Вы находитесь на странице: 1из 38

Energy Conversion and Energy Storage Devices

Fundamentals and Applications

Philip I. Cohen and Paul J. Imbertson

January 29, 2019


Contents
1 Introduction 4
1.1 Sources for further investigation . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.2 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
1.3 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2 Photovoltaic Solar Cells 9


2.1 The Solar Spectrum . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
2.2 Overview of Current Technologies . . . . . . . . . . . . . . . . . . . . . . . . . . 12
2.3 Basic photovoltaic mechanism . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2.4 Brief Review of Selected Semiconductor Properties . . . . . . . . . . . . . . . . . 15
2.4.1 Energy bands . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
2.4.2 Optical Absorption . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.4.3 Carrier densities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.4.4 Minority carrier lifetime . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.4.5 Diffusion of carriers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.5 Diodes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
2.5.1 Photovoltaic diode: the simple qualitative picture . . . . . . . . . . . . . . 22
2.5.2 The diode: simple quantitative picture . . . . . . . . . . . . . . . . . . . . 22
2.5.3 Uniformly illuminated long diode (optional) . . . . . . . . . . . . . . . . . 23
2.5.4 Diode with exponential absorption (optional) . . . . . . . . . . . . . . . . 25
2.5.5 Band Diagrams (optional) . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.6 Power and efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.6.1 Maximum Power . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 26
2.6.2 Efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
2.6.3 Maximum Power Point Tracking . . . . . . . . . . . . . . . . . . . . . . . 29
2.6.4 Losses at surfaces . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.6.5 Summary . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
2.7 Losses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.8 Recent Innovations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.8.1 Si Cell . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
2.8.2 PERT cell . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.8.3 Half Cell . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.8.4 Bifacial Cell . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.9 Beyond Silicon . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.9.1 Heterostructure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
2.9.2 CdTe . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
2.9.3 Perovksites . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
2.10 Next Generation Structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
2.11 Concentrator Designs . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34
2.12 Exercises . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
2.13 References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

2
Contents

http://xkcd.com/556/

3
1 Introduction
The current mix of energy consumption in the U.S. is dominated by liquid fuels and natural
gas. The share of coal, once dominant and a large carbon polluter, has decreased significantly with
coal-fired plants converted to natural gas. Fig. 1a and b show two views of this mix of fuel from the
US Energy Information Administration (USEIA). In 1a the mix is plotted vs time, with the current
state near the vertical dashed line. For future times a model is used, in this case based on known
developments of current technology as well as predictions of future economics, termed a reference
model by the USEIA. In their annual energy report other scenarios are examined [1] but we limit
ourselves to this. One surprise is how little renewables are contributing to the current energy mix

Fig. 1.1: (a) Energy consumption by fuel in the U.S. for a reference model. Here 45 quads or 45 × 1015 BTU is 47.5 × 1018 J or
1.32 × 1013 kWh. (b) Flow chart showing both input and output, including waste heat.

despite the public attention that they are given. The publicity given to solar is probably reflected
in the large relative increase in solar capacity indicated in the plot as compared to other sources.
Another is that in the future, wind energy is expected to level off, allowing solar to dramatically
surpass its contribution. It is not until about 2050 that solar reaches coal which is not expected to
change much from its current level. For this long term picture, it is expected that photovoltaics will
dominate over solar thermal in which the heat from solar radiation is used to drive a steam engine.
Solar thermal represents a small portion of renewables currently and is expected to be relatively
more expensive in the future. In Fig. 1b, the picture describing both the inputs and outputs of
only a few years ago, the biggest surprise is how much energy is wasted. It is tempting to think
that with more efficient designs this waste heat could be better utilized, but there are fundamental
thermodynamic limits that must be considered. This is an important question – how much time
and effort should be expended to make power plant or a solar cell more efficient. Thermodynamic
Carnot efficiencies and the Schockley-Quiesser Limit will be discussed later.
According to the U.S. Department of Energy [2], the world uses about 4.1 × 1020 J/yr or
about 13 × 1012 W (13 TW) in continuous use. This is projected to be about 30 TW by 2050 but
due to green house considerations this needs to be reduced to about 10 TW. By comparison the
Sun deposits about 120,000 TW on the earth on average. Assuming 10% efficient solar farms and
using 0.16% of the available land (in the U.S. this would be about 10× the area of single-family
rooftops or about the area covered by the nation’s highways), one would obtain about 20 TW. For

4
1 Introduction

comparison hydroelectric can produce about 0.5 TW and wind turbines can obtain about 2-4 TW.
Nuclear is probably not a long term answer. As the DOE puts it, while wind and other sources will
contribute, solar is the only large number available [2]. With this in mind our focus will be on the
operation and application of solar energy.
Once generated solar power must be transmitted to the end user. Locally generated power
from solar cells obviously has some advantage in this regard since transmission losses scale with
distance. By contrast, utility scale generation in solar farms, even those that must be transmitted
over long distances, has significant advantages of economies of scale and can use more intense
light that is available without complications of variable cloud cover. The rapid increase in solar
energy generation discussed before and the grid necessary to distribute it is illustrated in Figs. 1.2
and 1.3. The main contributions are from the U.S. southwest, The eastern seaboard, Texas, and
Minnesota.
A real problem is that solar energy production is maximum at mid day, causing the demand
to power companies to be reduced. Coal and nuclear only make economic sense to operate when
near full capacity. If the demand is reduced then without some means of energy storage the grid
risks being overloaded and solar must be curtailed and wasted.
At this point there is not very much storage capacity. The grid that connects these sources to
users is designed to be balanced and this balance must be maintained at all times by appropriately
adding or removing generation capacity. Renewables make a relatively small contribution to this
energy distribution problem but it is notable that solar energy from the continental regions tend
to be maximum during the day while solar from the coast as well as wind energy are largest at
night [3]. This complementary variation will help the balance needed for this demand problem.
The U.S. grid has three separate networks, as shown in Fig. 1.2 with Western, Eastern, and Texas
subdivisions. Each sector delivers power at 60 Hz but they are not synchronized with each other.
One can see in Fig. 1.2 that there are a few high voltage dc connections between the sectors. To

Fig. 1.2: U.S. existing and proposed electric grid from


www.NPR.org Fig. 1.3: Existing and planned solar generating utilities (from
SEIA and Argonne National Laboratory, 2018,
. www.seia.org, http://solarprojects.anl.gov/).

minimize losses, high voltage transmission, perhaps 100 kV, is preferred. This means that the
output from a 1/2 V photovoltaic cell must, through a combination of series connections and power
electronics, be transformed to very high voltages.
To assist in balancing the grid and to make applications more convenient by separating gen-
eration from power output, some sort of storage is required. Large scale hydroelectric or pumped
storage dominates with about 22 GW of power capacity. Most new large scale storage is in the
form of Li ion batteries with a U.S. total of about 1 GW currently.
The topics to be discussed are based on these introductory comments. In regard to power

5
1 Introduction

generation we will focus on solar photovoltaics as it is expected to be the dominant renewable by


2050 and give somewhat less emphasis to wind power. We will discuss the power grid and the
requirements it places on load leveling and interconnections. The power electronics necessary to
obtain optimal use of photovoltaic devices and transmission over the grid will be a key issue for
both large and small applications. Concerning storage, the physics, chemistry, and operation of Li
ion batteries will be discussed, limiting ourselves to these since they are pervasive and simple. To
get a sense of the limitations and capabilities of solar and the capacities of these batteries we will
discuss applications to electric vehicles, including cars, bicycles and scooters.
Thus our focus will be on energy conversion: starting from some fundamental form, perhaps
solar or chemical, we will examine the engineering requirements and methods to convert to elec-
tricity. We will take an engineering view in this, worrying about fundamental concepts only when
trying to determine limits to device efficiencies. In fact energy itself is a poorly defined notion. A
very well known scientist, Richard Feynman, described the issue as: “It is important to realize that
in physics today, we have no knowledge of what energy is. We do not have a picture that energy
comes in little blobs of a definite amount. It is not that way. However, there are formulas for
calculating some numerical quantity, and when we add it all together it . . . always gives the same
number. It is an abstract thing in that it does not tell us the mechanism or the reasons for the various
formulas.” (from ‘Lectures on Physics’). We will just use the formulas.
Finally, though fundamentally crucial, it is difficult to talk much more about the economics
of various energy sources. However two concepts are useful. First, the levelized cost of energy
(LCOE) is an estimate of the average minimum cost that energy should be sold over the lifetime of
the generator in order to break even. It looks at the average cost to build and operate the generator
and the amount of energy that it generates. It is usually given in $/kWh and is probably the best
pricing method. Second, installed costs or module costs are usually given in $/W and represent the
cost to the installer without considering maintenance or long term costs. Assumptions on discount
rates and operating and maintenance need to be made to determine the LCOE.
For example some system costs and LCOE costs of PV as determined by NREL:

Fig. 1.4: System cost of PV components by NREL in 2017 from the perspective of an installer, including installer profits and
averaged over the US weighted by installed capacities [4].

6
1 Introduction

Fig. 1.5: LCOE NREL 2017 [4].

1.1 Sources for further investigation

1. https://www.pveducation.org/
2. https://batteryuniversity.com/
3. http://ecee.colorado.edu/ bart/book/welcome.htm Bart Van Zeghbroeck, “Principles of Semi-
conductor Devices,” which is freely available on line and at the level of the excellent text
“Semiconductor Device Fundamentals,” by R.F. Pierret.
4. https://www.eia.gov/ The U.S. Energy Information Administration is part of the U.S. De-
partment of Energy. In its authorization, Congress explicitly exempted the administrator
from requiring approval from any other officer or employee of the U.S. government prior to
publication of EIA’s analyses.
5. https://www.nrel.gov/research/data-tools.html is a useful starting point to examine the site
sponsored by the National Renewable Energy Laboratory.

1.2 Exercises

1. Should I buy a 100 W Sunpower solar cell (Amazon $218.99 (Jan, 2019)) and put it on my
roof (assume no other costs or rebates) or should I connect to the grid at $0.10/kWh ?
2. The Luddington resevoir in Michigan was originally intended as a utility scale electric stor-
age facility. It is now used to take advantage of the low cost of electricity at night and the
high cost during the day. At this facility, water is pumped 111 m up to a resevoir above
Lake Michigan. It can hold 27 billion gallons of water and can achieve a maximum flow
of 33 million gal/min to run a turbine generator. Assume that the efficiency of the turbine
generator and of the pump needed to fill the resevoir are both η = 90 %.
a) What is the maximum power output?
b) If daytime electricity is sold for $0.08/kWh, what night time electricity cost will allow
a profit?

7
3. (a) What force (in Newtons) is required to lift one liter of water? (b) What energy in Joules
is required to lift one liter of water 1000 ft?
4. The sun can provide about 1000 W/m2 at peak. Estimate the power one could expect to be
able to use to power a typical car or a typical house (roofs only). Assume that amorphous
silicon solar cells have an efficiency of 20% and that there are no other losses.
5. What are the main power requirements of a typical house, breaking this down to water heater,
home heating, refrigerator, computers, . . . ?
6. Download the solar spectrum for AM1.5 (google AM1.5 and xls). Use the “global tilt” data.
Note that the wavelengths might not be equally spaced in the downloaded data.
a) Plot the number of photons per unit wavelength per second striking the earth per square
meter vs wavelength.
b) At what wavelengths are (a) the number of photons maximum and (b) photon energy
maximum. Compare to the absorption spectrum of chlorophyll.
c) From the “Global tilt” data, what is the total power per sq m?
7. Assume that the top surface of a Honda Civic is covered with a 20% efficient solar panel.
How much power can be provided at peak sun? How much power is needed to operate the
car (assume 40 mpg and go 1 mile in 1 min with an efficiency of 30%)?
8. How many photons are there per second per sq m with an energy > 1.12 eV? If each photon
produced an electron, what current would this correspond to?
9. A flywheel consisting of a cylindrical shell with a 1m radius, R, spins at 20,000 rpm. The
rotational energy is (1/2)Iω 2 where I = mR2 . For an angular velocity of 1500 rpm, what
diameter flywheel is needed to store 5 W·hr/kg.
10. The U.S. generated 4.1 × 1015 Wh of electricity in 2011. Use a tilted flat plate solar cell with
an efficiency of 20% to generate an equal amount of electricity from the sun. What fraction
of Utah (220,000 km2 ) would be needed. You can use the graph of the incident incident solar
energy in NREL annual map jpg.
11. You are asked to construct a photovoltaic cell that only uses light over a very narrow wave-
length interval, say 0.01 nm (different parts of the device will use different wavelength inter-
vals, but for this question just consider one interval). Assume that every photon is absorbed,
that each absorbed photon creates one electron, and perfect charge collection efficiency. Us-
ing the irradiance data in the graph below:
a) What is the current density at 1500 nm?
b) At which wavelength is the current density maximum?

1.3 References
[1] U.S. Energy Information Administration. Annual Energy Outlook. Tech. rep., Department of
Energy (2018).
[2] J. Horwitz. Solar energy challenges and opportunities. Tech. rep., Department of Energy
(2007).
[3] R. Fares. Renewable Energy Intermittency Explained: Challenges, Solutions, and Opportuni-
ties. Sci. Amer. .
[4] R. Fu, D. Feldman, R. Margolis, M. Woodhouse, and K. Ardani. US Solar Photovoltaic Sys-
tem Cost Benchmark: Q1 2017, NREL/TP-6A20-68925. Tech. Rep. NREL/TP-6A20-68925,
National Renewable Energy Laboratory (2017).

8
2 Photovoltaic Solar Cells
A. Becquerel discovered the photovoltaic effect in 1839. Around 1890, C. Fritts coated Se
with a transparent layer of gold to obtain a cell with 1% efficiency. The first Si cell was described
by R.S. Ohl at Bell labs in a patent in 1941 and then further developed there by Pearson, Chapin,
and Fuller in the early 1950’s, the latter with an efficiency of about 6% [1]. In these cells sunlight
is converted into electricity by using the energy of the photons to produce electrons and holes.
Random motion of these free carriers in the region of a semiconducting pn junction prevents the
carriers from immediately recombining. If they are allowed to recombine by connecting the device
across a load, then the energy can be used to generate electrical power. Not all photons are absorbed
and not all that is absorbed can be recovered. The early history is described by Green [2].

2.1 The Solar Spectrum

Any material or body at a temperature T emits radiation (photons). If the size and shape
of the material satisfy certain constraints then the intensity of light that is emitted vs wavelength,
a spectrum of light, will depend only on the temperature. This spectrum is called the black body
spectrum and its shape depends only on the temperature. The main concept here is that in any
material, held at a temperature T , there is some probability that an electron orbiting an atom nucleus
will absorb some thermal energy and be excited into a higher energy state; subsequently it will relax
back to a lower energy state with the possibility of emitting a photon. If this were the only process,
the spectrum would consist of relatively sharp atomic lines, specific to the particular material. But
for a solid body the density is high enough that the light is reabsorbed before leaving the cavity and
then re-emitted, perhaps repeating multiple times. An atomic absorption line is not perfectly sharp.
The absorption probability is largest if the photon energy matches the center of the transition, but
it occur if the photon energy is somewhat off this energy as well. It is just less likely. Similarly in
the emission process, with the most likely photon emitted at the center frequency of the transition.
With enough transitions in the material, the photons become “thermalized” and all materials at the
temperature T emit the same distribution of light. Photons at the center frequency of a transition
are easily absorbed; then re-emit, usually at a different frequency. Photons off the center are not
absorbed easily and are mainly transmitted. This provides a mechanism for all materials to emit
the same spectrum. This idealized emission is called black body radiation. Though the Sun’s
temperature is not uniform throughout its mass, its surface acts as a black body with a characteristic
temperature of 5760K. Further, it appears as a disc that subtends an angle of θs = 0.26◦ at the earth
and so the radiation that reaches the earth is check Eq and plots and y values; check units – I think
# density)

E3
 
2Fs
b(E) = 3 2 (2.1)
h c eE/kTs − 1

where Fs = π sin2 θs reduces the fraction that reaches the earth. This is the light intensity (W/m2)
reaching the Earth surface per unit energy. It is sometimes more convenient to write this as a

9
2 Photovoltaic Solar Cells

function of wavelength; in that case the light intensity per unit wavelength is given by

2Fs hc2 1
b(λ ) = 5 hc/kT
(2.2)
λ e sλ − 1

where one can see the conversion since the integral over energy must equal the integral over wave-
length. Eqs. 2.1 and 2.2 are plottled as the smooth curves in Fig. 2.1. For comparison, measured,
average data, to be discussed next, is also plotted.

1.6 ASTM Direct+Circumsolar 600 ASTM Direct+Circumsolar


1.4 500
Irradiance (Wm−2 nm−1 )

Irradiance (Wm−2 eV−1 )


1.2
1.0 400
0.8 Integral = 900 300
0.6 200
0.4
0.2 100
0.00 500 1000 1500 2000 2500 3000 00.0 0.5 1.0 1.5 2.0 2.5 3.0 3.5 4.0 4.5
λ (nm) Energy (eV)
Fig. 2.1: Solar irradiance at the earth’s surface. These two plots show the same data: The plot vs λ is reversed from that vs
energy (low wavelength corresponds to high energy) and the increments on the abscissa are not equal. These data are
from ASTM G173-03 Reference Spectra Derived from SMARTS v. 2.9.2 and are discussed later in this section. Both
plots integrate to 900 W/m2 .

Note that the units on the ordinate (y-axis) of Fig. 2.1(left) are a bit odd. It is written as
Wm−2 nm−1 rather than Wm−3 . This form is used because each data point corresponds to W per
area of surface for wavelengths in the interval (λ , λ + dλ ). Integrating over dλ yields W/m−2 . If
the measured, average data is available it is usually simple to use it for a numerical calculation, but
depending on the analysis, it might be more convenient to use this black body approximation to the
spectrum.
The amount of light available for PV varies throughout the year, during the day, and, of
course, by location. To maximize the amount of light collected one would need two axes of rotation
since the earth spins on an axis that is tilted relative to the plane of the ecliptic (Fig. 2.3).

23.5 o

o
48.2 AM1.5

Fig. 2.2: Two axes are required to track the sun


d
atmosphere x = d/cos θ = 1.5 d
throughout the day and year.

Fig. 2.3: Atmosphere.

The spectra shown in Fig. 2.1 is an average over the northern hemisphere (U.S. ?) referred
to as air mass 1.5 (AM1.5). In this case, at a given location on the earth, light incident on the upper

10
2 Photovoltaic Solar Cells

atmosphere (AM0) is attenuated by an amount exp (−αd/ cos θ ) = exp (−1.5αd) where α is an
attenuation coefficient, d is the effective thickness of the atmosphere, and θ = 48.2◦ is taken as an
average angle with respect to the normal.
Three distributions are given by an ASTM standard. First there is extraterrestrial radiation
which is the radiation at the top of the atmosphere with a total power of 1348 W/m2 . Second, useful
for concentrating solar cells there is light from the solar disk plus a small amount of radiation that
is scattered by the atmosphere, but at not too large an angle so that it can still be focused by optical
systems. The total power in this direct plus circumsolar radiation is 900 W/m2 . This is the spectrum
one should use for analyzing a concentrator system. Finally, there is radiation from the solar disk
plus scattered light and light reflected from the earth. Since this includes the diffuse light as well
as the direct light, it would be appropriate to use for modeling a titled, flat plate collector. The total
power in this “Global tilt” radiation is 1000 W/m2 . It is normalized to this value for convenience
(there is some arbitrariness here since it is an average over the continental U.S. over the day and
years).
Tracking the sun over the course of the year and during the day will cost energy and/or
maintenance; however it is feasible, and does not add substantially to the capital costs of a utility
scale system [3]. At noon, when the sun is closest to over head, the cell should be pointing south,
at an angle that depends on the latitude. NREL suggests tilting at the latitude for a fixed panel.
5
23.

Equ N )
ato 45
r P (
MS

ecliptic

Fig. 2.4: (a) Sketch showing the relevant angles of a solar cell with respect to the sun for summer solstice. Since the sun subtends
an angle of about 0.25◦ at the earth, the light incident is drawn horizontal. (b) Actual tilt angles of utility scale fixed-tilt
arrays in the U.S.

Others suggest different angles, perhaps to account for reflection from the ground and/or to opti-
mize energy collection over the course of the day. Fig. 2.4 shows that in the summer one should
tilt by 45-23.5 = 21.5◦ and in the winter one should tilt by 45+23.5 = 68.5◦ , assuming a latitude
corresponding to the Minneapolis airport. NREL’s recommendation for a fixed pannel is thus an
average of the two. Fig. 2.4(b) shows the fixed tilt, utility scale utilities in the U.S. and their tilt
angles, as indicated by the symbol diameter. They do not follow this recommendation, with most
tilted between 20 and 30◦ , tending to emphasize the summer sun [4].
The first question might be to determine which part of the solar spectrum should we try to
use. One way to approach this is to look at biological systems as a guide for manufactured devices.
Evolution has had hundreds of millions of years to try various chemistries. The most common
user of solar radiation are green plants in a process known as photosynthesis and perhaps we could
emulate these with manufactured systems.
According to Kiang [5] plant molecules have evolved to use two parts of the solar spec-
trum. The short wavelength portion of the spectrum provides energy rich photons while the long

11
2 Photovoltaic Solar Cells

wavelength portion provides large numbers of low energy photons. Fig. 2.6 shows the fraction
of light absorbed vs wavelength for a green plant as well as a cartoon of energy gathering in the
photosynthesis process. The high energy photons are downshifted in energy to enable use in the

Fig. 2.5: In photosynthesis, a large func-


tion of the plants is to act as Fig. 2.6: Green plants reflect or transmit the green part of the
an antenna, absorbing and then spectrum, making use of the energetic blue photons
downshifting and transferring so- and the numerous red.
lar energy.

photosynthesis reaction. The figure shows pigment molecules that down shift the photon to an
energy that can be utilized to split water to release electrons for other reactions. The energy of the
long wavelength photons is already suitable, though for a different reaction. It takes eight photons
for the photosynthetic reaction [5]. The green photons in between are insufficiently numerous and
do not have enough energy to make an adaption worthwhile. Not needing these, green light is
reflected giving rise to the colors of many plants.

2.2 Overview of Current Technologies

Which photovoltaic cell material or design that is most cost effective or most desirable by
whatever measure is not so obvious. Many material combinations have been suggested and exam-
ined, mostly guided by predictions of a theoretical maximum efficiency. The most straightforward
comparison of the status of different technologies is to show the development of efficiencies de-
vices fabricated in the various materials with different designs over time. This is a famous plot
collected by scientists at NREL, though unfortunately, containing too much information to be eas-
ily presented. Figs. 2.7 and 2.8 show the results of measurements that have been confirmed by
independent governmental agencies (for example NREL) on devices that had been submitted by
university, industrial, and other governmental laboratories. Fig. 2.7 shows the best efficiencies
for a few of the leading materials as of 2018. Fig. 2.8 shows the best efficiencies measured from
most technologies over time. Some of the important takeaways are: (1) Si, CdTe, and CuInGaSe2
(CIGS) are the leading materials that are economical and stable, (2) the efficiency of the various
technologies has increased slowly with time with the very notable exception of the perovskites,
(3) organics, dye-sensitized, and amorphous technologies, which are expected to be inexpensive to
produce, have relatively low efficiencies, and (4) a current high quality cell can have an efficiency
of 20-25%. The designs used in these various devices are very different.
The leaders in a few of the technologies as of 2018 are using Si, GaAs, CdTe, CuInGaSe2 ,
and perovskites. Apart from the perovksites, the technologies have appeared to have leveled off,

12
2 Photovoltaic Solar Cells

Fig. 2.8: NREL (D. Levi and others), figure via Wikipedia but
likely in a journal someplace; looking.

Fig. 2.7: Confirmed single junction AM1.5 cell efficiencies.


From Green et al. [6]

with percent level improvements considered major events. There is a clear trade off between cell
cost and efficiency. The least efficient cells are a-Si and organic thin films. But these are very
inexpensive. Amorphous Si is used for roof shingles and organics can be deposited using all of
the inexpensive techniques that chemical engineers have developed for plastics. Crystalline Si,
CdTe, and CIGS have an intermediate efficiency, with thin film materials having some advantage
in energy requirements. The most efficient are compound semiconductors, which however require
a concentrator to make cost effective. The main competition is between thin films devices and
crystalline silicon.
The two leading U.S. companies are Sunpower that make high performance single crys-
talline Si cells and First Solar that make similarly good CdTe cells. Chinese companies sell silicon
devices.
CdTe cells have some nice properties: Their light absorption is well matched to the solar
spectrum, they can be deposited at a very high rate, nearly 1 µm/s, grain boundaries do not en-
hance recombination of electrons and holes, they are direct bandgap semiconductors with high
absorptance, and the efficiency of current cells is relatively high. They can be deposited in a pro-
cess that does not require high vacuum systems, leading to low cost at manufacturing scales. They
suffer from the significant disadvantage that cadmium is toxic. Note that CdTe is stable and non-
toxic but tellurium is expensive and when it is recovered one must deal with the Cd waste. Further
CdTe cell efficiencies are not as high as one would like.
Silicon cells are the most established and are competitive with the thin film materials because
the technology is so well established in electronics. The material costs are high because one must
slice large single crystals. However, they have become very thin with less waste, reducing the
energy required to fabricate the devices.
It is always interesting to examine the availability of these somewhat exotic materials. The
figure shows the relative abundances of materials found in the earth’s crust Note that the ordinate is
a log scale. Thus though Cd is moderately abundant, Te is one of the rare materials and there is an
economic incentive to recover it. For a while there was concern that availability of these materials
was an issue, leading to rapid investment in CZTS technologies. This has become less of a concern.

13
2 Photovoltaic Solar Cells

Fig. 2.9: Relative abundances of materials in the earth’s crust from Wikipedia.

2.3 Basic photovoltaic mechanism

To understand the basic photovoltaic mechanism it might be easiest to consider the formation
of a very idealized Si pn-junction, with the initial joining shown in Fig. 2.10 and the depletion
region shown schematically in Fig. 2.11. Doped Si is formed by adding a very small amount

junction
V=Q/C

+
p and N A− n and N D p n

-xp xn
x=0 Fig. 2.11: Driven by the concentration gradient, holes and
Fig. 2.10: An abrupt, nonequilibrium, junction of n-Si and electrons will diffuse to opposite sides of the
p-Si. Initially both components consist of equal junction until the field in the depletion region
numbers of carriers and fixed ions. establishes a steady state.

of impurity atoms, say 1 ppm, which replace Si atoms in the Si lattice. These neutral impurity
atoms have either one more (donor) or one less (acceptor) electrons than Si. At room temperature
the electrons from the donors become mobile, leaving behind a positively ionized donor atom.
Similarly, the acceptors contribute what looks like a mobile hole and a negatively charged acceptor
ion. The ions are fixed in their lattice positions. We imagine placing a material that contains
only neutral acceptors (P-type) adjacent to a material that contains only neutral donors (N-type)
to form a junction is formed, as illustrated in Fig. 2.10, before any motion of carriers is allowed
to take place. This non-equilibrium junction has an abrupt change in the concentration of holes
and electrons at x = 0, which drives interdiffusion of these carriers but not the fixed ionized atoms.
After interdiffusion, in Fig. 2.11, the carriers have recombined and form a depletion region which
produces an electric field. In this depletion approximation picture (since it is not actually abrupt)
there are no free carriers in the depletion region. The built in voltage across the junction is Q/C
where Q = qND xn where Q and C are charge and capacitance per area.
Silicon has a small gap between the valence and conduction bands and it takes light that
has hν > Eg to be absorbed. The depletion region is small and most of the absorption takes place
in the neutral region. There electron-hole pairs (EHP) will be created that diffuse randomly. At

14
2 Photovoltaic Solar Cells

this point there is no current since the electron and hole move together to maintain neutrality. An
EHP that diffuses into the depletion region will be split by the field, giving rise to a current that
is in the same direction as the normal reverse current of a pn junction diode. Under open circuit
conditions a current in a direction opposite to this photocurrent must be generated in order to
balance it. To do this the device develops a forward bias voltage that injects carriers back across
the junction. In steady state open circuit, the currents balance. If there is a load, a smaller forward
bias is developed, just enough to allow whatever load current is required. One could also look at the
charge in the depletion region. At open circuit, the photogenerated carriers that become separated
by the depletion field will tend to reduce the depletion charge on either side of the junction. Since
the charge Q is reduced, the voltage across the junction is reduced and the a small forward bias
voltage appears. It the diode is shorted, then all of the carriers separated by the depletion field
contribute to the current, termed the short circuit current (assuming no losses). This rough and
qualitative explanation will be made quantitative in subsequent sections.

2.4 Brief Review of Selected Semiconductor Properties

2.4.1 Energy bands


An isolated atom such as silicon has 4 electrons that are involved in bonding to other atoms
in its outermost shell of energy orbitals or levels. The isolated atom also has empty levels at slightly
higher energies. It is possible for the atom to absorb energy, causing an electron to be excited into
one of these empty levels or even, if too much energy is absorbed, removed from the atom. When
many Si atoms are placed close together to form a solid, the electrons of different atoms interact
and the the energy levels shift slightly. They are so closely spaced in energy that it is more useful
to treat them as a continuous band of energy states. The same thing happens to the empty, higher
states. For Si the result is that there is a filled band, called the valence band, and an unfilled band,
called the conduction band, separated by a small energy gap. Since the valence band is full one
cannot move an electron from a filled state to a higher energy empty state within the same band.
This means that the valence band cannot contribute to electrical conduction since an electron that
remains in the band cannot be accelerated (energy increased). However there can be a few electrons
in the conduction band. This could happen if there is enough thermal energy to excite an electron
from the valence band to the conduction band or if there is an impurity atom in the lattice that
has a weakly bound electron that could be similarly be excited. These electrons in the conduction
band can contribute to conduction since there are many empty states in this band in which an
electron excited by an electric field can be placed. Similarly when the valence band is missing one
electron, the other electrons in the band can be thought of as if a positively charged hole is moved
by an electric field. This is called ambipolar transport. Finally, an electron in the valence band
can absorb energy from a photon and be excited from the valence band into the conduction band,
forming an electron hole pair (EHP). This last process underlies the basic optical absorption that
we require for the photovoltaic effect.
As an approximation one can take the energy of an electron in the conduction band to be

1 p2
E = m∗n v2 = (2.3)
2 2m∗n

where one needs to redefine the mass to include all interactions. For electrons this “effective mass”
is denoted by m∗n and for holes m∗p . By the de Broglie hypothesis of wave-particle duality, the
electron has a wave nature and a wavelength λ = h/p where h is Planck’s constant. It is usually

15
2 Photovoltaic Solar Cells

more convenient to think in terms of momentum and defines k = 2π/lambda. This k can be a
vector, just like the momentum p by using this relation for the components. Then

h̄2 k2
E= (2.4)
2m∗n

with a similar relation holding for holes. Thus near k = 0 the valence band and conduction band
can be described by a parabola. At each point on the parabola there is an energy state and this
constitutes the energy bands. This is illustrated in Fig 2.12 where on the left shows the band edges
of a valence band and conduction band. An electron can be excited from one to the other. On the
right is a slightly more sophisticated view, where the energy of a hole or electron is plotted vs k as
a parabola. Any point along the parabolas is a state and there can be an excitation from one band
to the other. The energy difference between the minima, in this case a k=0, is the bandgap.

Ec Ec
Ev
Ev

Fig. 2.12: An optical transition causing an electron to be excited from the valence band to the conduction band. On the right the
energy of each state is plotted vs k and is a parabola with the center at k=0. This parabolic behavior is an approximation
that holds near the top or bottom of the bands.

Away from k = 0, the E(k) relation is more complicated. The allowed energy states fall into
bands of energies, which are illustrated for Si and CdTe in Fig. 2.13. Because of the interference
effects the electron energy is a periodic function of k and the maximum unique value of k is about
π/a where a is a lattice parameter. This is typically no more than about 2 Å1− . The curves drawn
in Fig. 2.13 connect a series of very closely spaced points plotted over a range (Γ < k < X) and for
(L < k < Γ). These are curves in three dimensions, but plotted in two √ dimensions for simplicity.
The jargon is that Γ is k = 0, X is k = (π/a)[100] and L is k = (π 3/2a)[111], where a is the
lattice parameter.
By contrast a photon has an energy given by E = hν where h = 6.63 × 10−34 Js is Planck’s
constant. Since c = νλ , E = hν = hc/λ = 12400/λ , where the last holds when E is in eV and λ is
in Å. For example a visible photon at 6000 Å(orange) has k = 2π/λ = 0.001 Å−1 and E = 2.1 eV.
This says that if an electron absorbs energy from a photon, then the value of its momentum, k, is
not going to change very much compared to typical electron momenta. The absorption of a photon
by an electron is therefore called a vertical transition. This is illustrated in the left side of Fig. 2.13
for an absorption process in which a photon of about 3.4 eV causes an electron from the valence
band to jump to the conduction band minimum at k = 0, creating a hole in the valence band and
an electron in the conduction band. The key idea is that for Si this is not the bandgap energy.
For Si there is a minimum in the band gap at a value of k 6= 0. However, for this transition to
occur some additional k (recall h̄k is momentum) must be supplied. Interaction with the crystal
lattice, creating a lattice vibration can supply this momentum but it is less likely. Consequently Si
is much more weakly absorbing than CdTe, in which the valence band maximum and conduction
band minimum are both at k = 0 and so thick Si absorbers are required as opposed to thin CdTe

16
2 Photovoltaic Solar Cells

Blakemore
3.4 eV

=1.12

Fig. 2.13: Si has an indirect bandgap, while CdTe has a direct bandgap.

absorbers. Amorphous Si (written a-Si) does not have this simple band structure and it has many
defect sites that can couple in the absorption process so that optical absorption is more likely and
thinner absorber layers can be used. Your calculator, for example, uses a thin a-Si layer.

2.4.2 Optical Absorption


The absorption of light at the relatively low intensities involved in photovoltaics generally
follow Beer’s law. In this case the rate of change of the intensity passing through a material is
proportional to the light intensity. More simply, Fig. 2.14(a) show the variation in light intensity

I(λ,x)

x
x+dx
I(λ,x+dx)

“Thin Film Solar Cells,” Poortmans and Arkhipov eds.,

(2006), 285

Fig. 2.14: Optical absorption coefficient, α, definition and data for several materials. The absorptance, α, is a function of
wavelength which will be used in the design of heterostructure cells. The characteristic length for light absorption is
1/α. Note that in a-Si light is absorbed more strongly than in crystalline Si (for longer wavelengths).

passing through a material at planes through x and x + dx. The photon flux, or the number of pho-
tons per area per second, is reduced due to absorption. We assume no defects that cause scattering.
The flux exiting a the slice at x + dx would be expected to be proportional to the number entering
the slice at x, the thickness of the slice, dx, and the absorption process, described by α. If one
multiplies the flux by hν of the photon then the same description could be applied to the energy

17
2 Photovoltaic Solar Cells

per time (power) per area and one has for the light intensity:

I(x) − I(x + dx) = α · I(x) · dx

or
dI
= −αI
dx
where α is the absorptance and is a function of the wavelength. Integrating gives

I = I(0)e−α(λ ) x (2.5)

where I(0) is the intensity at x = 0 and the absorptance, α, is a function of the wavelength.
Fig 2.14(b) and (c) show the absorptance of several materials, including CdTe, crystalline Si (c-Si)
and amorphous Si (a-Si). From Eq. 2.5 shows that in a distance 1/α the intensity is reduced by a
factor of 1/e which is termed the absorption length and is a good measure of the strength of the
absorption. In these semilog plots one can see that the absorption in a-Si is much stronger than
C-Si and that CdTe at 500 nm is 105 while c-Si at 2.5 eV is 104 , i.e., a 10 times stronger absorber1 .

2.4.3 Carrier densities


An important result that we will need when examining the efficiency of photovoltaic cells
concerns the carrier densities in the conduction and valence bands of a semiconductor. For the
simple “free electron” band structure shown in Fig. 2.12(b) with energy given by Eq. 2.4 the number
of electrons in the conduction band is given by the integral of the number of states between E
and E + dE, times the probability that an energy state is occupied by an electron, all of which is
integrated over E. Following ref. [7] this is written per volume as
Z ∞
n= gc (E) f (E)dE (2.6)
EC

where f (E) = 1/(1 + exp (E − EF )/kT ) is the Fermi-Dirac distribution function and gc (E) is the
density of states in the conduction band

mn (2m∗n (E − Ec ))1/2
gc (E) =
π 2 h̄3
where m∗n is the electron effective mass, Ec is the lowest energy in the conduction band, and h̄ is
Planck’s constant. There is a similar relationship for the valence band. In this case the probability
that a state is unoccupied (a hole state) is 1− f (E) and then after integration the density of electrons
and holes becomes

n = Nc e(EF −Ec )/kT (2.7)


(Ev −EF )/kT
p = Nv e (2.8)

where Nc and Nv are called the effective density of states where N = 2(m∗ kT /2π h̄2)3/2 and m∗ is
either the effective carrier mass in the conduction band or valence band. In some sense one can use
Eq. 2.8 to treat the Fermi energy, EF , as a parameter that describes the carrier concentration even

1 Need to check this comparison. Seems too close.

18
2 Photovoltaic Solar Cells

though it is much more fundamental than this. One can see that if a material is P-type with a high
density of holes, EF will be close to Ev and toward the middle of the gap, far from Ev when the
material is low doped. If there are no acceptor impurities then the material is termed intrinsic and
EF is very close to the gap middle. Similarly an N-type material EF will be close to Ec if heavily
doped. When it is intrinsic, n = p = ni . At this carrier concentration, the Fermi level is called the
intrinsic level, Ei .
It is not too hard to show [7] that n and p for the same material are

n = ni e(EF −Ei )/kT (2.9)


(EF −Ei )/kT
p = ni e (2.10)

where ni = Nc exp ((Ei − Ec )/kT ) and that the product of n and p for the same material is indepen-
dent of EF and depends only on temperature, i.e.

np = n2i (2.11)
−Eg /kT
np = Nc Nv e (2.12)
1/2 −Eg /2kT
ni = (Nc Nv ) e (2.13)

so that the density of carriers in an intrinsic (undoped) material depends exponentially on the
bandgap. Since the bandgap is the energy barrier for an electron from the valence band to jump
into the conduction band to form an electron and hole pair (EHP), this exponential term represents
the process difficulty. Many physical processes that involve activation by jumping over an energy
barrier have a similar exponential term.
At a junction there are both types of materials and some notation is important to avoid
confusion. In an N-type region, nn is density of the majority carriers, pn is the density of minority
carriers (holes), and nno and pno are the majority and minority carrier densities in an equilibrium
situation – for example light is not being absorbed and creating EHP pairs or there is no forward
bias that injects carriers. However, for parameters such as electron mobility that depends on the
total number of acceptor and donor impurities we will use µN .

2.4.4 Minority carrier lifetime


It is useful to examine the np product from another point of view. For a semiconductor at
equilibrium there will be some number of electrons, no , and holes, po . This semiconductor is in
thermal equilibrium with its surroundings and the heat from the surroundings generates electrons
and holes at a rate G(T ). In equilibrium one requires that the thermal generation rate be balanced
by recombination of electrons and holes. Otherwise the carrier concentration would continue to
change. One expects the electrons and holes to recombine with a rate proportional to density of
electrons and holes so that at equilibrium

dno
= 0 = −Ano po + Gthermal (T ) (2.14)
dt
where A is some rate constant. For low carrier densities it is not expected to depend on carrier
concentration. This just says that how often an electron and hole recombine depends on how often
they find each other, which will depend on how many there are, similar to a chemical reaction.
The result is that in steady state no po = G(T )/A which as in Eq. 2.11 is a constant that depends
on temperature and not the doping density. If the material is intrinsic the constant must be n2i , by

19
2 Photovoltaic Solar Cells

substitution. Since G(T )/A does not depend on doping, no po = n2i , as before.
Now suppose that in addition to the thermal generation, there is a uniform distribution of
light that produces elecrons and holes. For specificity, assume an n-semiconductor that is now
away from equilibrium with free carrier densities nn and pn , so that the time evolution in minority
carriers is given by

d pn
= −Ann pn + Gthermal (T ) + GL (2.15)
dt
where the three terms on the right are: (1) recombination of electrons and holes, (2) thermal gen-
eration of carriers at the same rate as before, and (3) light induced generation of carriers.
Expressing this in terms of changes from the equilibrium density

d∆p
= −A(nno + ∆n)(pno + ∆p) + Gthermal (T ) + GL (2.16)
dt
= −Anno ∆p − A∆n∆p − Apno ∆n + GL (2.17)
= −Anno ∆p + GL since the middle terms are small (2.18)
d∆p ∆p
= − + GL (2.19)
dt τp

where τn = 1/Anno is the minority carrier recombination time. In steady state one easily sees that

Minority Carrier Lifetime (Virginia Semiconductor)


102
Electrons in p-Si
Holes in n-Si
101

100
Lifetime ( s)

1
10
2
10
3
10
4
10
1017 1018 1019 1020
Majority Doping Density (cm ) Fig. 2.16: Minority carrier lifetime in p-doped GaAs, col-
lected by Boone and Woodall and Yale data(?)
Fig. 2.15: The minority carrier lifetime in Si from Virginia and is about a factor of 50 less than Si at 1 ×
Semiconductor. The lifetime depends on the im- 1019 cm−3 .
purity (dopant) density.

for an n-type material ∆p = GL τ p .

2.4.5 Diffusion of carriers


Carriers in a semiconducting material are always constantly in a random motion. If the
distribution is uniform then there are as many moving in, say, the +x̂ direction as there are in the
−x̂ direction and there is no net current. If there is a non-uniform distribution of carriers then one
expects the carriers to move away from regions of high density towards regions of low density.

20
2 Photovoltaic Solar Cells

There is a net current given by

dn
JN,di f f = qDN (2.20)
dx
dp
JP,di f f = −qDP (2.21)
dx
where q = |q| is the electron charge and DN , DP are the electron and hole diffusivities, respectively,
and JN,di f f is the electron diffusion current. Note that like mobility, DN and DP each depend on the
total number of acceptors and donors.
The continuity equation expresses the statement that the difference between a current enter-
ing or leaving a slice (in one dimension) is due to charge recombination or generation via light in
terms of electron and hole total (drift plus diffusion) currents, i.e.

∂n 1 ∂ JN
= (2.22)
∂t q ∂x
∂p 1 ∂ JP
= − (2.23)
∂t q ∂x

The carrier concentrations can decrease due to recombination and increase due to light ab-
sorption. Hence if n = no + ∆n then

∂ n ∂ no ∂ ∆n ∂ ∆n
= + = (2.24)
∂t ∂t ∂t ∂t
and similarly for p(t). For minority carriers

∂ ∆n p ∆n p
= − + GL (2.25)
∂t τn
∂ ∆pn ∆pn
= − + GL (2.26)
∂t τp

Combining the diffusion equation and the continuity equation one obtains the so-called mi-
nority carrier diffusion equation [7, 8]:

∂ ∆n p ∂ 2 ∆n p ∆n p
= DN − + GL (2.27)
∂t ∂ x2 τn
∂ ∆pn ∂ 2 ∆pn ∆pn
= DP − + GL (2.28)
∂t ∂ x2 τp

This important result will mainly be of interest to us in steady state when the left hand side is zero.
If in addition GL = 0 and the region of interest is long then each has a solution like

∆n p (x) = ∆n p (0)e−x/LN (2.29)

where LN = (DN τn )1/2 is interpreted as the diffusion length of minority carrier electrons before
recombination. We will take this as an important collection length in a photovoltaic cell.

21
2 Photovoltaic Solar Cells

2.5 Diodes

2.5.1 Photovoltaic diode: the simple qualitative picture


A semiconductor with a bandgap is schematically indicated in Fig. 2.17 by two horizontal
lines separated by the gap energy. A photon must have an energy greater than this to be absorbed.
The x-axis in this schematic does not really correspond to anything as the separation between the
horizontal lines correspond to the minimum separation of the valence and conduction bands a one
k-state. Fig. 2.18 shows a N-type (left) and P-type (right) placed together to form a PN junction.
If there is no applied voltage the Fermi levels (EF ), which correspond to the energy at which the
occupation probability is 1/2, are equal. In this case there is a slight forward bias and EF,N > EF,P .
In an energy band picture, when a photon is absorbed by a semiconductor, the photon’s
energy (= hν) is given to an electron in the valence band, putting this electron in the conduction
band. It is also possible for an electron already in the conduction band to absorb the energy but the
absorption probability is lower and there are far fewer such electrons to begin with. The main ab-
sorption process is illustrated in Fig. 2.17 and in Fig. 2.18 these photons are shown being absorbed

GL

V
Ln
Fig. 2.17: Absorption of a photon creates an
electron and hole pair. Fig. 2.18: Long Diode

on the P-type side of the pn junction. Those electrons within a diffusion length of the junction will
occasionally wander there and fall down the energy hill, until the barrier height is reduced by a
forward bias voltage V . At this voltage, the injected carriers balance the light induced carriers. A
solar cell under open-circuit conditions is shown in Fig. 2.18 so that the light induced current is
balanced by the diode current and there is no net current. The voltage, V , across the diode is the
voltage required to balance the light induced current. Since this is no longer equilibrium these are
quasi-Fermi levels and the voltage is their difference.

2.5.2 The diode: simple quantitative picture


The reverse saturation current is the current of carriers that are thermally excited that can
diffuse to the barrier region. For example, in the P-type region there will be n po of electrons that
are
√ thermally excited into the conduction band. In a time τn , those that can diffuse a distance
Dn τn will reach the depletion region and fall down the energy√hill. The reverse current, Io , to the
depletion region is then the number of electrons in a volume A Dn τn times the charge, q, divided
by the time, τn , that it takes these electrons to diffuse to the depletion region per unit area.

A DN τn DN 1/2
Io = Jo A = qn po = qn po (A ) (2.30)
τn τn
Jo = qn po DN /LN (2.31)
qn2i DN
Jo = (2.32)
NA LN

22
2 Photovoltaic Solar Cells

Similarly, minority carrier holes on the n-side can diffuse to and cross the barrier, contributing a
similar term qn2i DP /ND LP so that the total reverse current in the diode current is the sum

qn2i DN qn2i DP
Jo = + (2.33)
NA LN ND LP
At zero bias the current to the right in the diode of Fig. 2.18 will be Jo given in Eq. 2.33 and this
has to be balanced by a forward current to the left. The electrons moving to the right or the holes
to the left need to surmount a barrier and expect an exponential like that in Eq. 2.9 to control the
current. At zero bias there is no current and so at a forward bias the reduction of the barrier height
must enter as

J = Jo (1 − eqV /kT ) (2.34)

with Io given above. This is the Shockley equation.


To find the diode current if there is also light present, one can take Eq. 2.19 add a light
induced current. As in the case of the minority carriers diffusing to the junction. Now the density
of carriers in, say the n-side, is GL τn so that in a time τn , qGL τn LN charges will reach the depletion
region. Then the current due to the light is just qGL LN . The total current is the Shockley current
plus the light induced current or
 
J = Jo 1 − eqV /kT + qGL LN (2.35)

or the light induced current is IL = qAGL LN . Note that the light induced current is in the same
direction as the reverse saturation current and that in this example we just consider the absorption
in the P-side of the junction.
The equivalent circuit, with the addition of some realistic resistances is shown in Fig. 2.19.

Rs

+
ID IL RL −V

Fig. 2.19: Equivalent circuit. Note that IL is in the same direction as the reverse saturation current (and the mechanism causing
them are quite similar). In a slightly more advanced model one would put in a shunt resistance in parallel with IL .

2.5.3 Uniformly illuminated long diode (optional)


Assume GL is uniform over the p-side of a long n+ p, one-sided, long diode. This is equiva-
lent to saying that the absorption coefficient, α  1/length. As usual for pn junctions, we need to
find the minority carrier diffusion current at x = 0, just outside the depletion region on the p-side
of the junction.2

2 Recall
the subtle argument here: We neglect the minority carrier drift current so that the total current is the sum of
the majority carrier drift and diffusion plus the minority carrier diffusion. For the majority carrier current, use the
minority carrier diffusion current on the other side of the junction. We neglect recombination in the depletion region.

23
2 Photovoltaic Solar Cells

n + − Si

p−Si

u=x−xp

−Wn− xn −xn 0 xp Wp+ x p

Fig. 2.20: Uniformly illuminated long diode. Not sure about first coordinate label in this figure. Hold off on fixing as might want
for properly including light absorption.

First we need to solve the minority carrier diffusion equation in steady state, i.e.,

d 2 ∆n ∆n
0 = DN − + GL (2.36)
dx2 τn
which has a general solution

∆n = Ae−u/LN + Beu/LN +C (2.37)



where LN = DN τn . Substituting this solution into Eq. 2.36 one sees that C = GL τn . At u = +∞
the solution diverges and so B must be zero. The result is that

∆n(u) = Ae−u/LN + GL τn (2.38)


−u/LN
= (∆n(0) − GL τn )e + GL τn (2.39)

The electron diffusion current at u = 0 is qDN d∆n/du there. Or

∆n(0) − GL τn
JN = −qDN (2.40)
LN
∆n(0) qDN GL τn
= −qDN + (2.41)
LN LN
∆n(0)
= −qDN + qGL LN (2.42)
LN

where the first term, after substituting the law of the junction for ∆n(0), is the usual Shockley diode
current and the second term can be interpreted as the current due to all photons collected within
a diffusion length of the junction. Note that the light induced current is in the same direction as
the reverse saturation current. On the n-side it can be assumed that there is little addition to the
photocurrent. Since it is heavily doped the lifetime is short and hence the diffusion length is small.
Depending on dimensions few carriers can reach the depletion region and half will hit the surface
recombination difficulty at the front surface. 3 Putting in the law of the junction
qDN
J = − n po (ev/VT − 1) + qGL LN (2.43)
LN
= Js (ev/VT − 1) − JL (2.44)

where JL = −qGL LN and Js = −qDN n2i /NA . The signs have been adjusted to be consistent with
3 Expand?

24
2 Photovoltaic Solar Cells

diode conventions. The light induced current is in the same direction as the reverse saturation
current.
A similar result could hold on the n-side but since it is heavily doped and thin, the main
contribution will be Eq. 2.43. Also note that because of the assumption that the diode was long we
did not have to worry about surface recombination at the boundary.

2.5.4 Diode with exponential absorption (optional)


A more realistic situation would be for a diode to absorb light entering from the front with
exponential absorption. The simplest case would be to consider one wavelength, where IL (x) =
Io exp (−αx), where Io is the power per unit area incident on the front surface. . The diffusion
equation on the p-side is then

d 2 ∆n ∆n αIo −αx
0 = DN − + e (2.45)
dx2 τn hν
This has a solution Check subscripts
here
αΦo τn −αx
∆n(x) = A1 e−x/LN + A2 ex/LN + e (2.46)
1 − α 2 LN2

where Φo = Io /hν is the number of photons per unit area or equivalently

αΦo τn −αx
∆n(x) = B1 sinh x/LN + B2 cosh x/LN + e (2.47)
1 − α 2 LN2

where the equivalent, but different constants Ai or Bi are to be determined by the conditions at the
ends of the p-region. To go further we need to be able to determine the excess carrier density at the
boundary. At an interface there is usually enhanced recombination because of dangling bonds that
trap carriers until they recombine with those of the opposite sign. This is treated according to

∂ pn
qDP |x=xo = qS p ∆p(x)|x=xo (2.48)
∂x
∂ np
qDN |x=xo = −qSn ∆n(x)|x=xo (2.49)
∂x
(2.50)

where S p has units of cm/s and so is identified with a velocity. The left hand side is a (diffusion)
current and the right hand side is a current proportional to the difference between the carrier density
at xo and its equilibrium value. The current is always a finite quantity and so if S p is infinite, then
∆p(xo ) = 0. When S p = 0 then the slope of pn (x) at x = xo is zero. The exact value of SP depends
on surface treatment. A high value would be 104 or 105 ; a low value would be of the order of
10 cm/s.
Once again when W  L only the decreasing exponential is needed and one can find the
diffusion current at x = 0. Compare the result to an absorption within the diffusion length of
1 − exp −αL.

25
2 Photovoltaic Solar Cells

2.5.5 Band Diagrams (optional)


At open circuit the quasi-Fermi levels for holes and electrons are constant. The band diagram
would look something like Fig. 2.21 in which the generation of EHP’s causes the quasi Fermi levels
to shift apart. Assume the n-side is heavily doped and we only consider the carriers generated in
the p-side. For the p-side, EF,n moves towards the conduction band while EF,p , describing the
majority carrier density, does not change much. The open circuit voltage can be obtained from the

EF,n
EF,p

Fig. 2.21: Band diagram at open circuit showing the constant quasi-Fermi levels.

separation of these levels. Far from the junction, on the p-side:

n p = ni e(EF,n −Ei )/kT = Gτn + n po

p p = ni e(Ei −EF,p )/kT = Na


Taking their product

(Gτn + n po )Na = n2i eqVoc /kT

or
kT Gτn
Voc = ln( + 1)
q n po

kT IL
Voc = ln( + 1)
q Is
This last uses IL = qAGL LN and Is = qADN n po /LN .

2.6 Power and efficiency

2.6.1 Maximum Power


The maximum power one can obtain from a photovoltaic device is an issue that is important
at a number of levels. From a very practical point of view, this determines the specific circuit that
extracts the power from a device. From a more theoretical point of view one needs to know if a
photovoltaic structure based on a given material has reached its full potential and whether more
effort can reasonably expected to improve its performance.
The first issue is addressed by considering the ideal diode equation, Eq. 2.51. We will add
series and parallel resistance later. Assume that the diode current is given by

I = +Is (eV /VT − Is ) − IL (2.51)

26
2 Photovoltaic Solar Cells

Fig. 2.22: Plot of IV and PV including power square, showing the fill factor – area of filled rectangle divided by that of the open
rectangle. The current is arbitrarily taken to be negative to distinguish it from the power.

where the signs are chosen so that the diode reverse current and the light generated current are in
the same direction. This equation is often written with +IL since that is usually the main term of
interest. Here it is plotted in Fig. 2.22 with IL = 100 mA and Is = 1 nA as the black curve. Then
one can easily compute the power, p = iv and plot that vs the voltage across the diode, v. At open
circuit, the power is zero since there is no current and at short circuit the voltage is zero. The
maximum power developed is at a point in between and can be found [9–11] numerically or by
differentiation.
The power per area from a photovoltaic device is thus
 
P = JV = JL + Js (1 − eV /VT ) ×V (2.52)

and one wants to determine the voltage and current at maximum: Pm = JmVm . Unfortunately this
requires a numerical solution. One can modify this with some arithmetic, which will be discussed
below, but it is difficult to do better than calculating a list of powers and using a max(P) function.
Graphically the maximum power is represented by a square with one corner at the origin
and the other at (Vm , Im ). A measure of how square the IV curve is the fill factor defined as:4
ImVm
FF = (2.53)
ILVoc
In 2018, the best single crystalline Si cell had a FF of 84.9%, GaAs 86.5%, CIGS 74.3%, CdTe
79.4%, and perovskite 74.5%.
First we obtain an approximation for the maximum power that can be developed by a pho-
tovoltaic cell. From Eq. 2.52 and setting dP/dV = 0 one finds that the current and voltage at
maximum power satisfy:

 
JL /Js + 1
Vm = VT ln ≈ Voc −VT ln (1 +Vm /VT )
1 +Vm /VT

4 M. Green has a paper on FF and series and parallerl R.

27
2 Photovoltaic Solar Cells

with
JL
Voc = VT ln ( + 1)
Js

Unfortunately, to find Vm one must solve

JL /Js + 1
Vm = VT ln ( ) (2.54)
1 +Vm /VT

which is nonlinear, though it can be iterated.

2.6.2 Efficiency
It is crucial to know how Eq. 2.54 varies with a material’s bandgap. Without this knowledge
we would not know which material to focus our effort nor the maximum efficiency that we could
expect. Given the solar spectrum and a material’s bandgap we can find JL if we assume that all
photons with energy greater than Eg contribute one charge to the current:

I(λ )
Z λg
JL = q dλ (2.55)
0 hc/λ

where I(λ ) is the irradiance in (λ , λ + dλ ) and λg = hc/Eg . We also need to know how JS varies
with bandgap.
There is an elegant means to use thermodynamics to find Js (Eg ) [12] but the argument is
complicated. For simplicity we note from the Shockley ideal diode model that

qD p n2i qDn n2i


 
qD p qDn
Js = + = + NC NV e−Eg/kT
LN ND Ln NA LN ND Ln NA

and note that the fastest material dependence on the bandgap is in the exponential. For simplicity
we take the approach suggested by Sze [13] and fit this last exponential dependence to data for
champion cells from 2018 [6] as shown in Fig. 2.23. Doing this we obtain

Js = 2.68 × 109 e−Eg /kT mA/cm2 . (2.56)

This estimate gives a very similar result to the calculation of energy efficiency using a more elegant
thermodynamic analysis [9–11].
To find the efficiency of the diode one needs to find the maximum power it can develop
divided by the incident power. This will depend on the material in three ways in this formulation.
First, the current JL depends on the bandgap Eg since the diode cannot absorb a photon with hν < Eg
and all energy hν − Eg is wasted. The consequence is that a photon, despite its energy, can create
at most one EHP. This is illustrated in the figure Excess energy above Eg will be given off as lattice
vibrations in the order of femtoseconds; then the electrons and holes recombine over microseconds.
Second the diode dark current, Js , will vary from material to material. This will affect the open
circuit voltage. Last, the cell has to operate under forward bias in order to provide power. Carriers
will be injected across the barrier, which then recombine giving off light and heat.
We can now calculate the efficiency of a photovoltaic cell as a function of the bandgap, Eg .
From Eq. 2.56 we can determine Js . Knowing the incident solar spectrum and if we assume that any

28
2 Photovoltaic Solar Cells

10 -4

10 -6

10 -8

Js (mA/cm 2 )
10 -10

10 -12

10 -14

10 -16

10 -180.9 1.0 1.1 1.2 1.3 1.4 1.5


Eg (Volts)
Fig. 2.23: Fit of the reverse saturation current to champion cells in 2018 [6].

~ fs

~µ s
V

~ fs
Fig. 2.25: To generate power the diode must be
Fig. 2.24: Photons with energy greater the bandgap slightly forward biased. This will inject car-
create electron hole pairs that quickly (∼ riers across the built in barrier, which re-
100 fs) deexcite to the band edges, wasting combine, giving off heat and light, wasting
energy that now cannot be recovered. energy. This is a cost of doing business.

photon with energy greater than the bandgap is absorbed and creates one EHP that is collected, we
can calculate the short circuit current. Then from Eq. 2.54 we can calculate the maximum power
that can be developed. From the incident spectrum we can find the incident power. The ratio gives
the efficiency.

2.6.3 Maximum Power Point Tracking


A useful result is to notice that at maximum power

d dI
IV = ·V + I = 0
dV dV
or
dI I
=−
dV V

29
2 Photovoltaic Solar Cells

Fig. 2.26: Efficiency of an ideal solar cell material based on the reverse saturation current of Si, scaled to the material’s bandgap.
The ideal cell shows a broad maximum with an optimum bandgap near 1.4 eV. The closed circles are record, laboratory
efficiencies from Ref. [14]
.

and that the current lies on the hyperbola described by


Pmax
I=
V
and is tangent to the current voltage curve at the maximum power point.

2.6.4 Losses at surfaces


Show for W ≈ L a better solution involves sinh and cosh. Find the amount absorbed and
compare to previous. Now need to include surface recombination at the back. Consider the two
limits of zero and infinite surface recombination.

2.6.5 Summary
The power absorbed from the photon flux is not all available for work in an external circuit.
Fig. 2.27 illustrates the absorption process and one can see how the energy is dissipated after the
absorption. For electron hole pairs created outside of a diffusion distance from the junction there
is radiative and non-radiative recombination that emits energy either as photons or as heat. EHP’s
created above bandgap will relax to the band edges, emitting heat. Current that is injected due to
the forward bias reduces the photocurrent. Net electrons going to the left and falling down (or holes
going to the right and bubbling up) the potential hill will lose energy equal to qVbi − qV , emitted
as heat. Electrons going from the semiconductor to the metal will lose energy corresponding to
the drop to the Fermi level of the metal (unlike the semiconductor, there are empty states there).
On the other side, electrons from the conductor will fall into hole states in the semiconductor,
losing energy corresponding to the difference in the Fermi level in the conductor and the valence
band edge in the semiconductor. The power remaining will be the current times the forward bias
voltage. An EHP that is created will, after energy loss to the band edges, will one way or another
give off Eg .

30
2 Photovoltaic Solar Cells

V
IL

diffusion length
Fig. 2.27: A solar cell connected to a load is under illumination.

2.7 Losses

2.8 Recent Innovations

2.8.1 Si Cell
An old multicrystalline cell is shown in Fig. 2.28

Fig. 2.28: Large multicrystalline Si cell, about 1m by 2 m.


One can see individual single crystalline regions Fig. 2.29: Close up showing that for each cell to be in
as well as the contact fingers in a grid. There are series, the finger from the front surface of one
36 cells wired in series to give an open circuit must contact the rear surface of the pn junction
voltage of 25 V. preceding.

31
2 Photovoltaic Solar Cells

Fig. 2.30: Hanwa Q cell: see technote for more

2.8.2 PERT cell


2.8.3 Half Cell
2.8.4 Bifacial Cell

Fig. 2.31: Soltec bifacial cell mount

2.9 Beyond Silicon

2.9.1 Heterostructure
Will the band offsets in InGaN permit a reasonable solar cell? Discuss reduction in surface
recombination. Discuss competing factors of series R, absorption, surface recomb, . . . .
An n-CdS—p-CuInSe pn junction device is shown in Fig. 2.32. It would be better to use a
CdS/CdTe cell here and to compare to vendors.

32
2 Photovoltaic Solar Cells

We assume that the CdTe is grown under Te rich conditions and is slightly P-type due to Te
interstitials [15]. Assume the following parameters: PRB66, 155211
(2002).

n−CdS p−CuInSe

W n Lp 0 Ln Wp

−xn xp

Fig. 2.32: Example heterojunction cell with a CdS window layer. Not even close to scale.

n-CdS p-CdTe Notes


N C-Si CIGS a-Si:H
mobility to get resistivity
xN , xP band offsets complicate this Lsc µm 0.3 1 0.3
λ LD (µm) >100 1 <0.1
α
Eg from J.D. Cohen
WN ,WP
Ask Trevor to
simulate this with
SCAPS. Make
certain we are
including the main
ingredients
properly.

Fig. 2.34: Ions are used to mill away material and at the same time
the surface chemical composition is measured by exam-
Fig. 2.33: CIGS Repins 2008 record cell;
ining the secondary electron emission. [16]
uncapped test sample

Contreras and coworkers have attributed the gains in performance needed to reach record
CdTe cells to:
1. optimized ZnO window layers
2. improvements in the diffusion length of minority carriers, partly due to composition grading
that induces a field
3. reduction in recombination in the space-charge region due an improved GIGS/CdS interface
M. Igalson [17] treats the n-CdS/p-CIGS as if Nd = 1017 cm−3 with a bandgap of Eg = 3.4 eV
and Na = 5×1015 cm−3 with bandgap of Eg = 1.2 eV. The n-layer is 100 nm wide n-layer. Assumes
a spike type alignment at the interface with 0.2 eV. This cell is complicated by defects. At the

33
2 Photovoltaic Solar Cells

junction, Igalson assumes a divacancy which adds a feature to the band structure that due to heavy
p-doping seems to impede the flow of carriers and acts like a series resistance.

MgF 2 (80 − 120 nm)


n−ZnO (300 − 650 nm)
i−ZnO (50 nm)
CdS (20 − 50 nm) n~10 17 cm −3
p~5x10 15 cm −3
Cu(In 0.7 Ga 0.3 ) Se 2 (2000 − 3000 nm)
~3nm, Cd doped
p+−region Eg =1.18 eV

Mo (1000 nm)

Soda Lime Silica Glass ( 2 − 4 mm)


(substrate)

Fig. 2.35: Schematic of a CIGS cell. This does not reflect defects properly. There could also be a Zunger divancy.

2.9.2 CdTe
2.9.3 Perovksites
2.10 Next Generation Structures

2.11 Concentrator Designs

Solar cells are used in either a flat plate or in a concentrator geometry. This section is in-
cluded for completeness but since they are not used in terrestrial environments it can be considered
optional reading. The difficulty is for significant concentration of light, the device becomes quite
hot and the cell efficiency is reduced. Further, when using lenses to concentrate the light one must
track, which is often less convenient.
Generally when the cell is very expensive to fabricate it is used in a concentrator geometry
in order to minimize the size of the cell needed. For example by using lenses or mirrors to collect
light from a large solid angle and focus onto a small spot, only a small device would be needed to
convert the light incident on a large area flat cell. A concentrator shown in Fig. 2.36 by SolFocus,
Inc. is one implementation of a reflection geometry. It is a simple construction but will need to
contend with heat transfer and wavelength dependent optical absorption. This reflection geometry
is similar to that used in some telescope designs. The main difficulty here is that only light from a
narrow region around the sun is focused onto the photovoltaic cell. Light that has bounced off of
a cloud would not be imaged onto the cell, nor would light from the sun after it moves in the sky.
As a consequence, the concentrator must track the sun as it moves throughout the day, adding an
additional expense and maintenance burden.
There are some designs, called luminescent solar concentrators [18], in which fluorescence
is used to concentrate light, rather than lenses. These are under active development by several
groups.

34
2 Photovoltaic Solar Cells

Fig. 2.36: A concentrator scheme used by SolFocus, Inc. to collect light over a large area and focused onto a small, more
expensive photovoltaic cell. This concentrator can be made by “slumping” which is a manufacturing technique used
in automobile headlights. Some kind of molding process. The center top would be coated with a metal reflector and
the cell placed at the bottom. This looks like a relatively inexpensive structure.

2.12 Exercises

1. Download the solar spectrum for AM1.5 (google AM1.5 and xls). Plot the number of photons
per unit wavelength per second striking the earth per square meter vs wavelength. Use the
“direct + circumsolar” data.
2. From the “Global tilt” data, what is the total power per sq m?
3. You are asked to construct a photovoltaic cell that only uses light over a very narrow wave-
length interval, say 0.01 nm (different parts of the device will use different wavelength inter-
vals, but for this question just consider one interval). Assume that every photon is absorbed,
that each absorbed photon creates one electron, and perfect charge collection efficiency. Us-
ing the irradiance data in the graph below:
a) What is the current density at 1500 nm?
b) At which wavelength is the current density maximum?
1.2
Irradiance (W m−2nm−1)

0.8

0.6

0.4

0.2

0
0 500 1000 1500 2000 2500
λ (nm)

4. From the Global tilt data, how many photons are there per second per sq m with an energy
> 1.12 eV? If each photon produced an electron, what current would this correspond to?
5. The U.S. generated 4.1 × 1015 Wh of electricity in 2011. Use a tilted flat plate solar cell with
an efficiency of 20% to generate an equal amount of electricity from the sun. What fraction
of Utah (220,000 km2 ) would be needed. You can use the graph of the incident solar energy
at the link: NREL annual map jpg.
6. Estimate the solar energy one could expect to be able to obtain (tilted flat plate) from panels
on a typical residential roof in Minneapolis. How much is needed – assume that natural gas
is used to power the heater, water heater, and clothes dryer.
7. What are the main power requirements of a typical house, breaking this down to water heater,

35
2 Photovoltaic Solar Cells

home heating, refrigerator, computers, . . . ?


8. A flywheel consisting of a cylindrical shell has a rotational energy of (1/2)Iω 2 where I =
mR2 . For an angular velocity of 1500 rpm, what diameter flywheel is needed to store 5
W·hr/kg.
9. The dose used in FDA sunscreen testing is 2.2 mg/cm2 of exposed skin. For an adult 5ft 4in
150lb this is about 1oz or 29gm. The face is about 1/4 to 1/3 tsp. The SPF of sunscreen is
approximately

SPF ≈ 1/T

where T is the transmission given by de Beer’s law. Assume you purchased SPF 20 sunscreen
but would like to get the benefit of SPF 40. How much thicker a layer must you apply?
(cf U. Osterwalder, Brit J Derm (2009) v161 (suppl3) 13)
10. (following Pierret, prob 6.8) Consider an ideal p+ n step junction where GL electron-hole
pairs per s per unit volume are created uniformly throughout the diode. (In class I dis-
cussed an n+ p junction.) In this case when a voltage is applied holes are injected into the
n-region. Assume low-level injection. Referring to Pierret’s texty by combining the continu-
ity eq. (3.46b) with eqs. (3.51) and (3.52) (but for holes, not electrons) and with the diffusion
equation eq. (3.50) also for holes, one obtains eq. (3.54b):

∂ ∆pn ∂ 2 ∆pn (x) ∆pn (x)


= Dp − + GL (3.54b)
∂t ∂ x2 τp

Then the main equation of interest that describes the minority carrier density vs distance
from the edge of the depletion region is:

∂ 2 ∆pn (x) ∆pn (x)


0 = Dp − + GL
∂ x2 τp

a) Explain what τ p corresponds to.


b) Explain why for this solar cell problem we set ∂ ∆pn
∂t = 0.
c) Why do we only solve this equation in the neutral n-side and ignore the depletion
region?
d) Assuming that the depletion region edge is at x = 0 show that

∆pn (x) = Aex/L + Be−x/L +C

is a solution for ∆pn (x) on the N-type side of the junction.


e) Determine A, B, C, and L in terms of constants and ∆p(0).
f) Let the boundary condition at x = 0 be the same you used in 3161, i.e.

∆pn (0) = pno (exp (qV /kT ) − 1)

Calculate the minority carrier diffusion current at x = 0.


g) Why is the last result the diode current

jD = Jo (evD /VT − 1) − JL

36
2 Photovoltaic Solar Cells

11. A solar cell has a short circuit current density of 40 mA/cm2 and an open circuit voltage
of 0.60 V under one-sun illumination at room temperature. Use the ideal diode equation to
calculate the open circuit voltage which is expected at 1000 suns.

 
iD = Io evD /VT − 1 − IL

12. A p-n junction solar cell has Voc = 0.5 V and Jsc = 20 mA/cm2 . A second cell, of the same
area, has Voc = 0.7 V and Jsc = 12 mA/cm2 . Assuming that both cells obey the ideal diode
equation, find the values of Voc and Jsc when the two are connected in parallel?
13. A p-n junction solar cell has Voc = 0.5 V and Jsc = 20 mA/cm2 . A second cell, of the same
area, has Voc = 0.7 V and Jsc = 12 mA/cm2 . Assuming that both cells obey the ideal diode
equation, find the values of Voc and Jsc when the two are connected in series? Determine the
voltage across each diode (a) approximately and (b) exactly.
14. You are asked to construct a photovoltaic cell that only uses light over a very narrow wave-
length interval, say λ0 ± 0.01 nm. Assume that one photon can create one electron, and
perfect collection efficiency. Use the irradiance data in the graph below. At λ0 = 500 nm,
how much power per m2 is wasted due to bandgap mismatch? Assume a Si cell, Eg = 1.1 eV

1.2
Irradiance (W m−2nm−1)

0.8

0.6

0.4

0.2

0
0 500 1000 1500 2000 2500
λ (nm)

37
2.13 References

[1] D. M. Chapin, C. S. Fuller, and G. L. Pearson. A new silicon p-n junction photocell for
converting solar radiation into electrical power. J. Appl. Phys. 25, (1954) 676.
[2] M. A. Green. Photovoltaic principles. Physica E 14, (2002) 11.
[3] R. Fu, D. Feldman, R. Margolis, M. Woodhouse, and K. Ardani. US Solar Photovoltaic
System Cost Benchmark: Q1 2017, NREL/TP-6A20-68925. Tech. Rep. NREL/TP-6A20-
68925, National Renewable Energy Laboratory (2017).
[4] U.S. Energy Information Administration. Today in Energy. Tech. rep., U.S. Department of
Energy (2018).
[5] N. Y. Kiang. The color of plants on other worlds. Scientific American 298, (2008) 48.
[6] M. A. Green, Y. Hishikawa, E. D. Dunlop, D. H. Levi, J. Hohl-Ebinger, and A. W. Ho-Baillie.
Solar cell efficiency tables (version 52). Prog Photovolt Res Appl. 26, (2018) 427.
[7] R. Pierret. Semiconductor Device Fundamentals (Addison-Wesley, 1996).
[8] V. Zeghbroeck. Principles of Semiconductor Devices and Heterojunctions (online and print
(Prentice Hall PTR, 2007).
[9] W. Shockley and H. J. Queisser. Detailed balance limit of efficiency of p-n junction solar
cells. J. Appl. Phys. 32, (1961) 510 .
[10] C. H. Henry. Limiting efficiencies of ideal single and multiple energy gap terrestrial solar
cells. J. Appl. Phys. 51, (1980) 4494 .
[11] A. Marti and G. L. Araujo. Limiting efficiencies for photovoltaic energy conversion in multi-
gap systems. Solar Energy Materials and Solar Cells 43, (1996) 203 .
[12] H. J. Queisser. Detailed balance limit for solar cell efficiency. Mater. Sci. Eng. B 159-160,
(2009) 322 .
[13] S. M. Sze and K. K. Ng. Physics of Semiconductor Devices (New York, John-Wiley & Sons,
2007).
[14] M. A. Green, K. Emery, Y. Hishikawa, and W. Warta. Solar cell efficiency tables (version
34). Progress in Photovoltaics: Research and Applications 17, (2009) 320 .
[15] S.-H. Wei and S. B. Zhang. Chemical trends of defect formation and doping limit in II-VI
semiconductors: the case of CdTe. Phys. Rev. B 66, (2002) 155211 .
[16] M. A. Contreras, B. Egaas, K. Ramanathan, J. Hiltner, A. Swartzlander, F. Hasoon, and
R. Noufi. Progress toward 20% efficiency in Cu(In,Ga)Se2 polycrystalline thin-film solar
cells. Prog. Photovoltaics 7, (1999) 311 .
[17] M. Igalson, P. Zabierowski, D. Przado, A. Urbaniak, M. Edoff, and W. N. Shafarman. Under-
standing defect-related issues limiting efficiency of CIGS solar cells. Solar Energy Materials
and Solar Cells 93, (2009) 1290 .
[18] W. H. Weber and J. Lambe. Luminescent greenhouse collector for solar radiation. Appl
Optics 15, (1976) 2299.

38

Вам также может понравиться