Part 2 Avasthi
Part 2 Avasthi
Silicon/Organic Heterojunction to
Block Minority Carriers
4.1 Introduction
The best silicon solar cells are typically fabricated on thin high-quality silicon wafers
(< 200µm) so most of the minority carrier recombination losses happen at silicon
surfaces, specifically the silicon/metal contacts (Fig. 4.1(a)). Conventional silicon
solar cells use diffused p/p+ back-surface fields to reduce the surface recombina-
tion (Fig. 4.1(b)). A potentially cheaper and more effective alternative to diffused
back-surface fields could be a minority-carrier blocking silicon/organic heterojunction
(Chapter 2), e.g as an electron-blocking p-type contact at the anode (Fig. 4.1(c)).
To demonstrate such a heterojunction, two critical issues were needed to be solved:
midgap states at unpassivated Si surfaces and band-alignment at Si/organic interface.
The former can be solved using using PQ-passivated Si surfaces (Chapter 3), but
finding an organic with the correct band-alignment is still an open problem. The two
specific band-alignment criteria for an electron-blocking layer are:
(a) The LUMO of the organic should be much higher than conduction band edge
78
Surface Reduced Surface
Recombination Recombination
Dark Current Dark Current
Photocurrent − Photocurrent − p+-Si
EC p-Si EC p-Si
Cathode Cathode
Anode Anode
n+-Si n+-Si
EV
Photocurrent EV
Photocurrent
+ +
Dark Current Dark Current
(a) (b)
Reduced Surface
Recombination
LUMO
Dark Current Organic
Photocurrent − Passsivation
EC p-Si
Cathode
Anode
n+-Si
HOMO
EV
Photocurrent
+
Dark Current
(c)
Figure 4.1: (a) Band Diagram of solar cell in which recombination at the metal
contact dominates. The solid lines are desired direction of photogenerated carrier
flow and the dashed line represents dark current, or equivalently the loss mechanism.
(b) Effect of a p/p+ back surface field in reducing the electron recombination at the
anode. (c) Effect of a suitable silicon/organic heterojunction that reduces the electron
recombination at the anode due to the offset in the conduction band.
of silicon, i.e. there should be a large conduction band offset, so that electrons
are repelled away from the surface,
(b) The HOMO and the valence band edge of Si should be aligned, i.e. there should
be almost no valence band offset, so that photogenerated holes can be efficiently
extracted at the anode.
79
n+ Si
Anode
5 × 10 cm−3 500 nm
19
p Si Substrate
1016 cm−3 200 µm
Cathode Organic 10 nm
(a) (b)
80
to zero, similar to what is expected in a short-base diode (Fig. 4.3(a)). With the het-
erojunction, the electron density increases by several orders of magnitude because it
is no longer pinned to zero at the contacts. Since the diffusion current of the electrons
in p-type silicon is directly proportional to the slope of electron density profile, J0 of
the heterojunction device is lower than an equivalent device with no heterojunction
(Fig. 4.3(b)). Looking at the simulated current-voltage characteristics, an issue that
is immediately apparent is the high series-resistance of the heterojunction device. At
higher currents (>10 mA/cm2 ), ohmic losses in the organic layer are the primary
reason for the series resistance. This could be a potential issue in a solar cell.
To confirm the importance of surface defects at the silicon/organic heterojunc-
tion, simulations were also performed with a varying values of surface recombination
velocity (SRV). As expected, at higher recombination velocities, the electron density
at the contact decreases and the electron current increases (Fig. 4.3(b)).
The total current in the simulated device consists of two components, electron
current and hole current. At a forward bias of 0.5 V the hole and electron-current
components of the device were separately plotted as a function of surface recom-
bination velocity (Fig. 4.3(c)). Also shown in the data are the asymptotic values:
the minimum current obtained at an SRV of zero and the maximum current from a
Si/metal device (no heterojunction). While the electron current continues to decrease
as the SRV at the silicon/organic interface decreases, the total current stagnates at
≈ 35 µA. This is because the hole current from the p-type to the n-type region does
not depend on the SRV of the p-type contact and at a SRV of ∼30 cm/s the total
current becomes dominated by holes. This highlights an issue that will reoccur again,
a single heterojunction can reduce only one of the carriers, electron or holes, and the
device performance soon becomes limited by the current due to the other carrier.
Finally, the analytical expression for electron current, mentioned in Section 2.4.1
and derived in Appendix A.5, is compared with the simulated electron current.
81
(a) (b)
(c) (d)
Figure 4.3: (a) Simulated electron-density profile and (b) simulated current-voltage
characteristics of a n+ -p diode with and without the electron-blocking organic layer.
Curves for the non-ideal Si/organic interface, with a surface recombination velocity
of 500 cm/s, is also shown. (c) Contribution of hole and electron current to the total
current at 0.5 V forward bias. (d) Comparison between the simulated and analytically
derived (Eq. (2.4)) values of electron current at 0.5 V forward bias.
Clearly, the theory and simulations qualitatively agree over a wide range of SRV.
82
shown in Fig. 4.2(a), and see if J0 is reduced. However fabricating high-quality n+ -p
diodes with millisecond bulk lifetimes requires access to high-quality Si wafers and
ultra-clean furnaces. Furthermore, to validate the performance of Si/organic hetero-
junction, we want to measure a decrease in the device J0 . If the J0 of the fabricated
devices varies from sample-to-sample, due to differences in processing and/or wafer
quality, the experiments will be inconclusive. In order to reliably compare the perfor-
mance of different silicon/organic heterojunctions for the role of a electron-blocking
p-type contact, a test structure that is very sensitive to surface recombination of
minority-carriers (a “Minority Carrier Probe”) is required.
4.3.1 Design
The structure designed for this purpose is shown in Fig. 4.4(a). Simulations were
used to optimize the doping levels and thicknesses of all the layers, such that total
current is dominated by the electron injection current from n+ to p layer (Fig. 4.4c.
Additionally the top p-Si implant layer was kept thin so that most of the electron
recombination happens at the top Si/metal interface (diffusion length of electrons >>
thickness of the p-Si implant layer).
If an electron-blocking silicon/organic heterojunction is introduced at the p-type
contact between silicon and the anode (Fig. 4.5(a)), the electron recombination at
the Si/metal interface will decrease and a lower J0 will be measured as confirmed by
the simulated current-voltage characteristics (Fig. 4.5(c)).
The surface recombination velocity at the Si/organic interface is parameter that
is least under our control. To increase the chances of observing a decrease in electron
recombination, the minority-carrier probe should should work over a large range of
surface recombination velocities. Suppose the probe is too sensitive to surface recom-
bination, capable of showing a lower of J0 only when SRV <50 cm/s, no electron-
blocking effect would ever be observed at PQ-passivated silicon surface (SRV ∼100
83
90 µm
100 µm
p Si Implant
Anode 3 × 1017 cm−3 (500 nm)
n Si Epi
5 × 1016 cm−3 (5 µm)
n+ Si Substrate
2 × 1019 cm−3 5 µm
Cathode
200 µm
(a)
(b) (c)
Figure 4.4: (a) The structure of the minority-carrier probe used to test electron-
blocking in silicon/organic heterojunctions. (b) The band diagram of the minority-
carrier probe. (c) The simulated forward-bias current-voltage characteristics of the
minority-carrier probe. The electron current dominates the hole current.
cm/s, measured in Chapter 3). Simulations with different values of Si/organic recom-
bination velocity show that the designed minority-carrier probe is fairly insensitive to
SRV, and will show a decrease in J0 if SRV <10000, good enough for PQ-passivated
surfaces (Fig. 4.4(a)).
A practical minority-carrier probe that follows the exact specifications of Fig. 4.4(a)
should be sensitive to reduction in electron current, i.e. it should have a current se-
84
90 µm
100 µm
Organic 10 nm
p Si Implant
Anode
3 × 1017 cm−3 (500 nm)
n Si Epi
5 × 1016 cm−3 5 µm
n+ Si Substrate
2 × 1019 cm−3 5 µm
Cathode
200 µm
(a) (b)
(c) (d)
Figure 4.5: (a) The device structure and (b) band diagram of the minority-carrier
probe with a silicon/organic heterojunction at the p-type contact. (c) The simu-
lated forward-bias current-voltage characteristics showing reduction in current. (d)
The simulated forward-bias current-voltage characteristics at different levels of silicon
surface recombination velocity.
lectivity ratio (Ielectron /Ihole ) of >10. However, practical devices always have some
deviations from the intended structure. To insure that the minority-carrier probe
design is robust against such unintended variations, dependence of selectivity on pa-
rameters, like doping, carrier lifetime, and implant depth were also analyzed, As
shown in Fig. 4.6, simulations reveal that the design of the probe is quite robust and
85
(a) (b)
(c) (d)
Figure 4.6: The selectivity of the designed minority-carrier probe as a function of (a)
implant doping, (b) implant depth, (c) Epi layer doping, and (d) average minority-
carrier lifetime.
electron current dominates the hole current by at least a factor of ten for a wide range
of parameters.
4.3.2 Fabrication
86
to grow a 6 µm thick epitaxial layer of silicon. The precursor gas was dichlorosilane
(26 sccm) diluted in Hydrogen (4 slpm). The growth temperature was 1000 (29
% lamp power) and growth rate was approximately 200 nm/min. The epitaxial layer
was unintentially doped n-type (<5×1016 cm-3 ). Next, a 750 nm thick wet oxide
was grown at 1050 in the Thermco furnace. The oxide layer served as hard mask
for the subsequent ion-implantation. Photolithography was used to etch holes into
the oxide layer to define the implant area. Patterned samples were implanted with
boron at Core Systems, CA. Unless otherwise stated, the implant dosage and energy
were 1013 cm-2 and 50 keV, respectively. After implantation, samples were cleaned
and again loaded into the Thermco furnace, this time for dopant activation and
annealing. Annealing temperatures were around 1000 and annealing time was 20
minutes. Samples were always introduced into the furnace in an oxygen atmosphere,
and after 10 minutes nitrogen was turned on. To estimate the final doping profiles
of the structure, simulations were done in Taurus TSuprem4 (a Synopsys® tool).
The expected doping profile of the structure, before and after annealing, is shown
in Fig. 4.7(a). To form the Si/metal structure (Fig. 4.7), samples were dipped in a
10:1 buffered HF solution for 10s to remove any native oxide and then Ag or Al was
deposited by thermal evaporation.
4.3.3 Characterization
Before the minority-carrier probe can be used to test Si/organic heterojunctions, its
baseline performance needs to be characterized. The critical questions to demonstrate
performance of the minority-carrier probe are:
(a) Is the electron current the dominant component of the total current?
(b) Is the device in short-base condition? i.e. do most of the electrons recombine,
at the Si/metal interface?
87
(a)
p Si Implant
Ag (50 nm) 3 × 1017 cm−3 (400 nm)
Dark Current −
p-Si
EC
n Si Epi n-Si Cathode
n+-Si
5 × 10 cm−3 (6 µm)
16
Anode
n+ Si Substrate EV
2 × 1019 cm−3
(500 µm)
Al (100 nm)
(b) (c)
Figure 4.7: (a) Simulated doping profile of the implanted region before and after
annealing. The implant dosage and energy was 1013 cm-2 and 50 keV, respectively.
(b) Structure and (c) band-diagram of the minority-carrier probe.
(c) Are the current-voltage characteristics of the diodes ideal? Specifically, can
values of J0 be extracted reliably from the data?
In a short-base diode, the electron current scales with p-type doping (NA ) and length
of the quasi-neutral region (W ):
1
Jelectron ∝ (4.1)
NA W
88
Table 4.1: Implant doses, energies, and annealing conditions used to fabricate devices
with different p-type doping. The measured value of J0 are also given. Device area
is 200 µm × 200 µm
(a) (b)
Figure 4.8: Effect of increasing implant doping on the characteristics of the minority-
carrier probe. (a) Simulated doping profiles of the structures after annealing at 1000
. (b) Measured I-V characteristics for the three devices. The device area is 200 µm
× 200 µm.
89
levels in these devices is ∼ 1017 , 1018 , and 1019 cm-3 .
Current-voltage characteristics of the finished Si/metal diodes are shown in Fig. 4.8(b).
As expected, the highest current is measured in the sample with p- doping of ∼ 1017
cm-3 and a junction depth of 400 nm. As the doping level and junction depth are
increased to ∼ 1018 cm-3 and 560, respectively, the current falls by a factor of 10.
When the doping is increased even further (to ∼ 1019 cm-3 ), the current does reduce
more, but only marginally. This probably signifies that the diode is now dominated
by hole current or the bandgap narrowing in the heavily-doped layer is very signifi-
cant. Overall the data confirms that in the ∼ 1017 cm-3 doped structure, the electron
current is at least 10 times larger than the hole current.
One possible objection to the implant test is that all three samples were processed
completely separately during the implant and anneal steps and the difference in I-V
characteristics may just be an artifact of unaccounted variations in process conditions
and wafer quality. Another, and possibly more serious objection is that the test does
not tell if the diode is in short-base regime or not. Current reduction would be
observed even if most of the electrons recombine in the bulk and not the surface.
To work around this issue, two implanted and annealed devices were chosen from
the same silicon wafer. One of the devices was put back into the RTCVD and a 200
nm thick highly-doped p-Si epitaxial layer was grown on top of the p-type implant
layer (Fig. 4.9(a)). Finally, metal electrodes were deposited on both of the samples.
The device with the extra epitaxial layers has as surface field due to the p-p+
junction which is expected to reduce the electron recombination at the anode in
much the same way as a silicon/organic heterojunction. Significantly, the epitaxy
process is only 1 min long (at 1000 ), so no major changes in the doping or depth of
the p- /n- junction are expected. A decrease in current can only happen if the diode
90
Reduced Surface
p+ Si Epi Recombination
1019 cm−3 p+-Si
(200 nm) p Si Implant Epi
Ag (50 nm) 3 × 1017 cm−3 (400 nm)
Dark Current −
p-Si
EC
Cathode
n Si Epi n-Si n+-Si
5 × 1016 cm−3 (6 µm) Anode
n+ Si Substrate EV
2 × 1019 cm−3
(500 µm)
Al (100 nm)
(a) (b)
(c) (d)
Figure 4.9: (a) Structure and (b) band-diagram of the ‘Epitaxial’ device. (c) The
simulated IV characteristics of the minority-carrier probe with different p-type con-
tacts: silicon/metal, ideal silicon/organic heterojunction, and p-p+ homojunction.
(d) Measured I-V characteristics of devices with and without the ‘Epi’ layer. The
device area is 200 µm × 200 µm.
91
Metal Implant
180 µm 200 µm
180 µm’ 10 µm
90 µm 200 µm
Figure 4.10: Do currents in minority-carrier probe scale with contact area? (a) Top
view of the minority-carrier devices with different contact shapes. Black area repre-
sents metal. (b) Measured I-V characteristics of devices with different contact shapes.
Table 4.2: Test if current in the minority-carrier probes of Fig. 4.10 scales with contact
area or contact perimeter. Implanted area is fixed at 4×10-4 cm2 . Ideality factor of
the diodes was extracted from data between 0.2-0.4V.
Contact Current
Current
Device at 0.4V n
Area Scales Perimeter Scales scales by
µ
( m2 ) by µ
( m) by (nA)
µ
‘90 m, small P’ 8100 1 360 1 34 1 1.00
µ
‘90 m, large P’ 8100 1 3240 9 26 0.8 1.01
µ
‘180 m’ 32400 4 720 2 111 2.9 1.00
current. The experiment reaffirms that electron injection is the dominant current
carrying mechanism in the minority-carrier probe and conclusively shows that the
I-V characteristics are sensitive to electron recombination at the top silicon surface.
92
Ametal Aoxide
Metal Oxide
Imetal Ioxide
p-Si
− −
n-Si Ijunction
− −
Metal
Ajunction
Figure 4.11: Explanation for dependence of total current on Si/metal area. The total
electron injection current (Ijunction ) and its two components: Jmetal and Joxide
planted area (200 µm × 200 µm) but different Si/metal contacts (Fig. 4.10(a)). The
device named ‘180 µm’ is the same structure as the standard Si/metal device used
before. The devices, named ‘90 µm, small P’ and ‘90 µm, large P’, both have the
same contact area (90 µm × 90 µm) but different perimeter. The perimeter of ‘90
µm, large P’ is 9 times larger (‘large P’) than ‘90 µm, small P’.
If the Si/metal interface has any edge effects, the current in the ‘small P’ and ’large
P’ devices would be substantially different. The measured current-voltage character-
istics show that this is not the case (Fig. 4.10(b)). The differences in currents between
‘small P’ and ’large P’ device are minimal - at 0.4 V the currents in the two diodes
are 34 nA and 26 nA, respectively (Table 4.2).
The currents in the ‘90 µm, small P’ device and the ‘180 µm’ device are not equal,
even though the junction area is equal. At 0.4 V, the current in the ‘90 µm, small
P’ device is 34 nA while in ‘180 µm’ device it is 111 nA - a factor of 2.9 different.
The only difference in the two devices, to justify the different currents, is the contact
area which is 4× larger in the latter device. This dependence on Si/metal contact
area is not surprising because the minority-carrier probe is dominated by electron
recombination at the top electrode and larger metal area should increase electron
current.
A simple one-dimensional model of the p-n junction diode does not capture this
93
dependence. To visualize the argument, consider a simple two dimensional p-n junc-
tion (Fig. 4.11). Assume that all minority-carrier current in the p-region of the diode
flows vertically, i.e. there is no lateral current flow. The total current flowing across
the junction, characterized by J0,junction , is a area-weighted sum of two components:
the current recombining at the metal (characterized by a J0,metal ) and current recom-
bining at the oxide (characterized by a J0,oxide ).
where Ametal and & Aoxide represents the area covered with metal and oxide, respec-
tively. If the oxide passivation is as bad as the Si/metal interface, i.e. Joxide = Jmetal ,
and total current scales with total area (I0,junction = J0,metal (Ametal + Aoxide )). How-
ever, if the Si/oxide interface is perfectly passivated, i.e. Joxide =0, then I0,junction =
Jmetal Ametal , and the device current scales with area of the contact metal, not area
of the p-n junction. Practical devices are expected to lie somewhere in the middle.
This is confirmed in the I-V characteristics of devices, ‘90 µm, small P’ and ‘180 µm’,
where the total current in the latter device is 2.9× bigger (Table 4.2).
From the given value of Ametal and Aoxide (=Atotal - Ametal ), we can estimate J0,oxide
and J0,metal using Eq. (4.2).
7600 32400
J0,‘180µm0 = J0,oxide + J0,metal = 4.7 × 10−11
40000 40000
31900 8100
J0,‘90µm,smallP 0 = J0,oxide + J0,metal = 1.8 × 10−11
40000 40000
J0,oxide = 8.3 × 10−12 A/cm2 & J0,metal = 5.6 × 10−11 A/cm2 (4.3)
Non-idealities can also arise at the edge of the p-n junction due to un-annealed
94
implant damage or trapped charges at the Si/oxide interface. In the I-V characteris-
tics, the tell-tale sign of these issues is evidence of currents that scales with the device
perimeter. To test for any perimeter effects at the edges of the implanted region, de-
vices with different implanted areas were fabricated: ‘200 µm’, ‘50 µm’, and ‘20 µm’
(Fig. 4.12(a)).
If the diodes are ideal and there are no edge-effects then the simple model derived
above (Fig. 4.11) would fully describe the currents. Using values of Joxide and Jmetal
calculated in Eq. (4.3) along with Eq. (4.2), the ‘expected’ scaling factor between ‘200
µm’ and ‘50 µm’ devices in the absence of edge effects is:
= 29 (4.4)
Between ‘20 µm’ and ‘50 µm’, expected scaling factor is:
= 0.13 (4.5)
The measured I-V characteristics of the three devices is given in Fig. 4.12(b). At
0.4 V, the currents measured in the ‘200 µm’, ‘50 µm’, and ‘20 µm’ device is 110
nA, 3.5 nA , and 0.13 nA, i.e. scaling factor of 31 and 0.14 (Table 4.3). These
experimentally measured values are remarkably close to what our model predicts
(Eq. (4.4) & Eq. (4.5)). This proves not only the predictive power of the model but
also confirms that perimeter effects are indeed absent in the “minority-carrier probe”
devices.
In summary, the designed minority-carrier probe has near-ideal characteristics: an
ideality factor of ≈1, and a J0 that predictably scales with area.
95
180 µm 200 µm
Metal Implant
‘200 µm’
30 µm
10 µm
50 µm 20 µm
Figure 4.12: Do currents in minority-carrier probe scale with area? (a) Top view of
the minority-carrier devices with different implant area. Black area represents metal.
(b) Measured I-V characteristics of devices with different implant areas.
Table 4.3: Test if the current in the minority-carrier probes of Fig. 4.8 scales with
device area or contact area. Ideality factor of the diodes was extracted from data
between 0.2-0.4V.
µ
‘200 m’ 40000 16 32400 36 110 29 31 1.00
µ
‘50 m’ 2500 1 900 1 3.5 1 - 1.01
µ
‘20 m’ 400 0.16 100 0.11 0.51 0.13 0.14 1.08
96
4.4.1 Materials and Fabrication
Devices from the same implanted and annealed wafer were processed in three different
ways: Si/metal diode (Fig. 4.13(a) & (b)), Si/p+ -Epi/metal diode (Fig. 4.13(c) & (d)),
and Si/TPD/metal heterojunction diode (Fig. 4.13(e) & (f)). The I-V characteristic
of these structures is shown in the Fig 4.14.
The metal-silicon diode is the control structure, in which the dark-current is mostly
due to recombination of electron at the Si/anode interface. The Si/p+ -Epi/metal
diode was fabricated by growing a 200nm highly-doped silicon layer (∼1019 cm-3 )
epitaxially on the p-Si surface before depositing the top metal. The thick p+ epitaxial
Si layer increases the surface field near the anode, which repels electrons and reduces
electron recombination. Consequently, the current of the p+ -Epi diode is around 5
times lower than the control structure: at 0.3 V, currents fall from 2.2 nA to 0.34 nA.
The extracted value of J0 , from the y-intercept on the semilog I-V plot, in ‘Epi’ diode
is around 9.8 pA/cm2 instead of 47 pA/cm2 for the Si/metal diodes (Table 4.4).
97
The silicon/TPD/metal heterojunction was fabricated by depositing TPD on PQ-
passivated silicon. Since PQ layer is thin, it was assumed that energy levels of PQ
layers will not matter and the carriers will just tunnel through PQ (This ultimately
turned out to be an incorrect assumption. For details please look at Chapter 5). TPD
is a hole-conducting organic with reported HOMO & LUMO levels at 5.3 eV and 2.3
eV, respectively. Silicon/TPD is expected to be an electron-blocking heterojunction
[105], so if it is introduced between silicon and the top metal in the minority carrier
probe device, the electron recombination at the top electrode is expected to decrease.
Experimentally, currents in the heterojunction device are lower: at 0.3 V, the
current in heterojunction device is only 0.34 nA. The extracted J0 of the hetero-
junction device is around 23 pA/cm2 . Thus compared to Si/metal device, currents
in the heterojunction device are reduced by a factor of at least two, proving the
electron-blocking nature of Si/TPD heterointerface. The values of the ideality factor
(extracted from data between 0.2-0.3 V) further show that diode characteristics are
near-ideal, so the effect is not expected to be due to some unaccounted non-linearity
(Table 4.4).
A big concern is the large series resistance of the heterojunction devices, probably
due to low conductivity of the undoped TPD layer. In a practical solar cell, such large
ohmic losses are unacceptable so a intrinsically more conductive or doped organic
thin-film would be needed.
Assuming the simple model of Fig. 4.11 holds and the J0,oxide is equal among them
(all samples were annealed together so oxide was formed together), one can calculate
the saturation current density associated with the Si/p+ -Epi homojunction (J0,pEpi )
and Si/organic heterojunction (J0,heterojunction ). The J0 of the Si/p+ -Epi diode and
98
p Si Implant
Ag (50 nm) 3 × 1017 cm−3 (400 nm)
Dark Current −
p-Si
EC
n Si Epi n-Si Cathode
n+-Si
5 × 10 cm−3 (6 µm)
16
Anode
n+ Si Substrate EV
2 × 1019 cm−3
(500 µm)
Al (100 nm)
(a) (b)
Reduced Surface
p+ Si Epi Recombination
1019 cm−3 p+-Si
(200 nm) p Si Implant Epi
Ag (50 nm) 3 × 1017 cm−3 (400 nm)
Dark Current −
p-Si
EC
Cathode
n Si Epi n-Si n+-Si
5 × 1016 cm−3 (6 µm) Anode
n+ Si Substrate EV
2 × 1019 cm−3
(500 µm)
Al (100 nm)
(c) (d)
Reduced Surface
TPD/PQ LUMO Recombination
(15/10 nm) p Si Implant
TPD
Ag (50 nm) 3 × 10 17
cm−3 (400 nm)
Dark Current −
p-Si
EC
n Si Epi n-Si Cathode
n+-Si
5 × 1016 cm−3 (6 µm) Anode
n+ Si Substrate EV
HOMO
2 × 1019 cm−3
(500 µm)
Al (100 nm)
(e) (f)
Figure 4.13: Structure and band diagrams of the minority-carrier probe with different
p-type contacts (a) & (b) Si/metal contact (baseline) (c) & (d) Si/p+ -Epi/metal and
(e) & (f) Si/TPD/Metal.
99
Figure 4.14: Effect of PQ-passivated Si/TPD heterojunction on minority-carrier
probe. Measured I-V characteristics of the device with and without heterojunction.
For reference the I-V of the Epi structure is also shown. The device area is 200 µm
× 200 µm.
7600 32400
J0,‘Si/p+ −Epi0 = J0,oxide + J0,pEpi = 9.8 × 10−12 (4.6)
40000 40000
⇒ J0,pEpi = 1.0 × 10−11 A/cm2 (4.7)
Table 4.4: Current scaling, extracted ideality factor, and extracted J0 diode of the
heterojunction from data between 0.2-0.3V. For calculating J0 implanted area was
used as the device area: 200 µm × 200 µm.
100
Table 4.5: Calculated J0 of the four types of p-type contacts fabricated in this study
Si/PQ/TPD 2.6×10-11
and
7600 32400
J0,‘Si/organic0 = J0,oxide + J0,heterojunction = 2.3 × 10−11 (4.8)
40000 40000
⇒ J0,heterojunction = 2.6 × 10−11 A/cm2 (4.9)
Table. 4.5 summarizes the J0 values for all fours of the p-type silicon interface mea-
sured in this study.
The results do provide a first proof-of-concept that silicon/organic heterojunctions
can be used to selectively reduce the minority-carrier recombination, while allowing
transport of the majority carriers, at the p-type Si/metal contacts. More work in
needed to reduce the series resistance, optimize the band offsets, understand the
interaction of metal on the silicon and organic surfaces, and further reduce the surface
recombination at the Si/organic interface.
4.5 Conclusion
101
heterojunction. Finally, using the PQ-passivated Si/TPD heterojunction we showed
reduced minority-carrier recombination at the p-type metal-silicon contact.
Using such minority-carrier blocking heterojunctions in place of back-surface field
in conventional silicon solar cells is a low-cost pathway towards more efficient solar
cells. In an silicon-based solar cell with two silicon/organic heterojunctions, like the
one depicted in Fig. 2.1(b) or Fig. 1.5(c), such a heterojunction would be one of the
heterojunction. It would play a key role in reducing dark-current to achieve a large
open-circuit voltage.
102
Chapter 5
Silicon/Organic Heterojunction to
Block Majority Carriers
5.1 Introduction
Conventional silicon solar cells use diffused p/n junction to block majority carriers and
separate photogenerated carriers, yielding a photocurrent (Fig. 5.1(a)). A potentially
cheaper but just as effective alternative to diffused junctions could be a majority-
carrier blocking silicon/organic heterojunction like that one introduced in Chapter 2
(Fig. 5.1(c)). The heterojunction is different, and technologically more useful, than
the silicon/organic heterojunction described in Chapter 4 because it functions as an
alternative to the p-n junction, a essential component of every solar cell.
As we shall demonstrate, silicon surface passivation is less important in majority
carrier blocking heterojunctions because of a large electric field at the surface of
silicon that reduces carrier recombination. The critical issue in this type of structure
is band-alignment at the Si/organic interface. The hetero-interface even in this case
is an electron-blocker so the specific band-alignment criteria remain the same as in
Chapter 4:
103
(a) The LUMO of the organic should be much higher than the conduction band
edge of silicon, i.e. there should be a large conduction band offset, so that
electrons are repelled away from the surface,
(b) The HOMO and the valence band edge of Si should be aligned, i.e. there should
almost no valence band offset, so that photogenerated holes can be efficiently
extracted at the anode.
First principle 1D simulations were used to confirm that p-n junction solar cells can be
substituted by an electron-blocking silicon/organic heterojunction. As before, all sim-
ulations were done using the Taurus Device tool from Synopsys® . The simulated
structure is shown in Fig. 5.2(a). The recombination lifetime in silicon was set at 100
µs. The organic had a conduction-band and valence-band offset of 0.5 eV and 0 eV,
respectively (Fig. 5.2(b)). The bulk mobility of the organic layer was 10-3 cm2 /V·s.
For simplicity, the organic semiconductor was modeled as a conventional semicon-
ductor with density of states that were 10 times lower than the density of states of
silicon. The reduced density of states were supposed to simulate the fact that at
room-temperature only those states are occupied that are within around 50 meV of
104
Dark Current Dark Current
−
Photocurrent
− Photocurrent
EC
p-Si EC Cathode
Anode
Cathode
Anode n-Si
n-Si
Photocurrent + Photocurrent
EV + EV
Organic
Dark Current EC /LUMO Offset
Passsivation − Photocurrent EC
Cathode
Anode n-Si
EV /HOMO Offset
Photocurrent EV
HOMO +
Dark Current
(c)
Figure 5.1: Band diagram of a (a) p-n junction and (b) Si/metal Schottky solar cell.
The solid lines are desired direction of generated carrier flow and the dashed line
represents dark current, or equivalently the loss mechanism. (c) An electron-blocking
silicon/organic heterojunction with lower J0 than Schottky junction.
the band-edge. While not exact, this should qualitatively reproduce the expected
behavior. Since the surface recombination velocity (SRV) at the silicon/organic het-
erojunctions is important, simulations were performed with a SRV of 0, 500, and 5000
cm/s.
Under the assumption that surface defects do not cause any Fermi-level pinning
in Si/metal or Si/organic heterojunction, simulated band-diagrams show a depletion
region in silicon in both Si/metal and Si/organic/metal junctions (Fig. 5.2(a)). The
strength of the built-in field is also the same in both cases, decided only by the differ-
ence in work-function between the top metal (5.1 eV) and n-type silicon. However due
to the conduction-band offset, the simulated forward-bias currents in the passivated
Si/organic device are much lower (Fig. 5.2(c)). Under a simulated illumination of
105
Anode Organic (10 nm)
n Si Substrate
5 × 1014 cm−3
(500 µm)
Cathode
(a) (b)
(c) (d)
Figure 5.2: (a) The structure and (b) The band diagram of a metal-organic-silicon
diode used in simulations. The inset shows the heterojunction at the top of the struc-
ture in more detail. (c) The simulated current-voltage characteristics in dark and (d)
under a illumination of ≈ 1.1 Sun illumination, for a device with and without het-
erojunction. Non-ideal heterojunctions with non-zero surface recombination velocity
are also shown. Performance of the cell has a relatively weak dependence on surface
recombination velocity.
approximately 1.1 Sun (photon-flux of 3×1017 cm-2 ), the depletion region in the two
devices cause the same photocurrent in both the devices, but the lower dark-current
allows a higher open-circuit voltage in heterojunction device. From the results it is
clear that a) heterojunctions can yield diodes with lower J0 than metal/Si diodes and
b) that the built-in electric field in a heterojunction can efficiently separate photo-
106
(a) (b)
Figure 5.3: (a) The simulated electron-density profile at the silicon/organic interface
at 0.5V forward bias and 1.1 Sun illumination. (b) The simulated conduction band
profile at the silicon/organic interface at 0.5 V forward bias under 1.1 Sun. Notice
there is an electric field in silicon that responsible for the keeps electrons away from
the interface.
107
and ps in cm-3 ).
ns ps − n2i
Usurf ace = s
ns + ps + 2ni
≈ sns (5.1)
The approximation is valid because under illumination ps , ns >> n2i , and at the edge
of the depletion region, ps >> ns . In the simulated diode, at 0.5 V forward bias
and 1.1 Sun illumination, the electron and hole densities at the silicon surface are
only ≈1012 cm-3 and ∼ 1016 cm-3 , respectively (Fig. 5.3(a)). The current lost due to
surface recombination is directly related to the rate of surface recombination (U ).
Jloss,SRV = qU
= qsns (5.2)
So even for a large SRV of 5000 cm/s, the current lost due to surface recombination
is only ≈0.5 mA, i.e a loss of only around 1% of the total photocurrent. The reason
for such a low surface carrier density is the depletion region that exists at the silicon
surface due to work function difference between anode and n-type silicon (Fig. 5.2(b)).
As the voltage increases from 0.5 V to 1 V the surface field starts to reduce till it
turns negative (Fig. 5.2(b)), and effect of surface recombination starts to affect the
currents (Fig. 5.2(d)). However, near the peak-power point of the solar cell (around
0.5 V), the recombination losses are insignificant.
108
5.3 Silicon/Pentacene Heterojunction
5.3.1 Pentacene
Pentacene is a well known hole-conducting organic small molecule, with HOMO and
LUMO edges at 5.1 eV and 3.2 eV, respectively [106]. The silicon band edges are at
5.17 eV and 4.05 eV for EV and EC , respectively, and so the silicon/pentacene interface
is expected to have a large EC /LUMO barrier (0.8 eV) but a negligible EV / HOMO
barrier. Another motivation for choosing pentacene to form the heterojunction was its
high field-effect hole mobility, which in the best crystalline films is >1 cm2 /V·s [107].
The bulk-mobility of pentacene in amorphous thin-films is, however, considerably
lower and bulk resistivity of 1014 Wcm have been reported [108]. The high resistivity
of amorphous pentacene films, as we shall in the last paragraph of this section, is an
issue with these devices.
Sublimed pentacene was bought from H. W. Sands and was deposited on silicon by
thermal evaporation using the Angstrom Evaporator in C405A. Prior to pentacene
deposition, silicon wafers were cleaned and PQ-passivated by the method described
in Section 3.6.
Test structures with a Si/PQ/pentacene heterojunction were fabricated on crys-
talline CZ n- and p-type Si substrates (Fig. 5.4(a) & (c)). The function of the PQ
layer was to passivate the silicon surface. The pentacene layer was supposed to set
the effective band-offsets of the Si/organic heterojunction because the PQ layer is
very thin. The top electrode was patterned by a shadow mask with an active area of
3.1×10-2 cm2 (1-mm radius circles). Adjacent devices were not explicitly separated.
Rather it was assumed that the high lateral-resistance of the organic layer would be
enough to prevent any current spreading, and the device area would be defined by
109
LUMO
Pd (15 nm)
0.8 eV
PQ (10 nm) Pentacene (20 nm) Pentacene Dark Current
PQ − Photocurrent EC
n-Si Substrate
5 × 1014 cm−3 Top Bottom
n-Si
(500 µm)
Photocurrent
EV
HOMO +
Ag (50 nm) Dark Current
(a) (b)
LUMO
Ag (50 nm)
(c) (d)
Figure 5.4: (a) Structure and (b) expected band structure of a n-Si/PQ/pentacene
test device at a small positive voltage on top electrode. (c) Structure and (d) expected
band structure of a p-Si/PQ/pentacene test device at a small negative voltage on top
the electrode. It is assumed that carriers tunnel through PQ.
the top electrode. The top electrode was designed to be semi-transparent, by making
the metal (Pd) very thin (15 nm), allowing the transmission of some of the incident
light to Si. All electrical measurements were done using the Agilent 4155 parameter
analyzer.
All spectroscopic measurements were performed in Prof. Kahn’s lab by Dr. Yabing
Qi.
110
5.3.3 Current-Voltage Characteristics of Silicon/PQ/Pentacene
Heterojunctions
We initially assumed that the passivating PQ layer is thin enough that carriers can
tunnel through, so the band structure of PQ is not considered in the band digram of
Fig. 5.1(c). The n-Si/pentacene device is expected to have a large electron barrier at
the Si/organic heterointerface (Fig. 5.4(b)) and should function like a heterojunction
diode. For the case of p-Si, there is no equivalent barrier for holes to from silicon to
the top contact and the structure should simply function as a resistor (5.4(d)).
For n-Si/pentacene/metal heterojunction, experimentally measured dark I-V char-
acteristics are diode-like. The J0 of the pentacene device is lower than the J0 of an
equivalent Si/metal device (Fig. 5.5(a)). This suggests that the electron barrier-
height at the semiconductor/metal interface is larger in the pentacene device than in
the Shcottky barrier device, arguably due to the larger conduction band offset. Un-
der illumination, the field in silicon helps separate the photogenerated carriers, and
an open-circuit voltage of 0.37 V is observed, which is higher than the open-circuit
voltage of the Si/metal device (Fig. 5.5(a)).
The thin Pd is semi-transparent. With no heterojunction, i.e. for the n-Si/metal
Schottky diode, the short-circuit current is approximately 2.1 mA/cm2 . The mea-
sured value of short-circuit current in n-Si/PQ/pentacene device is not only far lower,
only 0.45 mA/cm2 , but also increases with negative bias, giving an ‘S’ shaped charac-
teristics (Fig. 5.5(b)). The voltage dependence suggests that the lower photocurrent
at zero-bias is not due to poor light absorption, since higher internal electric field
is not expected to improve absorption. Intuitively too, poor absorption should not
be the problem because most of the light absorption is expected to occur in silicon,
which is known to have a high light absorption efficiency. More likely, the problem is
in carrier collection - photogenerated holes are not being efficiently collected at the
top electrode.
111
(a) (b)
(c) (d)
One hypothesis for the poor collection could be that valence band offset at the
Si/organic interface is not zero but significant. At positive bias the barrier blocks the
flow of photogenerated holes from silicon to anode, leading to a lower photocurrent
(Fig. 5.6(a)). As the bias is made more negative, the field at the surface increases
and more of the carriers are able to cross the valence-band barrier to get collected at
the anode, leading to a higher photocurrent (Fig. 5.6(b)).
112
LUMO
LUMO
Pentacene Pentacene
PQ PQ
− Photocurrent −
EC Photocurrent
Bottom EC
Top Bottom
Top n-Si
n-Si
EV HOMO
HOMO + + EV
Low Photocurrent Higher Photocurrent
(a) (b)
Figure 5.6: Effect of surface field on photocurrent when the n-Si/PQ/pentacene diode
has a valence band offset. Band structure at (a) positive-bias and (b) negative-bias.
Photocurrent is higher in negative-bias.
The dark characteristics for p-Si/PQ/pentacene heterojunction are not ohmic but
diode-like and under illumination a small open-circuit voltage is observed (0.07 V),
hinting at the presence of a small internal electric field in silicon (Fig. 5.5(c)). These
observations too can be explained by considering a small valence-band offset at the
Si/organic interface (Fig. 5.5(c)). The photocurrent on the p-Si diode also shows a
bias voltage dependence (‘S’ shaped characteristics of Fig. 5.5(d)), suggesting poor
collection of electrons at the top electrode. However this is expected because we
expect a LUMO/conduction-band barrier.
In summary, I-V characteristics of the PQ-passivated Si/pentacene heterojunc-
tions are not consistent with the band structure of Fig. 5.4(b) & (d). There is
evidence that not just the electrons, but holes also see a barrier at the Si/organic
interface. The hole-barrier could be present either because of HOMO of pentacene
does not align with the conduction band edge of silicon or because PQ is not thin
enough and the carriers are being blocked at the Si/PQ interface.
113
5.3.4 Spectroscopy of the Silicon/PQ/Pentacene Heterojunc-
tion
To investigate the reasons for the anomalous I-V characteristics, the band-alignment
of the Si/PQ/pentacene heterojunction was studied using X-ray and ultra-violet pho-
toelectron spectroscopy (XPS & UPS).
For the experiment PQ-passivated n+ and p+ -Si samples loaded into the UHV
chamber and the band-alignment and Si band bending were measured using the same
process described in Section 3.6.3. The band-bending in PQ-passivated silicon was
higher than measured before (Table 3.2), 0.4 eV and 0.3 eV for n-Si and p-Si, respec-
tively, suggesting imperfect passivation but the band-offsets at the Si/PQ interface
were consistent with previous experiments (Fig. 3.11).
Next, pentacene was deposited (in-situ) at a very slow rate on top of the Si/PQ
interface. At regular intervals pentacene deposition was paused and UPS spectrum
was measured again. Plotting the work-function (Φ) and ionization energy (IE) as a
function pentacene thickness (Fig. 5.7(a) & (c)), allows us to observe the evolution
of the Si/PQ/pentacene interface for the first few nm of pentacene. After only 2-3
nm the energy levels of pentacene stabilize, suggesting that most of the interface
effects have been accounted for and bulk properties of pentacene dictate properties
for thicker films.
The Si/organic valence-band offset (∆EV,org ) in any system is the difference be-
tween the ionization energies of the Si/organic interface (IEorg ) and silicon surface(IESi ),
adjusted for the ‘net’ interface dipole (∆IDSi/org ).
Assuming IESi , i.e. energy of the valence-band edge, for pristine silicon is 5.2 eV
[95], the band offset for Si/pentacene interface (EV,pent ) can be calculated. For n+ -
114
0.3 eV
EV ac
0.2 eV
LUMO
IE Φ 1.4 eV
2.2 eV 3.2 eV
EF CB
0.4 eV
HOMO
0.8 eV 0.7 eV VB
Pentacene PQ n+ Si
(a) (b)
0.2 eV 0.8 eV
EV ac
0.3 eV
LUMO
1.1 eV
IE Φ 2.2 eV CB
3.2 eV
EF 0.3 eV
VB
HOMO
0.7 eV 1.0 eV
Pentacene PQ p+ Si
(c) (d)
Si and p+ -Si the measured valence band-offset for pentacene (∆EV,pent ) is -0.1 and
0.2 eV, respectively (Table 5.1). While not perfectly zero, these offsets are in the
ballpark of our original assumption that ∆EV,pent ≈0, suggesting that pentacene is
not responsible for the valence band offset. The only other source of a valence band
offset is the passivating PQ-layer. Similar calculation done on the data presented in
Fig. 3.11 (Section 3.6.3) for the valence band-offset at Si/PQ interface reveals a large
115
Table 5.1: Measured ionization energy (IE), interface dipoles, and valence-band offset
(∆EV ) at Si/pentacene and Si/PQ interface for n+ -Si and p+ -Si. IESi is assumed to
be 5.17 eV.
n+ -Si p+ -Si
Interface
IEorg ∆IDSi/org ∆EV,org IEorg ∆IDSi/org ∆EV,org
(eV) (eV) (eV) (eV) (eV) (eV)
Si/pentacene 4.8 0.3 -0.1 4.8 0.6 0.2
Si/PQ 5.6 0.3 0.7 5.4 0.8 1.0
LUMO
LUMO
1.1 eV
1.4 eV − Current
Dark Current Pentacene EC
Pentacene PQ
PQ p-Si
− Photocurrent EC Top Bottom
Bottom
Top Current + EV
n-Si 1.0 eV
0.7 eV
EV
Photocurrent + HOMO
barrier (0.7 to 1.0 eV) at both n+ and p+ -Si surfaces (Table 5.1). So if PQ is not thin
enough for carrier tunneling, holes in silicon could experience a barrier and the I-V
characteristics would be limited by the properties of the Si/PQ interface and not the
Si/pentacene interface.
Data from spectroscopic measurements shows that if there is a valence-band off-
set, then the most likely culprit is not the pentacene layer but PQ. Most likely, the
carriers are unable to tunnel through the PQ layer. Assuming that the trend of the
spectroscopic data is the same for lightly and highly doped Si, the revised band struc-
ture of the Si/PQ/pentacene heterojunction, which includes the PQ band structure,
116
is shown in Fig. 5.8.
Cell
If the large valence-band offset at the Si/PQ layer is the reason for poor hole collection
in Si/PQ/pentacene devices, then thinner PQ layers should help. Once the thickness
of the PQ layer is reduced sufficiently, carriers will be able to tunnel through the
barriers at the Si/PQ interface. Devices with thinner PQ layers, 2 nm as opposed to
10 nm, were fabricated to test the hypothesis. To reduce the large series resistances
observed in previous sets of devices, the pentacene thickness was also reduced to from
20 nm to 5 nm.
The measured J-V characteristics of the p-Si device are ohmic (Fig. 5.9(a)) with a
specific contact resistance (Rvert ) of ∼ 100 kWcm2 , confirming that the hole barrier is
negligible at all the Si/organic interfaces. Assuming that most of the resistance is due
to the 7 nm (5 nm of pentacene and 2 nm of PQ) thick organic layers, the resistivity
(ρvert ) of the organic layer stack is calculated to be ∼ 108 Wcm. Dark characteristics
of the n-type device are still diode-like, and under illumination an open-circuit voltage
of 0.23 V is observed (Fig. 5.9(b)). The short-circuit current is ≈ 2.2 mA/cm2 - in line
with what we expected from 0.1 Sun illumination through the semi-transparent metal
(Fig. 5.9(c)). There is no ‘S’ shape curve, and the photocurrent is independent of
bias-voltage. The Rvert and ρvert of this device are consistent with the values measured
for the p-type Si device. Overall, the Si/(thin)PQ/pentacene interface functions as
an electron-blocking heterojunction.
Unfortunately, the photovoltaic performance of the device is not very good. While
the short-circuit current is high, the obtained open-circuit voltage is barely higher
than the Si/metal control diode (0.23 V instead of 0.2 V), and is too low for a
practical solar cell.
117
(a)
(b) (c)
Figure 5.9: (b) J-V characteristics of the p-Si/PQ/pentacene device with thinner or-
ganic layers (2 nm of PQ and 5 nm of pentancene) in dark and ≈ 0.1 Sun illumination
plotted on a linear scale. (c) J-V characteristics of the n-Si/PQ/pentacene device so-
lar cell with thinner organic layers (2 nm of PQ and 5 nm of pentacene) in dark and
≈ 0.1 Sun illumination plotted on a log and (b) linear scale.
One reason for the low VOC could be that we reduced the thicknesses of the
pentacene layer too much, and somewhere between 5 nm and 15 nm is a pentacene
thickness that will yield a higher VOC . Unfortunately, such an optimization is not
useful simply due to the high series resistance of the device. For just for 5 nm thick
pentacene layer the measured specific contact resistance is around 100 W.cm2 (ρvert ∼
100 MWcm). The vlaue of resistivity is lower than some previously reported values,
e.g. 1014 Wcm [108], but it is still a very high number. The expected I 2 R loss at 1 sun
118
illumination (current density of ∼ 40 mA/cm2 , assuming an ideal photocurrent for a
silicon solar cell) is 160 W/cm2 . Since the incident light at 1 sun is only 100 mW/cm2 ,
power dissipation due to the resistance will overwhelm any power generated. A more
conductive organic is required to make a practically useful heterojunction.
The implied bulk mobility of holes (µh ) in the organic stack can be calculated
from the measured value of ρvert . Since pentacene is a hole conductor the resistivity
(ρvert ) can be written as
1
ρvert = (5.4)
qpµh
Ideally at 1 Sun, the photon-flux in a solar cell is on the order of 1017 cm-2 . Assuming
a silicon wafer thickness of 500 µm, the carrier density in the device is on the order
of 1018 cm-3 . The carrier density in the organic layer should be of the same order. So
the implied bulk-mobility of the organic layer is
1
µh =
qpρvert
⇒ µh ∼ 10−7 cm2 /V s
In comparison, the required value of µh for the organic, such that voltage drop
across the organic layer (∆Vorg ) is less than 10 mV at the current density (J) expected
at 1 Sun (40 mA/cm2 ), is
∆Vorg
Rvert <
J
⇒ Rvert < 0.25 Ωcm2
119
5.4 Silicon/Poly(3-hexylthiophene) Heterojunction
5.4.1 Poly(3-hexylthiophene)
P3HT deposition were done using the recipes provided by the Prof. Lynn Loo’s group
in the Department of Chemical Engineering. First, P3HT inks were formulated by
dissolving the requisite amount P3HT in chlorobenzene and agitating the solution by
a magnetic stirrer on a hot plate set at 90, till the particles of P3HT dissolved away
and the solution turned clear. Typically this took only around a minute. Though the
solution was formulated in air, the P3HT stock and inks were stored in a glove box
under inert atmosphere. To prevent photoinduced degradation, solutions were stored
in dark-tinted vials. P3HT was bought from Merck KGaA under the brand name
120
Table 5.2: The spin-coating recipes of P3HT used in this work. Thicknesses were
measured by the Tencor profilometer.
SP001 LisiconTM (Batch number EE98202) with a stated average molecular weight
(MW ) of 46115 gm/mol. Chlorobenzene was bought from JT Baker (CMOS Grade).
Prior to depositing P3HT, silicon (100) substrates were first rinsed in solvents to
remove organic impurities - acetone, methanol and 2-propanol in an ultrasonic bath
for 5 minutes each. Next, trace metal contaminants were removed by the standard
RCA clean - first SC-1 step: Ammonium hydroxide + hydrogen peroxide + DI Water
(1:1:5) at 80◦ for 15 minutes, followed by 1:100 hydrofluoric acid for 1 min, and
finally SC-2 step: hydrogen chloride + hydrogen peroxide + DI Water (1:1:5) at 80◦
for 15 minutes [78]. Lastly, the wafers were dipped in a 1:100 aqueous hydrofluoric
acid solution for 1 min to strip the native oxide layer. On the resulting hydrogen-
passivated Si wafers, P3HT was spin-coated from a pre-formulated P3HT solution.
Table 5.2 shows the various recipes used in this work and the resulting P3HT layer
thickness. Unless otherwise specified, no thermal processes were used to dry the
P3HT layer.
The depletion region in silicon is set by the difference in the work function of the
top electrode and silicon. For high VOC the depletion region should be as large as
possible. Since the substrate were n-Si, high-work function metals were preferred for
top electrode. Best results were obtained with palladium, though devices with gold
were also fabricated. The bottom electrode was a simple ohmic contact to the silicon.
121
Dark Current
Pd 15 nm
− Photocurrent
EC
n-Si Substrate Cathode
Anode
5 × 1014 cm−3
(500 µm) n-Si
Photocurrent
+ EV
Ag 50 nm Dark Current
(a) (b)
LUMO
Figure 5.10: (a) Structure and (b) expected band structure of a Si/metal Schottky
test device which serves as the control. (c) Structure and (d) expected band struc-
ture [9] of the heterojunction device test device used to evaluate effect of Si/P3HT
heterojunction.
Due to the large area surface and etched surface, ohmic Si/metal contacts were formed
quite easily on the backside. Metals used were either aluminum or silver. Electrodes
were deposited by thermal evaporation in the Angstrom Evaporator in C405A. The
top electrode was patterned by a shadow mask with an active area of 3.14x10-2 cm2
(1-mm radius circles). The top electrode was also kept very thin (15 nm) so that it
is semi-transparent, allowing some of the light to be absorbed in Si. All electrical
measurements were done using the Agilent 4155 parameter analyzer, HP 4175 LCR
meter, and Techtronix oscilloscope. For AM 1.5 measurements, a xenon-lamp solar
simulator was used.
To investigate Si/P3HT heterojunctions, tests structures with and without P3HT
were fabricated (Fig. 5.10) on crystalline phosphorous-doped silicon wafers.
122
5.4.3 Band Modulation at the Unpassivated Si/P3HT Het-
erojunction
Unlike the Si/PQ/pentacene heterojunction diodes, the silicon surface at the Si/P3HT
heterojunction in the test devices is unpassivated (Fig. 5.10(c)). There are thee major
reasons for this design decision:
(a) PQ causes series resistance and valence-band offset in devices, both of which
degrade photovoltaic performance (as demonstrated in the previous section.).
(b) Surface passivation is not very crucial in this type of heterojunction device and
hydrogen passivation should be sufficient, as demonstrated by simulations in
Section 5.2.
(c) PQ passivation did not seem stable under solvent processing (Section 6.3.3). So
there was concern that PQ would get washed away while spin-coating P3HT on
silicon.
1 2
2
= (Vbi − VSi ) (5.6)
CSi qSi NSi
123
where, q is the electronic charge, Si is the permittivity of Si, NSi is the Si substrate
doping, and Vbi is the built-in voltage. In a Schottky junction such a depletion region
exists at the silicon surface, as confirmed by the linear dependence of 1/C 2 on VSi for
Si/metal diodes (‘No P3HT’ curve of Fig. 5.11(b)).
For the Si/P3HT heterojunction devices, the measured capacitance (Cmeasured )
and applied voltage (Vapplied ) include contribution from not just the depletion region
(CSi ) but also the organic layer (Corg )
1 1 1
= + (5.7)
Cmeasured CSi COrg
where VOrg is the voltage drop across organic layer. The capacitance due to the thin
organic layer can be approximated by
Org
Corg = (5.9)
torg
To extract the values of CSi and VSi in the heterojunction devices from the measured
values, following equations can be used:
1 1 1
= − (5.10)
CSi Cmeasured COrg
qSi NSi
VSi = Vapplied − (5.11)
COrg CSi
2
Irrespective of P3HT thickness, the resulting 1/CSi − VSi curves are linear with
the same slope (Fig. 5.11(c)). From the slope, the value of implied Si substrate
doping as function of distance can be extracted (Eq. (5.6)), which suggests that Si
is uniformly doped with doping density of ∼ 4×1014 cm-3 - consistent with expected
124
(a) (b)
(c) (d)
Figure 5.11: (a) Small signal capacitance of n-Si/P3HT heterojunctions with different
P3HT layer thicknesses. (b) 1/C2 -V characteristics the same n-Si/P3HT heterojunc-
tions. (c) 1/CSi 2 -VSi characteristics extracted from the raw data using Eq. (5.10) &
(5.11). (d) Implied silicon substrate doping for different Si/P3HT devices.
n-type doping density of the 10-20 W.cm wafer used in these experiments. The built-
in voltage inferred from the x-axis intercept varies from ∼0.6 V for metal/Si devices
to ∼ 0.9 V for thicker P3HT layers. Quantitatively, the difference in Vbi may not be
significant, but qualitatively it proves the existence of a depletion region in silicon
that can separate photogenerated carriers.
125
(a) (b)
126
it is conceivable that the current reduction observed in devices with thicker P3HT
layers is purely due to the resistive drop across the P3HT. However in that case the
current reduction would scale inversely and continuously with P3HT layer thickness -
something not seen in the data of Fig. 5.12b. The observed initial current reduction is
rapid and after a certain threshold (P3HT thickness of 10 nm) independent of P3HT
layer thickness. This allows us to further that the reduction in electron current is
strictly an interface effect due to the electron barrier at the Si/P3HT heterojunction.
To test the transport of holes, current-voltage characteristics of the same devices
were measured under illumination from a microscope lamp Fig. 5.12(b). The intensity
of the lamp caused a short-circuit current of ∼ 4 mA/cm2 in a commercial silicon diode
with anti-reflection coating and no blocking metal, roughly corresponding to 1/10th
the intensity of sunlight. Since the top metal in these devices is semi-transparent,
some of the carriers are expected to reach the underlying silicon layer, generating
electrons and holes. At zero bias the photogenerated holes in silicon would flow from
silicon to anode, across the Si/P3HT interface, giving rise to a photocurrent. As
discussed in Section. 2.5 and Section 5.3.3, any hole-barrier at the hetero-interface,
e.g. due to band-offset between the valence band of Si and the HOMO of P3HT,
would impede the flow of photogenerated holes leading to a reduced short-circuit
current. Experimental data suggests that this is not the case because in devices,
both with and without P3HT, the values of short-circuit current were similar (∼ 2
mA/cm2 ) and not dependent on the applied voltage (in contrast Si/PQ/Pentacene
devices shown in Fig. 5.5(b)). This points to a lack of a hole barrier at the Si/P3HT
interface, so the HOMO of the P3HT must be closely aligned to the valence band
edge of silicon.
The lower J0 enables higher VOC according to Eq. (1.4). This is reflected in the
data, where under around 0.1 sun illumination, the VOC increases from 0.20 V for
Si/metal device to 0.41 V for the best Si/P3HT device.
127
Table 5.3: The extracted device parameters for Si/P3HT device parameters of
Fig. 5.12. Implied J0 calculated from Eq. (1.4), using measured VOC and JSC and
Eq.(1.4).
Device Parameters
P3HT thickness
Extracted J0 n JSC VOC Implied J0
(nm) (A/cm2 ) (mA/cm2 ) (V) (A/cm2 )
No P3HT 9.2×10-7 1.01 2.06 0.20 9.0×10-7
2 nm 3.5×10-8 1.06 1.80 0.30 1.6×10-8
10 nm 7.0×10-9 1.21 2.08 0.41 2.6×10-10
25 nm 7.7×10-9 1.18 1.93 0.41 2.4×10-10
As mentioned before, the J0 (as measured from the intercept on the y-axis) of the
best Si/P3HT heterojunction device is lower than the J0 of the Si/metal Schottky
device, 7.0×10-9 A/cm2 instead of 9.2×10-7 A/cm2 . However, other than the Si/metal
Schottky device, whose n=1.01, the ideality factors of heterojunction devices are all
greater than 1 - from 1.06 to 1.25. These high ideality factors affect the extraction of
J0 from the semilog characteristics, which conflates the comparisons between device.
Fortunately, the ideality factors are not our primary concern, VOC and JSC are. From
the measured VOC , JSC , and assuming n = 1 an “implied J0 ” can be calculated from
Eq. (1.4), i.e. the J0 that an ideal diode should have to justify the measured VOC at
the measured JSC (Table 5.3). To compare the performance of devices, “implied J0 ”
is a better metric than the extracted J0 . The implied J0 of heterojunction devices is
more than 3 orders of magnitude lower than the Si/metal device (Table 5.3).
Overall the results of Si/P3HT heterojunctions are consistent with the proposed
theory and simulations of Section 2.5. At the Si/P3HT heterojunction, there exists
an electric field in silicon that can separate photogenerated charge carriers, yielding
a photocurrent. The n-type-Si/P3HT heterojunction blocks electrons due to the
electron barrier, resulting in a lower saturation current density, but allows unimpeded
transport of holes.
128
PQ (10 nm) P3HT (10 nm)
Pd 15 nm
n-Si Substrate
5 × 1014 cm−3
(500 µm)
Ag 50 nm
(a) (b)
129
improve the performance of Si/P3HT heterojunction, in fact it degrades it.
First principle simulations were performed to separately plot the electron and hole
currents in the heterojunction device of Fig. 5.2(a). The simulation assumed a con-
duction band offset of 0.5 eV at the Si/organic interface. The recombination lifetime
of the silicon was set at 100 µs and the doping of the silicon wafer was 5×1014 cm-3 .
Without the heterojunction (with Schottky barrier), the I-V characteristics show a
dominant electron current and a smaller hole current (Fig. 5.14(a)). When the hetero-
junction is introduced, the hole current remains essentially unchanged but the electron
current is reduced by several orders of magnitude (Fig. 5.14(b)). The reduction in
electron current is so dramatic in the heterojunction device, that the hole-current
becomes the limiting factor for further reduction in J0 . This is significant because it
130
tells us that further reduction in J0 will not be achieved by making the Si/organic
barrier larger (say by increasing it to 1.5 from the present 1.1 eV). Rather it must be
achieved by reducing minority-carrier (hole) injection from anode into silicon.
(a) (b)
Figure 5.14: Simulated values of current in a (a) Si/metal and (b) Si/organic device, tp
highlight the relative contribution of electron and hole currents to total dark-current.
Electron current dominates the former and hole-currents dominate the latter. The
conduction-band offset at Si/organic interface was set at 0.5 eV. The recombination
lifetime of the silicon was set at 100 µs and the doping of the silicon wafer was 5×1014
cm-3 .
J0 ≈ Jelectron J0 ≈ Jhole
131
VS,F
I Vout VS (V)
VS,R time
VS Test Diode
IF
Function
10 kΩ
Generator ID (µA)
IR time
22 µA
ts
(a) (b)
Figure 5.16: (a) The experimental set up to measurement storage times in hetero-
junction test diodes, in order to detect stored minority-carriers. (b) Applied and
measured waveforms for the experiment.
(Section 2.5) and J0 for Si/P3HT device were ∼ 10-33 A/cm2 , much lower than the
expected hole current.
Direct evidence of minority-carrier dominance can be obtained by measuring the
minority-carrier storage times and observing the effect of doping and recombination
lifetime. In the next few sections, these measurements will be used to experimentally
demonstrate that in Si/P3HT heterojunction devices the electron current has been
lowered to a point where the hole current starts to dominate J0 .
In a typical silicon p-n diode, the forward-bias current causes minority-carriers to get
stored into the quasi-neutral regions. The presence of these stored minority-carriers
can be probed by measuring the reverse-bias storage times [112, 113].
When a minority-carrier diode is suddenly switched from forward to reverse bias,
the stored charges (and hence the voltage across the p-n junction) cannot switch
instantaneously. Till the junction maintains its forward-bias, the junction remains in
the ‘On’ state, and a simple voltage loop equation tells us that a negative current will
flow in the circuit (Fig. 5.16(b)). Once the stored charges density has fallen enough to
turn off the junction, the current reverts back to its equilibrium value (Fig. 5.16(b)).
The circuit used to measure this transient delay, also called ‘storage’ time (ts ),
132
along with the voltage waveforms is shown in the Fig. 5.16(a) and (b). The circuit
consists of a pulse-waveform generator. The output of the generator (VS ) is such that
it first forward-biases the diode at VS,F and then quickly switches it to reverse-bias
voltage VS,R . The transient current though the diodes IF in forward-bias and IR in
reverse-bias) is estimated by a storage oscilloscope by measuring the voltage across a
resistor connected in series with the diode. The value of the reverse current IR can
be quite high and is set by the value of the resistor R (in this case R=10 kW) - higher
the value of R, lower the transient reverse-current (IR ). For a hole-current limited
diode, ts is given by the equation [112]:
r −1
ts IR
erf c = 1+ (5.12)
τr IF,hole
where τr is the hole recombination lifetime, IF,hole is the hole component of the total
forward-bias current (IF ) and IR is the total reverse-bias current (IR ) during the
storage transient. Also
IF = IF,electron + IF,hole
At fixed value of total forward-bias current (IF ), reverse-bias current (IR ) and
recombination lifetime (τr ), devices with higher hole-currents (IF,hole ) would yield
longer storage times. For this experiment, all the diodes were fabricated on the same
Si wafer, so the recombination lifetime can be assumed to be the same. The total
reverse-bias current was fixed at 0.22 µA by keeping resistance (R) at 10 kWand tuning
the VS,R , and storage times were measured at two different values of total forward-bias
current - 100 µA and 1 mA.
In the diode without P3HT, the contribution of minority-carriers towards the total
current is very small, so the measured storage time at 100 µA forward-current was
very low (ts <0.1 µs) (Fig 5.17(a)). For diodes with thick P3HT layers (>10 nm)
133
the storage times increased to ∼ 2.6 µs, indicating that hole current is a much larger
fraction of total current. Similar results were also obtained at a forward-bias current
of 1 mA for the same set of devices (Fig. 5.17(e)).
While the storage time results do not conclusively prove that hole-currents exceed
electron-currents in heterojunction devices, they do demonstrate that compared to
Si/metal Schottky devices, minority-carrier (hole) currents are a bigger component
of total current in Si/P3HT heterojunction devices.
r
n2 Dhole qV /kT
Jhole (V ) = q i
e −1 (5.13)
NSi τr
where, q is the electronic charge, ni is the intrinsic carrier density in silicon, Dhole is
the hole diffusion coefficient, and V is the applied voltage. We assume that the diode
is in the long-base regime, i.e. Lhole << Si wafer thickness. By increasing the n-type
doping level (NSi ), and/or increasing the minority-carrier lifetime in the silicon (τr ),
the hole-injection current can be reduced.
Simulation show the beneficial effect of increased doping. A structure similar to
the one shown in Fig. 5.2(a), but with a higher doping (ND =1016 cm-3 ), was simulated
in dark and under ∼1.1 sun illumination. Compared to the lower doped (ND =5×1014
cm-3 ) Si/organic devices, higher-doped devices have a lower J0 and higher VOC .
Typically, silicon wafers grown by the Czochralski (CZ) process have lower minority-
carrier recombination lifetime than those grown by Float-Zone (FZ) method : <1 ms
134
(a) (b)
(c) (d)
(e)
Figure 5.17: Output waveforms measured at the oscilloscope for test devices with (a)
no P3HT, (b) 2 nm P3HT layer, (c) 5 nm P3HT layer, and (d) 25 nm P3HT layer.
Devices were forward-biased at 100 µA current. (e) Measured storage time (ts ) at
forward-bias current of 100 µA and 1 mA.
135
(a) (b)
Table 5.4: The extracted device parameters for Si/P3HT device made on higher-
doped FZ wafers. Implied J0 calculated from Eq. (1.4), using measured VOC , n and
JSC .
Device Parameters
Device
Extracted J0 n JSC VOC Implied J0
(nm) (A/cm2 ) (mA/cm2 ) (V) (V)
No P3HT (5×1014 cm-3 , CZ) 9.2×10-7 1.01 2.1 0.20 9.0×10-7
25 nm (5×1014 cm-3 , CZ) 1.6×10-8 1.14 1.6 0.41 2.0×10-10
25 nm (7×1015 cm-3 , FZ) 2.9×10-9 1.25 1.6 0.44 6.3×10-11
for CZ compared [26] to ∼ 10 ms for FZ [27]. The wafers used in the experiments of
Fig. 5.12a and 5.12b were grown by CZ process. To experimentally observe the effect
of lowering hole-injection on J0 of Si/P3HT heterojunction devices, wafers grown by
the FZ process were used instead of CZ to make devices with the structure shown in
Fig. 5.10(c). The doping of the FZ wafers was also higher: 7×1015 cm-3 instead of
4×1014 cm-3 for CZ wafers used in earlier experiments. The P3HT layer was 25 nm
thick.
Compared to heterojunction devices on lower-doped CZ Si, Si/P3HT devices fab-
ricated on higher-doped FZ substrates show an marked decrease in J0 - from 1.6×10-8
136
(a) (b)
Figure 5.19: The I-V characteristics of Si/P3HT devices fabricated on 4×1014 cm-3
doped CZ Si and 7×1015 cm-3 doped FZ Si. (a) In dark. (a) Under ≈ 0.1 sun
illumination.
to 2.9×10-9 (Fig 5.19(a) and Table. 5.4). The open-circuit voltage under 0.1 sun il-
lumination also showed a corresponding increase; from 0.41 to 0.45 V (Fig 5.19(b)).
Ideality factors are once again high, which makes it hard to compare the performance
between devices. To make a fair comparison between the photovoltaic performance
of the devices, the “implied J0 ” is calculated. In terms of “implied J0 ”, the currents
in the FZ device are three times lower than the CZ device (Table 5.4).
These results conclusively prove that dark-currents in Si/P3HT devices are dom-
inated by minority-carrier injection from the anode into silicon, and not electron in-
jection from silicon into the anode. Furthermore, we have demonstrate devices with
an “implied J0 ” of only 63 pA/cm2 , while keeping processing temperatures below 150
.
137
Pd (50 nm) P3HT
3 mm
3 mm 4 mm
n-Si Substrate
5 × 1014 cm−3
(500 µm)
Ag 50 nm 4 mm
(a) (b)
Figure 5.20: Device structure of Si/P3HT heterojunction solar cells where the semi-
transparent blanket metal is replaced with a 20% metal grid for the top electrode.
(a) Side View and (b) top view.
All the Si/P3HT heterojunction devices fabricated until now had a semi-transparent
Pd-layer as the top electrode, which absorbs roughly 50% of the incoming light. In an
actual solar cell, such high tranmission losses would be unaccpetable and an anode
structure with less light absorption is required.
The easiest alternative is to replace the blanket Pd-layer with a metal-grid with
fingers. Depending on the density of metal fingers, ∼5 to 40% of the surface area
in our experiment is covered by metal and does not transmit light. The fabrication
procedure for making such devices was similar as before, except for anode metal
pattering. Instead of using a shadow mask with 1 mm radius holes, a shadow mask
with 100 µm wide fingers was used (Fig. (b)). The separation between fingers was
varied to get patterns with approximately 5, 10, 20 and 40 % shadowing. The active
area of the solar cell is not very well-defined, because the extent of lateral conduction
is unknown. In the results presented here, the area was assumed to be 4 mm x 4 mm.
Current-voltage characteristics of the metal-grid devices were measured under a
microscope light calibrated with an approximate intensity of 0.1 and 1 sun (Fig. 5.21).
138
(a) (b)
(c)
As before an open-circuit voltage of ∼ 0.4 V was obtained for all devices (Fig. 5.21(a)).
One expects that higher shadowing losses in a more dense metal-grid would lead
to lower JSC , but interestingly, the values of JSC show a monotonic increase with
increasing metal-grid density (Fig. 5.21(b)).
The issue lies with poor lateral conductivity of the P3HT layer, which forces the
photogenerated holes in silicon to diffuse laterally in silicon, to a location with a
metal finger, instead of going vertically towards the anode. The longer the distance
139
the carriers need to diffuse in silicon, the higher the chance they will be lost to
recombination. By reducing the spacing between metal fingers, i.e. increasing the
grid density, the lateral diffusion in silicon is reduced and consequently JSC increases.
The data for fill factor also confirm the presence of lateral resistance effects in these
devices. Since the resistance is limited by lateral conduction, smaller spacing between
metal fingers reduces total series resistance and increases fill factors (Fig. 5.21(c)).
Thicker P3HT layers might be used to ameliorate the problem of lateral resistance
to some extent, but the improvement would come at the expense of increased trans-
mission losses (lights transmits through a thicker P3HT layer) and a higher vertical
resistance (carrier need to travel longer in the vertical direction). A more robust
solution is to integrate a transparent conductor on top P3HT-coated silicon wafers.
The transparent conductive semiconductor used in this study is the organic Poly(3,4-
ethylenedioxythiophene) poly(styrenesulfonate) (PEDOT:PSS) which was bought from
Sigma Aldrich (High-conductivity grade, 2.2-2.6% in water). Chemically PEDOT:PSS
is a water-based dispersion of a wide energy-gap organic semiconductor, PEDOT,
doped with the acid PSS. The fact that PEDOT is a doped-semiconductor is not im-
portant, the reason it is used in this work is because it is transparent. PEDOT:PSS
dispersion is a bluish-colored ink that can be spin-casted to form thin tranparent
films that absorbs <5% of the solar light. The advantages PEDOT has over con-
ventaional tranparent conductors, like ITO and ZnO, are a) PEDOT does not need
to be sputtered, a process that can mechanically damage the delicate organic layers
underneath, instead PEDOT layers are spin-coated. b) PEDOT has a higher work
function of ∼ 5.1 eV than ∼4.5 eV of ITO, which should result in a higher built-in
field in silicon and consequently a higher VOC should be obtained.
To measure the vertical resistivity of PEDOT:PSS layer, a special test structure
140
PEDOT:PSS
(220 nm) Pd (50 nm)
Pd (50 nm)
Oxide 2 mm
Si
Al (200 nm)
Figure 5.22: (a) Structure and (b) IV characteristics of the device used to measure
vertical resistivity of PEDOT.
was fabricated (Fig. 5.22(a)) in which PEDOT was sandwiched between two layers of
metal. The bottom metal (200 nm Al) and top metal (50 nm Pd) were deposited by
thermal evaporation, while PEDOT was spin-coated before the top metal deposition.
If the lateral current-spreading is ignored then the top-metal electrode defines the
device area. The top electrode was pattered by a shadow mask with holes (1-mm
radius). The I-V characteristics were linear with voltage, as they should be for a
resistor-like structure. Ignoring the contact resistance, the vertical resistivity of the
PEDOT layer can be simply extracted from the measured resistance, given the thick-
ness of the PEDOT layer (t=220 nm) and top-metal electrode area (A = 3.14×10-2
cm2 ):
V A
ρvert = (5.14)
I t
= 8200 Wcm
For lateral resistance of PEDOT, a different test-structure was used. There was
no back-electrode. Instead PEDOT was spin-coated on an oxide covered silicon
wafer. The top electrode was then deposited through a shadow mask. In the fi-
141
Spacing (L)
Pd (50 nm)
PEDOT:PSS
Spacing (L) (220 nm)
W
Oxide
Si
(c) (d)
Figure 5.23: Structure and (b) I-V characteristics of the device used to measured
lateral resistivity of PEDOT. (c) Extracted lateral resistance for different spacing
(L). W = 0.5 cm. (d) Calculated power lost due to lateral and vertical resistivity of
PEDOT as a function of operating current level. An additional metal grid (with 10
% shadowing) reduces the lateral loss to ∼ 1% at a current density of 35 mA/cm2 .
nal structure metal pads were placed at various spacings(L) - from 250 µm to 5500
µm (Fig. 5.23(a)). The measured I-V characteristics between any two adjacent pads
were linear (Fig. (b)), and the extracted resistance (Rlat ) increased monotonically
with spacing L (Fig. (c)). The linear fit of the dependence of Rlat on L has an in-
tercept on the y-axis, which is a measure of the contact resistance of the test fixture
and is not a PEDOT property. On the other hand, the slope of the linear fit directly
142
relates to the lateral resistivity of PEDOT (ρlat ). Given pad width (W =5 mm) and
PEDOT thickness (t = 220nm)
= 0.003 Wcm
It is worth noting that the resistivity is the two orthogonal directions is different
by 6 order of magnitude (ρvert /ρlat ∼ 106 ). Anisotropy in conductivity has been
previously reported in PEDOT:PSS films [115], though the difference in conductivity
was only three orders of magnitude (ρvert /ρlat ∼ 103 ) and not six. The reason for the
large anisotropy was traced to the morphology of the PEDOT layer, where PEDOT
and PSS domains were formed in a manner that charge hopping in one lateral direction
was much easier than vertical direction [115]. In-depth morphological study of the
PEDOT:PSS layers used in this study have not yet been conducted.
Due to vertical and lateral resistivity, PEDOT will suffer ohmic losses. To estimate
the severity of the losses in the solar cell due to the PEDOT layer, ohmic losses were
calculated for different values of current density. We are particularly interested in the
losses at 1 sun illumination, or at a current density of 40 mA/cm2 .
Assume a the device is operating at a current density of Jop (in mA/cm2 ). The
ohmic losses due to vertical resistance, Ploss,vert (in W/cm2 ), is given by
2
Ploss,vert = Jop ρvert t (5.16)
143
Light
S/2
Busbar
Lateral Current
PEDOT
t
P3HT
Si Vertical Current
Figure 5.24: Power losses due to lateral resistance of PEDOT layer. Current in the
lateral direction is not uniform so loss calculations is not simple.
erated close to the busbar are minimal but the current generated far from the metal
finger needs to travel a long resistive path to the finger, so its lateral ohmic losses are
higher. The power losses (Ploss,lat , in in W/cm2 ) in such a scenario can be estimated
by [116]
2 S2
Ploss,lat = Jop ρlat (5.17)
12t
where S is the average distance between two fingers and t is the PEDOT thickness.
For a 4 mm× 4 mm device with one metal finger in the middle of the device S is 4
mm.
Using the above equation, the ohmic losses due to the vertical and lateral resistance
at different current density can be calculated (Fig. 5.23(d)). At AM1.5 the total
optical power incident on a device is 100 mW/cm2 . The fraction of the total optical
power lost due to finite vertical resistivity is
Ploss,vert
Ploss,vert (% of 1 Sun) = × 100
100 mW cm2
Ploss,lat
Ploss,lat (% of 1 Sun) = × 100
100 mW cm2
144
In Fig. 5.23(d), these normalized power-loss values are plotted as a function of operat-
ing current level (Jop ). While the losses due to vertical resistivity are minimal - <1%
of total incoming power at 1 sun illumination - the losses due to lateral resistivity are
significant - ∼ 10 %. The simplest way to reduce the lateral resistance is to add more
fingers connected by a busbar, i.e. add a metal grid on top of PEDOT layer. This
reduces the average S, which decreases lateral ohmic losses, a per Eq.(5.17). For a 10
% metal-grid with 100 µm wide fingers, the average finger spacing is 900 µm. With
the metal grid, the calculated ohmic losses due to the lateral PEDOT resitivity are
only 1% of total power (Fig. 5.23(d)). Of course, the decreased resistance comes at
the expense of a lower JSC due to increased light reflection losses at the top surface.
Algorithms to precisely optimize the trade-off between JSC and fill factor have been
published elsewhere [37, 117]. In devices fabricated in this study a 5-10 % metal grid
was added on top of the PEDOT:PSS layer.
Next, Si/P3HT heterojunction solar cells with a PEDOT and metal grid as the trans-
parent anode stack were fabricated. The fabrication procedure was same as before
- P3HT was spin-coated on clean and hydrogen-passivated silicon - except for two
additions:
(a) PEDOT layers were spin-coated at 2000-6000 rpm (for 120 s) on top of P3HT-
coated wafers to form the PEDOT-P3HT-Si structure of Fig. 5.25(a).
(b) In heterojunction solar cells of Fig. 5.20b, the high lateral resistance of P3HT
automatically isolated two adjoining devices on a silicon wafer. However, due
to lower lateral resistance of the PEDOT later, a more robust method is needed
to isolate adjoint devices. In this study, devices were isolated by scratching
away the organic layers (Both PEDOT and P3HT) using a thin probe. On a
16 mm × 16 mm Si piece, 3 vertical and 3 horizontal scratches were made,
145
each 4 mm apart, to give sixteen 4 mm × 4 mm sized devices. The scratching
method does lead to some inaccuracies in calculating device area, e.g. carriers
collected within a diffusion length of the device edge may also contribute to
the photocurrent, artificially increasing the short-circuit current. However, the
uncertainty due to these effects is small (∼ 100 µm for CZ wafers) compared to
the size of the device (4 mm), so it is ignored in the calculations.
P3HT PEDOT
Pd 50 nm 10 mn 220 nm
Ag 50 nm
4 mm
(a) (b)
(c) (d)
Figure 5.25: Photovoltaic response of Si/P3HT heterojunction solar cells with PE-
DOT:PSS, at approximately 0.1 and 1 sun illumination, with different density of top
metal grid - 5%, 10%, 20 %, and 40%. (b) Open-circuit voltage (c) Short-circuit
current (d) Fill factor
146
The photoresponse of heterojunction devices with PEDOT and different top metal
grids is shown in Fig. 5.25. The open-circuit voltages of these devices were around 0.57
V under a microscope light roughly with an intensity of approximately 1 sun. Due to
better carrier collection at the anode and reduced lateral resistance, the devices show
substantially higher JSC (Fig. 5.25(c)) than what was measured on devices with just
the metal grid (Fig. 5.21(b)). Also, unlike the device with just metal grid, the JSC for
device with PEDOT decreases with increasing metal grid shadowing, demonstrating
that carrier collection is not a problem anymore. The improved lateral resistance also
helped increase the fill factors at lower illuminations (where the current density is <5
mA/cm2 ) - from Fig. 5.21(c) to Fig. 5.25(d) verses ). However, at higher illumination
the current density was high and higher resistance losses caused deterioration in the
fill factors - from 60% to only 25% only. Fill-factors still increase by decreasing the
distance between metal fingers, suggesting that lateral resistance is still an issue.
Loss Mechanisms
The low fill factors are a result of ohmic losses in the heterojunction device, and to
demonstrate high-efficiency devices, series resistance needs to be decreased. Ohmic
losses in the PEDOT layer were estimated in previous section, and they are expected
to be lower (Fig. 5.23(d) than what is experimentally measured in the devices, so
other sources of resistance need to be analyzed.
Organic layers are known source of series resistance. Doping does decrease the
resistivity of the organic layers [118], however these methods are neither easy nor
well-understood.
Resistivity of the Si wafers used in the devices above was ∼ 10 Wcm. So Si is
expected to add around 30 W of series resistance to a 4 mm × 4 mm sized device.
By switching to a 0.8 Wcm doped wafer, the resistance due to Si can be reduced to
only ∼ 2 W. The added benefit of higher-doped wafers is reduced hole injection and
147
increased VOC .
Lateral Current
Finger
Busbar W
L
S/2
Lateral Current
Top View
(a) (b)
Figure 5.26: (a) Schematic of a metal finger attached to a busbar. The current is
collected laterally and transported to the busbar. (b) Calculated ohmic power losses
as a function of operating current level, due to resistivity of metal fingers. Compared
to the original ‘50nm Pd’ fingers losses in the improved ‘Ag/Pd’ fingers are lower.
Another sources of series resistance is the metal grid. Typical resistivity of palla-
dium is ∼ 10-5 Wcm. One 50 nm thick finger of the metal grid made out of Pd (3 mm
long × 100 µm wide) contributes 30 Ws to series resistance. To reduce the resistance
of metal fingers, the Pd layer (50 nm) can be replaced by Ag, whose resistivity is ten
times lower (∼ 10-6 ohm cm) than Pd. However, the work-function of Ag is only 4.5
eV, while that of Pd 5.0 eV. To maintain the high work-function of the anode, and
consequently the built-in field in silicon, the Pd finger is not replaced with Ag, but
replaced by a 15nm/200nm thick Pd/Ag stack. Compared to a simple Pd finger (30
W), the resistance of a single metal finger made with Pd/Ag is <1 ohm. Calculations
of the ohmic losses due to resistive metal fingers is slightly complicated, once again
due to non-uniform lateral current.
Assume that the metal-grid consists of metal fingers that are uniformly spaced a
distance S apart and run along the length of the device (L). All fingers are connected
in parallel to a busbar of zero resistance (Fig. (a)). Each metal finger is L long, w
wide, and d thick made of metal with resistivity ρm . The current is collected in the
148
lateral direction from a area ± S/2 around the metal finger. The ohmic losses (Ploss,m ,
in W/cm2 ) depend how far the current is from the busbar. If the operating current
density is Jop the ohmic power loss is given by[119]
2
2 SL
Ploss,m = Jop ρlat (5.18)
3wd
The calculated power loss values due to the Pd and Pd/Ag metal fingers are shown
in Fig. (b). The use of lower resistivity metal along with thicker finger (d = 200nm),
substantially reduces ohmic losses in the device.
Another area of potential improvement is the short-circuit current. Conventional
Si solar cells have a JSC of around 40 mA/cm2 at AM1.5, while the heterojunction
device above yielded a JSC of only 23 mA/cm2 . One obvious source of loss if the metal
grid which reflects 10% of the light. Parasitic absorption of photon in the organic
layers could also reduce the photon flux seen by silicon. To check for this possibility,
transmission measurements were performed using a standard Si photodiode.
Organic layers of differing thicknesses were deposited on glass slides. Next the
photodiode was covered by these glass sides and the photoresponse was measured by
a ammeter under 1 sun illumination. Assuming the internal quantum efficiency of
the silicon diode is unity, the fraction of photons being absorbed in the organic layers
were calculated (Fig. 5.27). The results show that losses in a device with a 10 nm
P3HT layer and 80 nm PEDOT layer, the organic layers account for only ∼ 7% of
loss in JSC (approximately 3 mA/cm2 ).
The present generation heterojunction cells do not have light trapping structures,
such as an anti-reflection coating [120, 121, 122] or surface texturing [123], so photon
losses due to reflection off the silicon surface are expected to be high. The refractive
index of silicon is around 3.4. If the pristine silicon surface is exposed to air (the
worst case), the amount of normally incident light lost due to reflections off the top
149
Top View
(a) (b)
(c)
2
3.4 − 1
R= ≈ 30%
3.4 + 1
The refractive index of PEDOT is ∼1.33-1.55 [124]. Thus there is enough contrast in
the refractive indexes of Si and PEDOT, that PEDOT can function as a single layer
anti-reflection coating on silicon. In the next sub-section, the PEDOT thickness was
set at 80 nm, a thickness where it can serve as a quarter-wavelength anti-reflection
150
Table 5.5: The extracted device parameters for Si/P3HT heterojunction solar cell
(Fig. 5.28(a)) under AM1.5. Implied J0 calculated from Eq. (1.4), using measured
VOC , and JSC .
Device Parameters
Device
Extracted J0 n JSC VOC Implied J0
(nm) (A/cm2 ) (mA/cm2 ) (V) (A/cm2 )
No P3HT, No PEDOT Cell 9×10-7 1.01 1.5 0.21 4.4×10-7
Si/P3HT heterojunction 9×10-9 1.37 29 0.59 3.4×10-12
Increased Si substrate doping from 5×1014 cm-3 to 1016 cm-3 , leading to reduced
151
10% Ag/Pd Grid P3HT PEDOT
200 nm/15 nm 10 mn 80 nm
Ag 200 nm
4 mm
(a) (b)
(c)
Figure 5.28: (a) The structure of the optimized Si/P3HT heterojunction solar cell.
(b) I-V characteristics of the heterojunction solar cell in dark. (c) I-V characteristics
of the heterojunction solar cell under AM 1.5 illumination.
circuit voltage of 0.6 - 0.7 V. The measured short-circuit current density is 29 mA/cm2 ,
which is only about 30% lower than the theoretical maximum of 42 mA/cm2 . Based
on the measured valued of VOC and JSC , the “implied J0 ” of the heterojunction is
only 3.4 pA/cm2 , an improvemt of over five orders of magnitude over the Schottky
device (with metal grid but no PEDOT or P3HT) (Table 5.5).
The PEDOT:PSS/P3HT stack and top metal account for around 7% and 10%
reduction in the short-circuit current. The rest of the loss in short-circuit current (∼
152
13%) is either due to the reflections off the silicon and organic surface, or due to loss
of the red photon wich are absorbed at depths greater than the diffusion lengths of
the silicon substrate. The overall power efficiency of the device is 10.1%, which to
the best of our knowledge is a record for these type of devices [125, 59, 54, 62] by a
factor of two and better than the state of the art all-organic solar cells [6].
5.5 Conclusion
153
Chapter 6
Stability of Silicon/Organic
Heterojunctions
6.1 Introduction
Organic thin-films degrade in oxygen and water rich environments. The problem is
especially acute in organic solar cells, because light accelerates many of the degrada-
tion mechanism, limiting the performance even further [10, 11, 126]. On the other
hand, silicon solar cells are very stable, with guaranteed performance for 25+ years.
In order to compete with conventional silicon solar cells, the heterojunction solar cells
also need to demonstrate 20 year lifetimes.
The first step toward improved stability is characterization. In this chapter data
from some preliminary stability studies are presented for PQ-passivated silicon sur-
faces and Si/P3HT heterojunction solar cells. The basic question we are asking is:
how quickly do silicon/organic heterojunctions degrade?
The stability of PQ passivation is studied using the minority-carrier lifetimes and
I- V characteristics of field-effect devices. The stability is found to be a strong function
of encapsulation. At an unprotected PQ-passivated silicon surface (Chapter 3), the
154
minority-carrier recombination lifetime were stable for only 1-2 days (Section 6.3.2).
However, at resin encapsulated silicon/PQ surfaces, the quality of the passivation not
only stays stable but improves over a period of 2-4 months (Section 6.3.3). A simple
mechanism to explain this behavior is proposed in Section 6.3.4.
“All-organic” solar cells degrade via many mechanisms [11] (Table 6.1). Out of
the known mechanisms, the first two are not relevant in the silicon/organic hetero-
junction cells simply because the materials and interfaces most susceptible to degra-
dation in “all-organic” cells are not present in heterojunction cells, e.g. there is no
ITO/PEDOT interface or low work-function electrodes such as Ca/Al. Furthermore,
most of the light absorption and carrier separation occurs in silicon, which is stable
material, so issues relating to stability of the bulk heterojunction are not relevant for
SOH solar cells.
The two materials that are used in SOH solar cells and whose stability is circum-
spect, are the organic polymer P3HT - which is known to photo-oxidize [127, 128], and
transparent conductor PEDOT:PSS - which degrades with both humidity [129, 130],
heat [131], and light [132]. Also of concern is the stability of Si/organic interface
because the unpassivated Si surface may get more “defective” with time and lose
its ability to efficiently separate photogenerated charge carriers. As of yet, we have
not conducted any stability studies on the Si/P3HT heterojunction solar cells, but
stability is a prime subject for future research.
All the PQ passivation steps were consistent with recipes described before (Chapter 3).
A 10 nm thick layer of thermally evaporated PQ was used to passivate silicon. All the
minority-carrier recombination lifetime measurements were done with WCT-120 using
test structures that are sensitive to recombination at the top surface. Metal-insulator-
155
Table 6.1: Known degradation issues of “all-organic” solar cells, and their relevance
to Si/organic heterojunction solar cells.
Degradation of ITO due to PEDOT [135, 136] No, ITO/PEDOT interface in the
anode stack
The PQ-passivated wafers showed a strong response to intense light and in certain
cases a marked improvement in passivation quality was measured. The effect is re-
ferred to as “light-annealing” in this thesis.
The effect was observed, quite by accident, while conducting routine minority-
carrier recombination lifetime measurements using the WCT-120. The process flow
was similar to lifetime experiments described in Section 3.6. As before, the recombi-
nation lifetime of a special silicon wafer was measured, whose front and back surfaces
were passivated with a high-quality thermal-oxide. Next a 1-inch diameter hole was
etched on the top surface. After chemical cleaning, PQ was deposited on the exposed
156
PQ
(10 nm)
Si n-type
5×1014 cm−3
Thermal Oxide
(150 nm)
(a) (b)
Figure 6.1: (a) Structure of the PQ-passivated test device. (b) Light from a Xe-flash
bulb and UV light increases the effective recombination lifetime of minority-carrier
in the test device. The extracted surface recombination velocity (SRV) at the Si/PQ
interface is also plotted.
silicon, and the recombination lifetime was measured again. The structure of the test
device is shown Fig. 6.1(a).
Each measurement by the WCT-120 is accompanied by a light pulse from a Xe-
flash bulb, to generate excess carriers in silicon (details of the instrument were dis-
cussed in Chapter 3). It was observed that after each successive flash, the PQ-
passivated samples showed a small but steady improvement in the recombination
lifetime. The effect was cumulative and stable so that, after 20-30 flashes, the effec-
tive lifetime of a p-Si wafers, at a minority-carrier density of 1015 cm-3 , increased from
31 µs to 52 µs (Fig. 6.1(b)). This corresponds to a decrease in surface recombination
velocity from 700 cm/s to 279 cm/s. Th effect is quite striking and points to some
light-mediated reaction between silicon and PQ that is reducing the electrically active
defect-states at the interface.
Some of the known reactions between organics and silicon are induced by light
[138, 139, 140]. It is possible that the wide-spectrum intense light from the Xe-bulb,
157
is enabling a similar chemistry between PQ and silicon. To investigate, samples were
exposed to ultra-violet light from a mercury lamp stabilized at an intensity of 2
W/cm2 at 365 nm wavelength (lamp of the MA6 mask aligner in the cleanroom at
CI1 setting). First the sample was exposed for just 10 s. The lifetime jumped from
56 µs to 69 µs. The sample was then exposed for another 20 s, and again the lifetime
increased to 72 µs. The lifetime remained at 72 µs even after a subsequent 90 s long
UV exposure, and so the experiment was paused and sample was left in air for the
next 5 hours.
After 5 hours when the sample was remeasured. Surprisingly, the slow increase in
lifetime continued, and with every flash the lifetime increased a little bit, till it got
stabilized at 80 µs. A subsequent 200s UV exposure yielded the peak lifetime of 88
µs. A further 300 s UV exposure caused a small drop in lifetime so no further UV
158
were flashed ∼ 10 times after PQ deposition.
It is possible that the light-annealing we observe is nothing special. The surface
states at the silicon/PQ interface might just be trapping charges which causes a
decrease is electrical activity of the defects without affecting their density. There are
two reasons that this scenario is unlikely:
(a) The effect was never observed in other samples, even those with large defect
densities such as native-oxide coated wafers or HF/I2 passivated wafers. If
anything a decrease in lifetime is the norm (Fig. 3.3(b)).
The PQ layer is a soft organic material and only a 10-nm-thick layer covers the
silicon surface, so there is not much protection for the silicon-PQ bonds. Fig 6.2(b),
shows the recombination lifetimes and surface recombination velocity of the same
PQ-passivated Si wafer that was measured in Section 6.3, as function of time that
the wafer was exposed to ambient air. The time is measured from the moment the
sample was unloaded from the PQ evaporation chamber, which was maintained under
vacuum, until it was exposed to atmosphere for unloading. During the whole window
of measurement, from 10 minutes to approximately 1 month, the sample was kept in
ambient air in the clean room, with no special protection from the water, oxygen or
light.
159
PQ
(10 nm)
Si n-type
5×1014 cm−3
Thermal Oxide
(150 nm)
(a) (b)
Figure 6.2: (a) Structure of the PQ-passivated test device. (b) The effective lifetime
and extracted surface recombination velocity (SRV) of the test device, demonstrating
air stability of an unprotected Si/PQ interface.
The details of measurements taken between 10 minutes and 600 minutes, were
discussed in previous section. After the “light-annealing”, the passivation remains
pretty stable for at least another 24 hours Fig 6.2(b). However, measurements taken
a month later reveal that the lifetimes falls and reverts bck to the value measured
when sample was covered with native oxide Fig 6.2(b). The lifetime could not be
improved by any further light-annealing.
There is a remote possibility that the improved passivation was simply a result
of good-quality oxide being formed during the UV treatment (we could be forming
ozone). However, it is hard to imagine that at room temperature a good quality oxide
can form on silicon. Secondly, if such an oxide did form, it is hard to find a reason
for it to disappear over 1 month. Therefore we believe that passivation is actually
result of PQ bonding on silicon. A possible hypothesis to explain the degradation in
passivation that is observed will be discussed in Section 6.3.4.
160
6.3.3 Temporal Stability: With Encapsulation
Lifetime Measurements
The recombination lifetime test was similar to the one described in the last section.
The test samples were PQ-passivated p-type silicon wafer with oxide on the back
surface. Other than the Xe-flash, no other light-annealing was performed on these
samples (No UV treatment).
Over a period of days, the effective recombination lifetimes improved from 43 µs to
69 µs, ultimately reaching 71 µs. Using the generalized lifetime analysis of Chapter 3
(Eq. (3.8)), surface recombination velocities were extracted from lifetime data. From
day 0 to day 6, the surface recombination velocity at Si/PQ interface fell from 250
cm/s to only ≈3 cm/s, comparable to very high quality Si/SiO2 interfaces.
On the seventh day the samples was washed with methanol to remove the pho-
toresist (and if possible PQ). If the improved passivation was due to some “magical”
oxide and not PQ, the lifetimes would continue to be high. However, the measured
lifetime permanently fell to only 18 µs, the value expected when the wafer was coated
with native oxide. These results show that a) the passivation is due to some sili-
con/organic interaction that can be washed sway with a solvent and b) if encapsu-
lated, PQ-passivation improves in passivation quality over a period of 8 days.
161
PQ Photoresist
(10 nm) (1.4 µm)
Si n-type
5×1014 cm−3
Thermal Oxide
(150 nm)
(a) (b)
Metal-Insulator-Semiconductor Capacitor
162
(a)
(b) (c)
inversion, initially there are’t enough minority-carriers in silicon to form the inversion
layer. According to Shockley-Read-Hall theory, thermal processes try to correct this
lack of carriers by a net generation process. The rate at which minority carriers are
generated in silicon is characterized by the generation lifetime (τg ) which, like the
recombination lifetime (τr ), increases with a decrease in the desnity of bulk/surface
defects in the silicon. If the generation lifetime is short enough, then the inversion
layer is quickly formed and a flat inversion capacitance is measured as one increases
the gate voltage (e.g. Fig. 6.4(b)). If however, the generation lifetime is high, the
generation rate is unable to generate carriers fast enough to form the inversion layer
and the device goes into “deep-depletion”, characterized by the decreasing capacitance
in one sweep direction - from negative to positive bias (e.g. Fig. 6.4(c)).
163
Since the “4 months later” characteristics (Fig. 6.4(c)) show evidence of deep de-
pletion, i.e. improved generation lifetimes, it stands to reason that PQ-passivation
has improved. In comparison, a similar device, coated with AZ5124 but without PQ,
does not show any modulation of the Fermi level or improvement in C-V characteris-
tics. In both “before” and “after” characteristics, the Fermi-level is pinned, indicating
very high surface defect densities. So clearly, the improvement is related to Si/PQ
interaction and not Si/AZ5124 interaction.
164
(a) (b)
(c) (d)
measurements show that the silicon surface passivation with the organic molecule
9,10-phenanthrenequinone improves with time, if the surface is protected with AZ5124.
AZ5124 on its own does not lead any improvements, so Si/PQ interface must account
for the improvements. The capacitors show deep depletion and n-channel transistors
show steeper subthreshold slopes and lower off currents. The results can be explained
165
(a) (b)
Table 6.2: The change in the PQ-passivated MISFET characteristics, over a period
of two months. Samples were stored in a glove box.
by asserting that, over time, PQ reacts with more of the unsatisfied Si dangling bonds.
In this section a simple model to explain these two observations at the Si/PQ
interface is presented. This is only a working hypotheses and more experimentation
166
is required to make a strong statement about its validity. The model is based on two
assertions:
2. Oxygen and humidity can diffuse through the thin PQ layer to reach the silicon
surface. Even at room temperature, a native oxide is formed within hours on
a pristine H-terminated silicon surface. During native oxide formation, oxygen
penetrates and oxidize the first few monolayers of silicon. If we assert that PQ
is a permeable organic layer that can allow oxygen and water vapor to pass
through, then silicon atoms underneath the PQ layers could oxidize over time,
forming a native-oxide layer underneath the PQ-passivated Si surface. The
underlying silicon atoms can continue to get oxidized, even if the uppermost
silicon atoms remain bonded to PQ.
The overview of the proposed mechanism is shown in Fig 6.7. When the PQ-
passivated surface is left unencapsulated, oxygen and water diffuse through the PQ
layer and, over time, start oxidizing the silicon atoms underneath PQ. The resulting
native-oxide underlayer is electrically very “defective”. When lifetime measurements
are taken, the surface recombination properties are limited not by the Si/PQ interface,
but by the Si/native-oxide interface. Thus, even though PQ atoms are bonded to
167
(a) (b)
Figure 6.7: Mechanism to explain stability of encapsulated Si/PQ interface. (a) When
un-protected, passivation degrades. (b) When encapsulated, passivation improves.
6.4 Conclusion
168
Chapter 7
169
junctions could be used in crystalline silicon photovoltaics as a replacement for
diffused p/p+ and p-n junctions, thereby eliminating all the high-temperatures
steps required for dopant diffusions.
2. A novel organic-based passivation scheme for Si (100) surfaces has been devel-
oped, that reduces silicon surface defect density, to less than 1012 cm-2 , with-
out the use of any high temperature steps. The passivation precursor, 9,10-
phenanthrenequinone (PQ), is a semiconducting small-molecule that forms a
type-I heterojunction with silicon. The low surface defect density is demon-
strates by the low surface recombination velocity, less than 150 cm/s, measured
at the Si/PQ interface. Furthermore, Fermi-levels at the Si surface are not
pinned and can be modulated over a wide range under the electric field of an
insulated gate. Low surface defect-density also allows high field-effect mobility
to be obtained at the PQ-passivated surfaces - 600 cm2 /V.s and 50 600 cm2 /V.s
for electrons and holes, respectively.
170
lation on the stability of silicon/organic heterojunctions were demonstrated.
1. Efficiency: The AM1.5 power conversion efficiency of the Si/P3HT solar cells
demonstrated in this work is 10.1%, the highest number reported for Si/organic
heterojunction devices. Furthermore, 10.1% is at least as high if not higher
then published efficiencies for all-organic solar cells (∼ 8 %) [6]. Further, sim-
ulations of double-sided silicon/organic heterojunction show that hybrid cells
can achieve power efficiencies of over 20%, matching the current state-of-the-art
homojunction and heterojunction HIT solar cells.
2. Cost: While cost advantages provided the motivation for the work, the thesis
does not analyze the cost implications in any detail. However, the energy and
capital-investment savings achieved by replacing high-purity/high temperature
processing with room-temperature spin-coating (as in all-organic photovoltaic)
are self-evident. There are further cost advantages. For instance, common
impurities in silicon are known to get activated (thereby killing carrier lifetimes)
at high temperatures, so a low temperature process may allow use of cheaper
silicon wafers. Overall, we expect significant cost savings in the fabrication costs
of crystalline Si solar cells.
171
layers degrade. Good news is that the device has not been optimized for sta-
bility yet. Many of the known solutions that increase the device lifetime have
not been integrated into the device, e.g. encapsulation barriers, more stable
organics, etc. The issue of stability is a more general problem that plagues
all organic-based devices and there is growing body of research that deals with
precisely this issue. Heterojunction cells will also benefit from these advances.
In fact this leads to a related question - are silicon/organic structures more
reliable than all-organic cells? We have reason to believe that they are. Simply
because the silicon/organic structures are inherently immune to many of the re-
liability problems that plague all-organic cells. For example, the silicon/organic
cells do not have a degradation-prone low work-function metal/organic inter-
face; unlike P3HT/PCBM organic devices, in heterojunction cells there is no
metastable inter-penetrating bulk heterojunction; silicon is the primary absorb-
ing layer which is known to be long-term stable; and in the ultra-thin P3HT
(∼10 nm)layers of heterojunction device there aren’t many excited states which
are usually the cause for photo-oxidize of the polymer.
172
Increasing the Short-Circuit Current
The fill factor of baseline silicon/organic heterojunction cell is only 0.59, mainly due to
the ohmic losses in the organic films, especially PEDOT:PSS. Addition of solvents like
DMSO, ethylene glycol, isopropyl alcohol, etc. is a well-known method of boosting
PEDOT conductivity [149, 150]. Certain surfactants, like Zonyl [151] and sodium
dodecyl sulfate [152], are also known to enhance the conductivity of PEDOT:PSS
thin films. Series resistance can also be reduced by doping P3HT layers with organics
173
Dark Current
EC − Dark Current
Photocurrent HOMO LUMO −
p+-Si EC
Anode Photocurrent
EC /LUMO Offset p-Si
EV
Cathode
n-Si Cathode Anode
Passsivation
Photocurrent + EV
Organic Passsivation + Photocurrent
Dark Current
LUMO Organic
Dark Current EV /HOMO Offset
Reduced Surface
Recombination HOMO
(a) (b)
Both the heterojunctions demonstrated in the thesis block electrons but no hole-
blocking heterojunction were presented. Complimentary heterojunctions that block
holes could be a possible direction of future work. As in the electron-blocking hetero-
junctions, hole-blocking heterojunction can also be in two flavors: minority-carrier
blocking (Figure 7.1a) and majority-carrier blocking heterojunctions (Figure 7.1b).
Many electron-conducting organic materials are known (PC60 BM, C60, DCN-
nCQA, etc.) with reported LUMO levels around 4.0 eV and HOMO levels greater
than 5.2 eV, so they should form hole-blocking heterojunction with silicon.
174
Reduced Surface
Recombination
LUMO
Anode EC
Cathode
n-Si
HOMO
EV Passivation
Photocurrent +
Organic 2
Dark Current
HOMO
Reduced Surface
Recombination
Figure 7.2: Band Diagram of double-sided heterojunction solar cell with a electron-
blocking layer at the top and hole-blocking layer on the bottom.
Heterojunction solar cells use organic materials, so the device lifetime is limited by
the stability of the organic layers. Fortunately, the organic layers perform a very nar-
row role in silicon/organic heterojunction solar cells and many of the most damaging
degradation mechanisms of “all-organic” solar cells are not active in silicon/organic
heterojunction solar cells (discussed in the introduction of Chapter 6). One future en-
deavor could be to characterize and improve the stability/reliability of silicon/organic
solar cells.
175
The design of a lifetime measurement setup for organic solar cells has been detailed in
literature [153]. Commercial systems are also available [154]. A similar set up needs
to be built/bought for silicon/organic heterojunction stability studies.
Degradation Mechanisms
176
More Stable Organics
If PEDOT limits the lifetime of heterojunction solar cells, it could be substituted with
more stable transparent conductors, such as SPDPA [160], Aedotron [161], ZnO and
ITO [145, 146, 147]. If the hydrogen-passivated Si/organic interface degrades quickly,
intermediate layers could be added between silicon and organic to stabilize the hetero-
junction interface, e.g. SAMs [73] or thin-oxide films [67]. If P3HT degradation limits
stability, any organic with the correct HOMO/LUMO offsets can be used in place of
P3HT for silicon/organic heterojunction cells, e.g. P3CT [126], transition metal ox-
ides, such as V2 O5 [162], solution-processed MO3 [163], and solution-processed NiOX
[164].
177