
Count','A_sCount');
}
$database_time = '2000-'.date('m-d H',strtotime($end_time.' -
1 hour')).':30:00';
for($i=0;$i<sizeof($fetch_array);$i++){
$final_value = $final_upload[$fetch_array[$i]];
$mysqli_query = "UPDATE $data_type$table_name SET
$fetch_array[$i] = $final_value WHERE Local = '$database_time'";
$query_run = mysqli_query($link,$mysqli_query);
}
}

9.1.2 ECONOMIC MODEL


// The economic model function, calculating four economic factors for
two discount rate scenarios.
function
EconomicModel($type,$total_production,$capacity,$module_choice,$ISF){
include '/data/www/vhosts/pvportal-
2.ewi.tudelft.nl/httpdocs/ModelInfo/ModuleInfo/'.$module_choice.'.php
';

if($type == 'BAPV'){
if($capacity<=10E3){ // The attributes of the
residential BAPV category
$stand_investment_spec = 1.45; // The standard
investment cost, which is adjusted by the panel choice and inverter
sizing factor.
$stand_panel_spec = 0.515; // The cost of a mono-Si
Trina Solar panel, chosen as the standard panel.
$stand_inverter_spec = 0.24; // The standard cost
of a fully sized inverter.
$OM_percent = 0.005; // The operation and
maintenance cost per year as a fraction of the initial investment.
$electricity_revenue = 0.181; // The value of one
kWh of electricity in €.
}elseif($capacity>10E3){ // BAPV systems with
capacities above 10 kWp are the category of commercial BAPV
$stand_investment_spec = 1.03;
$stand_panel_spec = 0.515;
$stand_inverter_spec = 0.15;
$OM_percent = 0.01;
$electricity_revenue = 0.105;
}
}elseif($type == 'free-standing'){ // The attributes of the
commercial free-standing PV category.
$stand_investment_spec = 0.97;
$stand_panel_spec = 0.515;
$stand_inverter_spec = 0.10;
$OM_percent = 0.015;
$electricity_revenue = 0.105;
}
// The total installation costs are calculated based on the user
choices.
$choice_panel_spec = $Wp_price; // The cost of the actual user
panel choice.
$choice_inverter_spec = $stand_inverter_spec/($ISF/100);
// The effect of inverter downsizing by the ISF on inverter cost.

146
$install_BoS_spec = $stand_investment_spec - $stand_panel_spec -
$stand_inverter_spec; // The contribution of installation and
balance of systems (BoS) costs to the total investment.
$choice_investment_spec = $install_BoS_spec + $choice_panel_spec
+ $choice_inverter_spec; // The final specific initial investment
taking the user choices into account.
$OM_spec = $OM_percent*$choice_investment_spec; // The
specific O&M costs per year.
$invest_cost = $choice_investment_spec*$capacity; //Total
installation costs
$OM_cost = $OM_spec *$capacity; //Annual fixed cost in €

// The standard economic factors used for all systems.


$lifetime = 25; // System lifetime in years
$panel_degradation = 0.007; // Panel degradation rate per
year.
$discount_low = 0.03; // The real interest rate, i.e.
discount rate, per year for the low scenario.
$discount_high = 0.07; // The real interest rate, i.e.
discount rate, per year for the high scenario.
$electricity_increase = 0.03; // Electricity price
inflation per year, based on the ECN prediction.

//Calculation of annual net present value of the system. The for


loop calculates cumulative savings for each year.
$NPV['nd'][0] = -$invest_cost; // A non-discounted net
present value array is initiated in order to calculate the non-
discounted payback period.
$NPV['low'][0] = -$invest_cost; // The net present value
array is initiated for the low interest scenario.
$NPV['high'][0] = -$invest_cost; // The net present value
array is initiated for the high interest scenario.

$total_revenue['nd'][0] = 0; // An additional section


is added for the non-discounted revenue, which is necessary to
determine the CAGR.
$total_revenue['low'][0] = 0; // The total discounted
revenue array is initiated for the low interest scenario.
$total_revenue['high'][0] = 0; // The total discounted
revenue array is initiated for the high interest scenario.

$total_cost['nd'][0] = $invest_cost; // An additional section


is added for the non-discounted cost, necessary to determine the
CAGR.
$total_cost['low'][0] = $invest_cost; // The total discounted
cost array is initiated for the low interest scenario.
$total_cost['high'][0] = $invest_cost; // The total discounted
cost array is initiated for the low interest scenario.

for($i=1;$i<=$lifetime;$i++){
$total_production_actual[$i] = $total_production * pow((1-
$panel_degradation),$i); // The annual production including
performance degradation.
$electricity_revenue_actual[$i] = $electricity_revenue *
pow((1+$electricity_increase),$i); // The electricity price in
that year.

$total_revenue['nd'][$i] =
($total_production_actual[$i]/1000)*$electricity_revenue_actual[$i];
// The non-discounted revenue of the system.

147
$total_revenue['low'][$i] =
(($total_production_actual[$i]/1000)*$electricity_revenue_actual[$i])
/pow(1+$discount_low,$i); // Discounted revenue from electricity
production (low scenario).
$total_revenue['high'][$i] =
(($total_production_actual[$i]/1000)*$electricity_revenue_actual[$i])
/pow(1+$discount_high,$i); // Discounted revenue from electricity
production (high scenario).

if($i==12){ // In year 12, the inverter is replaced,


adding a one-time additional cost to the system.
$total_cost['nd'][$i] = $OM_cost +
$choice_inverter_spec*$capacity; // Non-discounted O&M cost.
$total_cost['low'][$i] = ($OM_cost +
$choice_inverter_spec*$capacity)/pow(1+$discount_low,$i); //
Discounted O&M cost (low scenario).
$total_cost['high'][$i] = ($OM_cost +
$choice_inverter_spec*$capacity)/pow(1+$discount_high,$i); //
Discounted O&M cost (high scenario).
}else{
$total_cost['nd'][$i] = $OM_cost;
$total_cost['low'][$i] =
$OM_cost/pow(1+$discount_low,$i);
$total_cost['high'][$i] =
$OM_cost/pow(1+$discount_high,$i);
}

$NPV['nd'][$i] = $NPV['nd'][$i-1] + ($total_revenue['nd'][$i]


- $total_cost['nd'][$i]); // The non-discounted NPV, which is
just the total profit made up until year i. Note that this is not
actually a NPV, as the NPV, by definition, includes the time value of
money.
$NPV['low'][$i] = $NPV['low'][$i-1] +
($total_revenue['low'][$i] - $total_cost['low'][$i]); // The NPV
in the low scenario.
$NPV['high'][$i] = $NPV['high'][$i-1] +
($total_revenue['high'][$i] - $total_cost['high'][$i]); // The NPV in
the high scenario.
}

// The normal and discounted payback periods.

for($n=0; $n<count($NPV['nd']); $n++){


if($NPV['nd'][$n]>=0){
$PBP['nd'] = $n;
break;
}
}

for($n=0; $n<count($NPV['low']); $n++){


if($NPV['low'][$n]>=0){
$PBP['low'] = $n;
break;
}
}

for($n=0; $n<count($NPV['high']); $n++){


if($NPV['high'][$n]>=0){
$PBP['high'] = $n;
break;

148
}
}

// The compound annual growth rate. The revenue and cost are not
discounted because the CAGR should be compared with the base interest
rate, i.e. the discount rate itself.
$CAGR =
(pow(array_sum($total_revenue['nd'])/array_sum($total_cost['nd']),(1/
$lifetime)))-1;

//Levelised Cost of Electricity


$LCoE['low'] =
array_sum($total_cost['low'])/(array_sum($total_production_actual)/10
00);
$LCoE['high'] =
array_sum($total_cost['high'])/(array_sum($total_production_actual)/1
000);

$returnarray =
array($NPV,$PBP,$CAGR,$LCoE,$total_revenue,$total_cost);
return($returnarray);
}

149
9.2 MISSING WEATHER PARAMETERS

Table 25: Absent measurement parameters for each KNMI weather station and the stations that
supplied replacement data. Station 323 was not included as one of the final 46 stations as it was
discontinued in 2014, but was used to provide replacement data for 5 stations.
Station Wind Cloud Pressure Rain Irradiance Tambient Tground Source

209 (Ijmond) - - - - - - 240

225 (Ijmuiden) - - - 257

242 (Vlieland) - - 235

248 (Wijdenes) - - - - - - 269

249 (Berkhout) - - 235

257 (Wijk aan Zee) - - - 225

258 (Houtribdijk) - - - - - - 269

267 (Stavoren) - 270

273 (Marknesse) - - 269

277 (Lauwersoog) - - 279

278 (Heino) - - 279

283 (Hupsel) - - 290

285 (Huibertgat) - - - - - - 280

286 (Nieuw Beerta) - - 280

308 (Cadzand) - - - - - - 310

312 (Oosterschelde) - - - - - - 310

313 (Vlakte v.d. Raan) - - - - - - 310

315 (Hansweert) - - - - - - 323

316 (Schaar) - - - - - - 323

324 (Stavenisse) - - - - - - 323

331 (Tholen) - - - - - - 323

340 (Woensdrecht) - - 323

343 (R'dam Geulhaven) - - - - - 344

348 (Cabauw) - - 356

377 (Ell) - 380

391 (Arcen) - - 375

150
9.3 SUMMARY OF COMMUNICATION WITH STEDIN AND TENNET

9.3.1 TELEPHONE CALL WITH MARKO KRUITHOF AT STEDIN (25/11/2016)

The question of how much solar energy is being produced in the Netherlands in real-time or
annually is not an easy one. There are several reasons for this. Firstly, not all PV systems are being
registered by their respective owners. This is because they are not legally obliged to do so, and in
some cases panels have been acquired on the black market in which case registering does not make
sense for the owner. A second reason for incomplete data on total production is that self-
consumption of solar energy by households occurs, which leads to the fact that not all produced
energy is measured by a smart meter. Finally, some households have multiple energy production
installations. If these installations are all connected to the same grid entry point, it is impossible for a
TSO to distinguish which fraction of produced energy comes from solar energy.

As there is no registration requirement for PV systems, the most reliable data on installed
capacity comes from the Tax and Customs Administration which measures the number of panels
being imported into the Netherlands. The assumption is made that all panels that are not exported
elsewhere are installed. A factor that motivates system owners to register is that it allows them to
apply for subsidies. Mr. Kruithof indicated that he thinks about 80% of PV systems are registered on
the national level. Stedin has toyed with ideas to fly over areas to map all PV systems on roofs, and via
location find out which systems are registered and which are not.

For TSOs, the most common method for solar energy production estimation is through models.
These models use weather data, an estimate of installed capacity, and information on the relation
between weather and power production of a PV system to output the predicted power output for the
given weather.

A few projects exist which use real-time data on electricity production gathered by smart meters
in electrical substations, such as in the neighbourhood Lombok in Utrecht. It is useful for a TSO to
connect the smart meter data to data from other sources such as SCADA (supervisory control and
data acquisition) to calculate what behaviour will occur within the system. However, employing smart
meters at a large set of locations requires a large investment for the TSO, which at the moment does
not seem necessary to ensure network stability. The tactic Stedin employs at the moment is to check
whether the installed capacity i.e. the Wp installed is within the safety margin.

9.3.2 TELEPHONE CALL WITH FIEKE ‘T HOEN AT TENNET (18/01/2017)

TenneT is obliged by European law to provide predictions of solar energy production to ENTSO-E
(the representative organization of all European TSOs). As actual (real-time) production is not
measured or otherwise established by TenneT, the predicted solar electricity generation is equated to
the actual, realized, generation. Because the penetration level of solar energy is so low in the
Netherlands, PV electricity has little to no effect on the high-voltage network that is managed by
TenneT. In other words, they do not have an incentive to make the projections of solar energy
production as accurate as possible. TenneT do not make the predictions themselves, but outsource it
to a weather prediction company. The name of this company was not disclosed to me as this
information relates to their business case. Due to protection of company information, Ms. ‘t Hoen
was not at liberty to inform me of the exact methodology the weather company uses for the
production forecast. However, she mentioned several aspects that are accounted for in the prediction
model. The company uses an estimate of installed capacity. The value of this capacity is updated
several times a year and TenneT is notified of the installed capacity used in the predictions when the

151
forecast data are sent to TenneT. The installed capacity is coupled with weather predictions retrieved
from a number of other companies to arrive at power production. An estimate of regional distribution
of the PV system capacity is used together with regional weather data to ensure that local irradiance
changes do not affect the national forecast to a large extent.

Ms. ‘t Hoen mentioned the projections made by the German branch of TenneT as another source
of information. As Germany has a much larger installed PV capacity than the Netherlands (~40 GW at
the end of 2015 [110]), forecasts are of higher importance. Interestingly enough, TenneT found that
even though installed solar capacity amounts to 50% of the German base load, there are no large
stability issues. This is attributed to the fact that on sunny and completely cloudy days, power output
of PV systems changes very gradually or remains relatively constant. On partly cloudy days, the
impact of fluctuations on the high-voltage grid is low due to the large spatial spread of the systems.
The German branch uses information from 4-5 weather services. Predictions from the previous month
are evaluated to see whether they are realistic and if adjustments to the applied model should be
made. Two interesting and surprising weather influences they found were the effect of snow and of
Sahara sand. When there has been snowfall, a large portion of panels may be covered in snow. Even if
solar irradiance is relatively high, output from these systems will be insignificant. Likewise, when
strong winds blow sand from the Sahara over Europe, panels can be coated with this sand leading to a
lower system performance until they are cleaned or until rainfall occurs.

In TenneT’s view, the responsibility for the prediction of solar energy production should lie with
the energy companies that sell electricity on the market. As they are obliged to provide 15-min.
predictions for the power generation of their entire portfolio, it would make sense to have them
estimate the solar energy production of their customers as well. However, actual feed-in is not
measured by these companies. The energy companies use EDSN-profiles (standard consumption
profiles) for all customers. As neither real production nor consumption is measured, the feed-in into
the electricity network is unknown. In the future this may change with smart energy meters.

Ms. ‘t Hoen referred me to several other organizations for further research. TenneT does not tally
the provincial or regional total energy production in real-time or otherwise. The real-time penetration
rate of solar energy can therefore not be established with TenneT data. Alliander, Enexis, Stedin, and
Enduris may have more insight into provincial energy production as the four largest regional DNOs.
Alliander may be a good option to find out more about solar energy production, as the organization
had issues with the large increase in solar panels in Groningen after the compensation scheme for
earthquakes. A solar energy prediction company called iCarus has been contacting TenneT offering
their services. They state that they can provide a very accurate prediction of solar energy production.
Icarus is an initiative of Alliander and their website can be found at http://icarus.energy/.

152

You might also like