Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 13

Multi-junction solar cell

From Wikipedia, the free encyclopedia


(Redirected from Multi-junction)

Jump to navigationJump to search

Black light test of Dawn's triple-junction gallium arsenide solar cells[1]

Multi-junction (MJ) solar cells are solar cells with multiple p–n junctions made of different
semiconductor materials. Each material's p-n junction will produce electric current in response to
different wavelengths of light. The use of multiple semiconducting materials allows the absorbance
of a broader range of wavelengths, improving the cell's sunlight to electrical energy conversion
efficiency.
Traditional single-junction cells have a maximum theoretical efficiency of 33.16%.[2] Theoretically, an
infinite number of junctions would have a limiting efficiency of 86.8% under highly concentrated
sunlight.[3]
Currently, the best lab examples of traditional crystalline silicon (c-Si) solar cells have efficiencies
between 20% and 25%,[4] while lab examples of multi-junction cells have demonstrated performance
over 46% under concentrated sunlight.[5][6][7] Commercial examples of tandem cells are widely
available at 30% under one-sun illumination,[8][9] and improve to around 40% under concentrated
sunlight. However, this efficiency is gained at the cost of increased complexity and manufacturing
price. To date, their higher price and higher price-to-performance ratio have limited their use to
special roles, notably in aerospace where their high power-to-weight ratio is desirable. In terrestrial
applications, these solar cells are emerging in concentrator photovoltaics (CPV), with a growing
number of installations around the world.[10]
Tandem fabrication techniques have been used to improve the performance of existing designs. In
particular, the technique can be applied to lower cost thin-film solar cells using amorphous silicon, as
opposed to conventional crystalline silicon, to produce a cell with about 10% efficiency that is
lightweight and flexible. This approach has been used by several commercial vendors,[11] but these
products are currently limited to certain niche roles, like roofing materials.

Contents

 1Description
o 1.1Basics of solar cells
o 1.2Loss mechanisms
o 1.3Multi-junction cells
 1.3.1Material choice
o 1.4Structural elements
 1.4.1Metallic contacts
 1.4.2Anti-reflective coating
 1.4.3Tunnel junctions
 1.4.4Window layer and back-surface field
o 1.5J-V characteristic
o 1.6Theoretical limiting efficiency
 2Materials
o 2.1Gallium arsenide substrate
o 2.2Germanium substrate
o 2.3Indium phosphide substrate
o 2.4Indium gallium nitride substrate
 3Performance improvements
o 3.1Structure
o 3.2Spectral variations
o 3.3Use of light concentrators
 4Fabrication
 5Comparison with other technologies
 6Applications
 7See also
 8References
 9Further reading

Description[edit]
Basics of solar cells[edit]

Figure A. Band diagram illustration of the photovoltaic effect. Photons give their energy to electrons in the
depletion or quasi-neutral regions. These move from the valence band to the conduction band. Depending on
the location, electrons and holes are accelerated by Edrift, which gives generation photocurrent, or by Escatt, which
gives scattering photocurrent.[12]
Traditional photovoltaic cells are commonly composed of doped silicon with metallic contacts
deposited on the top and bottom. The doping is normally applied to a thin layer on the top of the cell,
producing a p-n junction with a particular bandgap energy, Eg.
Photons that hit the top of the solar cell are either reflected or transmitted into the cell. Transmitted
photons have the potential to give their energy, hν, to an electron if hν ≥ Eg, generating an electron-
hole pair.[13] In the depletion region, the drift electric field Edrift accelerates both electrons and holes
towards their respective n-doped and p-doped regions (up and down, respectively). The
resulting current Ig is called the generated photocurrent. In the quasi-neutral region, the scattering
electric field Escatt accelerates holes (electrons) towards the p-doped (n-doped) region, which gives a
scattering photocurrent Ipscatt (Inscatt). Consequently, due to the accumulation of charges, a
potential V and a photocurrent Iph appear. The expression for this photocurrent is obtained by adding
generation and scattering photocurrents: Iph = Ig + Inscatt + Ipscatt.
The J-V characteristics (J is current density, i.e. current per unit area) of a solar cell under
illumination are obtained by shifting the J-V characteristics of a diode in the dark downward by Iph.
Since solar cells are designed to supply power and not absorb it, the power P = V·Iph must be
negative. Hence, the operating point (Vm, Jm) is located in the region where V>0 and Iph<0, and
chosen to maximize the absolute value of the power |P|.[14]
Loss mechanisms[edit]

The Shockley-Queisser limit for the efficiency of a single-junction solar cell. It is essentially impossible for a
single-junction solar cell, under unconcentrated sunlight, to have more than ~34% efficiency. A multi-junction
cell, however, can exceed that limit.

The theoretical performance of a solar cell was first studied in depth in the 1960s, and is today
known as the Shockley–Queisser limit. The limit describes several loss mechanisms that are
inherent to any solar cell design.
The first are the losses due to blackbody radiation, a loss mechanism that affects any material object
above absolute zero. In the case of solar cells at standard temperature and pressure, this loss
accounts for about 7% of the power. The second is an effect known as "recombination", where
the electrons created by the photoelectric effect meet the electron holes left behind by previous
excitations. In silicon, this accounts for another 10% of the power.
However, the dominant loss mechanism is the inability of a solar cell to extract all of the power in
the light, and the associated problem that it cannot extract any power at all from certain photons.
This is due to the fact that the photons must have enough energy to overcome the bandgap of the
material.
If the photon has less energy than the bandgap, it is not collected at all. This is a major consideration
for conventional solar cells, which are not sensitive to most of the infrared spectrum, although that
represents almost half of the power coming from the sun. Conversely, photons with more energy
than the bandgap, say blue light, initially eject an electron to a state high above the bandgap, but this
extra energy is lost through collisions in a process known as "relaxation". This lost energy turns into
heat in the cell, which has the side-effect of further increasing blackbody losses.[15]
Combining all of these factors, the maximum efficiency for a single-bandgap material, like
conventional silicon cells, is about 34%. That is, 66% of the energy in the sunlight hitting the cell will
be lost. Practical concerns further reduce this, notably reflection off the front surface or the metal
terminals, with modern high-quality cells at about 22%.
Lower, also called narrower, bandgap materials will convert longer wavelength, lower energy
photons. Higher, or wider bandgap materials will convert shorter wavelength, higher energy light. An
analysis of the AM1.5 spectrum, shows the best balance is reached at about 1.1 eV (about 1100 nm,
in the near infrared), which happens to be very close to the natural bandgap in silicon and a number
of other useful semiconductors.
Multi-junction cells[edit]
Cells made from multiple materials layers can have multiple bandgaps and will therefore respond to
multiple light wavelengths, capturing and converting some of the energy that would otherwise be lost
to relaxation as described above.
For instance, if one had a cell with two bandgaps in it, one tuned to red light and the other to green,
then the extra energy in green, cyan and blue light would be lost only to the bandgap of the green-
sensitive material, while the energy of the red, yellow and orange would be lost only to the bandgap
of the red-sensitive material. Following analysis similar to those performed for single-bandgap
devices, it can be demonstrated that the perfect bandgaps for a two-gap device are at 1.1 eV and
1.8 eV.[16]
Conveniently, light of a particular wavelength does not interact strongly with materials that are of
bigger bandgap. This means that you can make a multi-junction cell by layering the different
materials on top of each other, shortest wavelengths (biggest bandgap) on the "top" and increasing
through the body of the cell. As the photons have to pass through the cell to reach the proper layer
to be absorbed, transparent conductors need to be used to collect the electrons being generated at
each layer.

Figure C. (a) The structure of an MJ solar cell. There are six important types of layers: pn junctions, back
surface field (BSF) layers, window layers, tunnel junctions, anti-reflective coating and metallic contacts. (b)
Graph of spectral irradiance E vs. wavelength λ over the AM1.5 solar spectrum, together with the maximum
electricity conversion efficiency for every junction as a function of the wavelength. [14]
Producing a tandem cell is not an easy task, largely due to the thinness of the materials and the
difficulties extracting the current between the layers. The easy solution is to use two mechanically
separate thin film solar cells and then wire them together separately outside the cell. This technique
is widely used by amorphous silicon solar cells, Uni-Solar's products use three such layers to reach
efficiencies around 9%. Lab examples using more exotic thin-film materials have demonstrated
efficiencies over 30%.[16]
The more difficult solution is the "monolithically integrated" cell, where the cell consists of a number
of layers that are mechanically and electrically connected. These cells are much more difficult to
produce because the electrical characteristics of each layer have to be carefully matched. In
particular, the photocurrent generated in each layer needs to be matched, otherwise electrons will be
absorbed between layers. This limits their construction to certain materials, best met by the III-V
semiconductors.[16]
Material choice[edit]
The choice of materials for each sub-cell is determined by the requirements for lattice-matching,
current-matching, and high performance opto-electronic properties.
For optimal growth and resulting crystal quality, the crystal lattice constant a of each material must
be closely matched, resulting in lattice-matched devices. This constraint has been relaxed somewhat
in recently developed metamorphic solar cells which contain a small degree of lattice mismatch.
However, a greater degree of mismatch or other growth imperfections can lead to crystal defects
causing a degradation in electronic properties.
Since each sub-cell is connected electrically in series, the same current flows through each junction.
The materials are ordered with decreasing bandgaps, Eg, allowing sub-bandgap light (hc/λ < e·Eg) to
transmit to the lower sub-cells. Therefore, suitable bandgaps must be chosen such that the design
spectrum will balance the current generation in each of the sub-cells, achieving current matching.
Figure C(b) plots spectral irradiance E(λ), which is the source power density at a given wavelength λ.
It is plotted together with the maximum conversion efficiency for every junction as a function of the
wavelength, which is directly related to the number of photons available for conversion into
photocurrent.
Finally, the layers must be electrically optimal for high performance. This necessitates usage of
materials with strong absorption coefficients α(λ), high minority carrier lifetimes τminority, and high
mobilities µ.[17]
The favorable values in the table below justify the choice of materials typically used for multi-junction
solar cells: InGaP for the top sub-cell (Eg = 1.8 − 1.9 eV), InGaAs for the middle sub-cell (Eg = 1.4
eV), and Germanium for the bottom sub-cell (Eg = 0.67 eV). The use of Ge is mainly due to its lattice
constant, robustness, low cost, abundance, and ease of production.
Because the different layers are closely lattice-matched, the fabrication of the device typically
employs metal-organic chemical vapor deposition (MOCVD). This technique is preferable to
the molecular beam epitaxy (MBE) because it ensures high crystal quality and large scale
production.[14]

absorption Hardness
Material Eg, eV a, nm µ , cm2/(V·s) τp, µs α, µm/K S, m/s
(λ = 0.8 μm), 1/µm n (Mohs)

c-Si 1.12 0.5431 0.102 1400 1 7 2.6 0.1–60


InGaP 1.86 0.5451 2 500 – 5 5.3 50

GaAs 1.4 0.5653 0.9 8500 3 4–5 6 50

Ge 0.65 0.5657 3 3900 1000 6 7 1000

InGaAs 1.2 0.5868 30 1200 – – 5.66 100–1000

Structural elements[edit]
Metallic contacts[edit]
The metallic contacts are low-resistivity electrodes that make contact with the semiconductor layers.
They are often aluminum. This provides an electrical connection to a load or other parts of a solar
cell array. They are usually on two sides of the cell. And are important to be on the back face so that
shadowing on the lighting surface is reduced.
Anti-reflective coating[edit]
Anti-reflective (AR) coating is generally composed of several layers in the case of MJ solar cells. The
top AR layer has usually a NaOH surface texturation with several pyramids in order to increase the
transmission coefficient T, the trapping of the light in the material (because photons cannot easily
get out the MJ structure due to pyramids) and therefore, the path length of photons in the
material.[12] On the one hand, the thickness of each AR layer is chosen to get destructive
interferences. Therefore, the reflection coefficient R decreases to 1%. In the case of two AR
layers L1 (the top layer, usually SiO
2) and L2 (usually TiO

2), there must be to have the same amplitudes for reflected fields and nL1dL1 = 4λmin,nL2dL2 = λmin/4
to have opposite phase for reflected fields.[18] On the other hand, the thickness of each AR layer is
also chosen to minimize the reflectance at wavelengths for which the photocurrent is the lowest.
Consequently, this maximizes JSC by matching currents of the three subcells.[19] As example, because
the current generated by the bottom cell is greater than the currents generated by the other cells, the
thickness of AR layers is adjusted so that the infrared (IR) transmission (which corresponds to the
bottom cell) is degraded while the ultraviolet transmission (which corresponds to the top cell) is
upgraded. Particularly, an AR coating is very important at low wavelengths because, without
it, T would be strongly reduced to 70%.
Tunnel junctions[edit]
Figure D: Layers and band diagram of the tunnel junction. Because the length of the depletion region is narrow
and the band gap is high, electrons can tunnel.

The main goal of tunnel junctions is to provide a low electrical resistance and optically low-loss
connection between two subcells.[20] Without it, the p-doped region of the top cell would be directly
connected with the n-doped region of the middle cell. Hence, a pn junction with opposite direction to
the others would appear between the top cell and the middle cell. Consequently,
the photovoltage would be lower than if there would be no parasitic diode. In order to decrease this
effect, a tunnel junction is used.[21] It is simply a wide band gap, highly doped diode. The high doping
reduces the length of the depletion region because

Hence, electrons can easily tunnel through the depletion region. The J-V characteristic of the
tunnel junction is very important because it explains why tunnel junctions can be used to have a
low electrical resistance connection between two pn junctions. Figure D shows three different
regions: the tunneling region, the negative differential resistance region and the thermal diffusion
region. The region where electrons can tunnel through the barrier is called the tunneling region.
There, the voltage must be low enough so that energy of some electrons who are tunneling is
equal to energy states available on the other side of the barrier. Consequently, current density

through the tunnel junction is high (with maximum value of , the peak current density) and
the slope near the origin is therefore steep. Then, the resistance is extremely low and
consequently, the voltage too.[22] This is why tunnel junctions are ideal for connecting two pn
junctions without having a voltage drop. When voltage is higher, electrons cannot cross the
barrier because energy states are no longer available for electrons. Therefore, the current
density decreases and the differential resistance is negative. The last region, called thermal
diffusion region, corresponds to the J-V characteristic of the usual diode:

In order to avoid the reduction of the MJ solar cell performances, tunnel junctions must be
transparent to wavelengths absorbed by the next photovoltaic cell, the middle cell, i.e.
EgTunnel > EgMiddleCell.
Window layer and back-surface field[edit]
Figure E: (a) Layers and band diagram of a window layer. The surface recombination is reduced. (b)
Layers and band diagram of a BSF layer. The scattering of carriers is reduced.

A window layer is used in order to reduce the surface recombination velocity S. Similarly, a
back-surface field (BSF) layer reduces the scattering of carriers towards the tunnel junction.
The structure of these two layers is the same: it is a heterojunction which catches electrons
(holes). Indeed, despite the electric field Ed, these cannot jump above the barrier formed by
the heterojunction because they don't have enough energy, as illustrated in figure E. Hence,
electrons (holes) cannot recombine with holes (electrons) and cannot diffuse through the
barrier. By the way, window and BSF layers must be transparent to wavelengths absorbed
by the next pn junction i.e. EgWindow > EgEmitter and EgBSF > EgEmitter. Furthermore, the lattice
constant must be close to the one of InGaP and the layer must be highly doped (n ≥
1018 cm−3).[23]
J-V characteristic[edit]
For maximum efficiency, each subcell should be operated at its optimal J-V parameters,
which are not necessarily equal for each subcell. If they are different, the total current
through the solar cell is the lowest of the three. By approximation,[24] it results in the same
relationship for the short-circuit current of the MJ solar cell: JSC = min (JSC1, JSC2,
JSC3) where JSCi(λ) is the short-circuit current density at a given wavelength λ for the subcell i.
Because of the impossibility to obtain JSC1, JSC2, JSC3 directly from the total J-V characteristic,
the quantum efficiency QE(λ) is utilized. It measures the ratio between the amount of
electron-hole pairs created and the incident photons at a given wavelength λ. Let φi(λ) be
the photon flux of corresponding incident light in subcell iandQEi(λ) be the quantum
efficiency of the subcell i. By definition, this equates to:[25]

The value of is obtained by linking it with the absorption coefficient , i.e. the
number of photons absorbed per unit of length by a material. If it is assumed that each
photon absorbed by a subcell creates an electron/hole pair (which is a good
approximation), this leads to:[23]
where di is the thickness of the subcell i and is the percentage of incident light
which is not absorbed by the subcell i.
Similarly, because

, the following approximation can be used: .

The values of are then given by the J-V diode equation:

Theoretical limiting efficiency[edit]


We can estimate the limiting efficiency of ideal infinite multi-junction solar
cells using the graphical quantum-efficiency (QE) analysis invented by C. H.
Henry.[26] To fully take advantage of Henry's method, the unit of the AM1.5
spectral irradiance should be converted to that of photon flux (i.e., number
of photons/m2/s). To do that, it is necessary to carry out an intermediate unit
conversion from the power of electromagnetic radiation incident per unit
area per photon energy to the photon flux per photon energy (i.e., from
[W/m2/eV] to [number of photons/m2/s/eV]). For this intermediate unit
conversion, the following points have to be considered: A photon has a
distinct energy which is defined as follows.

(1): Eph = h∙f = h∙(c/λ)

where Eph is photon energy, h is Planck's constant (h = 6.626*10−34 [J∙s]), c


is speed of light (c = 2.998*108 [m/s]), f is frequency [1/s], and λ is
wavelength [nm].
Then the photon flux per photon energy, dnph/dhν, with respect to certain
irradiance E [W/m2/eV] can be calculated as follows.

(2): = E/{h∙(c/λ)} = E[W/(m2∙eV)]∙λ∙(10−9 [m])/(1.998∙10−25 [J∙s∙m/s]) =


E∙λ∙5.03∙1015 [(# of photons)/(m2∙s∙eV)]

As a result of this intermediate unit conversion, the AM1.5 spectral


irradiance is given in unit of the photon flux per photon energy, [number of
photons/m2/s/eV], as shown in Figure 1.

Figure 1. Photon flux per photon energy from standard solar energy spectrum
(AM of 1.5).
Based on the above result from the intermediate unit conversion, we can
derive the photon flux by numerically integrating the photon flux per photon
energy with respect to photon energy. The numerically integrated photon
flux is calculated using the Trapezoidal rule, as follows.

(3):
As a result of this numerical integration, the AM1.5 spectral irradiance is
given in unit of the photon flux, [number of photons/m2/s], as shown in
Figure 2.

Figure 2. Photon flux from standard solar energy spectrum (AM of 1.5).
There are no photon flux data in the small photon energy range from 0 eV
to 0.3096 eV because the standard (AM1.5) solar energy spectrum for hν <
0.31 eV are not available. Regardless of this data unavailability, however,
the graphical QE analysis can be done using the only available data with a
reasonable assumption that semiconductors are opaque for photon
energies greater than their bandgap energy, but transparent for photon
energies less than their bandgap energy. This assumption accounts for the
first intrinsic loss in the efficiency of solar cells, which is caused by the
inability of single-junction solar cells to properly match the broad solar
energy spectrum. However, the current graphical QE analysis still cannot
reflect the second intrinsic loss in the efficiency of solar cells, radiative
recombination. To take the radiative recombination into account, we need to
evaluate the radiative current density, Jrad, first. According to Shockley and
Queisser method,[27] Jrad can be approximated as follows.

(4):

(5):
where Eg is in electron volts and n is evaluated to be 3.6, the value for
GaAs. The incident absorbed thermal radiation Jth is given by Jrad with V = 0.

(6):
The current density delivered to the load is the difference of the current
densities due to absorbed solar and thermal radiation and the current
density of radiation emitted from the top surface or absorbed in the
substrate. Defining Jph = enph, we have
(7): J = Jph + Jth − Jrad
The second term, Jth, is negligible compared to Jph for all semiconductors
with Eg. ≥ 0.3 eV, as can be shown by evaluation of the above Jth equation.
Thus, we will neglect this term to simplify the following discussion. Then we
can express J as follows.
(8):

The open-circuit voltage is found by setting J = 0.

(9):

The maximum power point (Jm, Vm) is found by stetting the derivative .
The familiar result of this calculation is

(10):

(11):
Finally, the maximum work (W m) done per absorbed photon, Wm is given by

(12):
Combining the last three equations, we have

(13):
Using the above equation, W m (red line) is plotted in Figure 3 for different
values of Eg (or nph).

Figure 3. Maximum work by ideal infinite multi-junction solar cells under


standard AM1.5 spectral irradiance.
Now, we can fully use Henry's graphical QE analysis, taking into account
the two major intrinsic losses in the efficiency of solar cells. The two main
intrinsic losses are radiative recombination, and the inability of single
junction solar cells to properly match the broad solar energy spectrum. The
shaded area under the red line represents the maximum work done by ideal
infinite multi-junction solar cells. Hence, the limiting efficiency of ideal
infinite multi-junction solar cells is evaluated to be 68.8% by comparing the
shaded area defined by the red line with the total photon-flux area
determined by the black line. (This is why this method is called "graphical"
QE analysis.) Although this limiting efficiency value is consistent with the
values published by Parrott and Vos in 1979: 64% and 68.2%
respectively,[28][29] there is a small gap between the estimated value in this
report and literature values. This minor difference is most likely due to the
different ways how to approximate the photon flux from 0 eV to 0.3096 eV.
Here, we approximated the photon flux from 0 eV to 0.3096 eV as the same
as the photon flux at 0.31 eV.

Materials[edit]
The majority of multi-junction cells that have been produced to date use
three layers (although many tandem a-Si:H/mc-Si modules have been
produced and are widely available). However, the triple junction cells
require the use of semiconductors that can be tuned to specific frequencies,
which has led to most of them being made of gallium arsenide (GaAs)
compounds, often germanium for the bottom-, GaAs for the middle-, and
GaInP2 for the top-cell.
Gallium arsenide substrate[edit]
Dual junction cells can be made on Gallium arsenide wafers. Alloys
of Indium gallium phosphide in the range In.5Ga.5P through In.53Ga.47P serve
as the high band gap alloy. This alloy range provides for the ability to have
band gaps in the range of 1.92eV to 1.87eV. The lower GaAs junction has a
band gap of 1.42eV.[citation needed]
Germanium substrate[edit]
Triple junction cells consisting of indium gallium phosphide (InGaP), gallium
arsenide (GaAs) or indium gallium arsenide (InGaAs) and germanium (Ge)
can be fabricated on germanium wafers. Early cells used straight gallium
arsenide in the middle junction. Later cells have utilized In0.015Ga0.985As, due
to the better lattice match to Ge, resulting in a lower defect density.[citation needed]
Due to the huge band gap difference between GaAs (1.42eV), and Ge
(0.66eV), the current match is very poor, with the Ge junction operated
significantly current limited.[citation needed]
Current efficiencies for commercial InGaP/GaAs/Ge cells approach 40%
under concentrated sunlight.[30][31] Lab cells (partly using additional junctions
between the GaAs and Ge junction) have demonstrated efficiencies above
40%.[32]
Indium phosphide substrate[edit]
Indium phosphide may be used as a substrate to fabricate cells with band
gaps between 1.35eV and 0.74eV. Indium Phosphide has a band gap of
1.35eV. Indium gallium arsenide (In0.53Ga0.47As) is lattice matched to Indium
Phosphide with a band gap of 0.74eV. A quaternary alloy of Indium gallium
arsenide phosphide can be lattice matched for any band gap in between the
two.[citation needed]
Indium phosphide-based cells have the potential to work in tandem with
gallium arsenide cells. The two cells can be optically connected in series
(with the InP cell below the GaAs cell), or in parallel through the use of
spectra splitting using a Dichroic filter.[citation needed]
Indium gallium nitride substrate[edit]
Indium gallium nitride (InGaN) is a semiconductor material made of a mix of
gallium nitride (GaN) and indium nitride (InN). It is a ternary group
III/V direct bandgap semiconductor. Its bandgap can be tuned by varying
the amount of indium in the alloy from 0.7 eV to 3.4 eV, thus making it an
ideal material for solar cells.[33] However, its conversion efficiencies because
of technological factors unrelated to bandgap are still not high enough to be
competitive in the market.[34][35]

You might also like