You are on page 1of 41

Metal Nano-Grids for Transparent Conduction in Solar Cells

Christopher P. Muzzilloa,b*
a
Department of Chemical Engineering, University of Florida, 1030 Center Dr,

Gainesville, FL 32611, USA


b
National Renewable Energy Laboratory, 15013 Denver West Pkwy, Golden, CO 80401,

USA

* Corresponding author. E-mail address: christophermuzzillo@gmail.com (C. P.

Muzzillo).

Abstract

A general procedure for predicting metal grid performance in solar cells was developed.

Unlike transparent conducting oxides (TCOs) or other homogeneous films, metal grids

induce more resistance in the neighbor layer. The resulting balance of transmittance,

neighbor and grid resistance was explored in light of cheap lithography advances that have

enabled metal nano-grid (MNG) fabrication. The patterned MNGs have junction

resistances and degradation rates that are more favorable than solution-synthesized metal

nanowires. Neighbor series resistance was simulated by the finite element method,

although a simpler analytical model was sufficient in most cases. Finite-difference

frequency-domain transmittance simulations were performed for MNGs with minimum

wire width (w) of 50 nm, but deviations from aperture transmittance were small in

magnitude. Depending on the process, MNGs can exhibit increased series resistance as w

is decreased. However, numerous experimental reports have already achieved

transmittance-MNG sheet resistance trade-offs comparable to TCOs. The transmittance,

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
neighbor and MNG series resistances were used to parameterize a grid fill factor for a solar

cell. This new figure of merit was used to demonstrate that although MNGs have only been

employed in low efficiency solar cells, substantial gains in performance are predicted for

decreased w in all high efficiency absorber technologies.

Keywords

transparent conduction; metal grid; metal nano-grid; transparent conducting oxide; series

resistance; transmittance

Graphical abstract

Drying cracked film


Electrospun nanofiber Laser print
EHD jet print Inkjet
Nanosphere Capillary flow
Screen print;
Nanoimprint assembly
flexographic

109 Grid fill Hybrid


perovskite
108 factor
Rsh,neighb (/sq)

CdTe front;
107 CIGS front;
106 CdTe back

105 HIT

104 CIGS back

103 GaAs back

102 Si;
GaAs front
101
102 103 104 105
w (nm)

Highlights

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
• General procedure developed for calculating metal grid performance in solar cells

• Processes have recently been developed to form metal nano-grids that outperform

TCOs

• Neighbor series resistance, transmittance and grid series resistance modeled for

MNGs

• Neighbor series resistance important, although neglected in previous work

• New metal grid figure of merit shows decreasing wire width enhances PV

performance

1. Introduction

Transparent conduction is crucial to many optoelectronic devices, and is

particularly vital to the photovoltaics (PV) community: efficient solar cells require high

optical transmittance of the solar resource along with high electrical power transmission.

Transparent conducting oxides (TCOs) are the industry standard for Si heterojunction with

intrinsic thin layer (HIT), CdTe, Cu(In,Ga)(Se,S)2 (CIGS), and hybrid inorganic-organic

perovskite technologies. The most common TCO is indium tin oxide (ITO; $1.6 billion

sales in 2013 [1]), but it has vulnerabilities: vacuum sputtering makes it relatively capital-

intensive, with low material utilization of expensive In [1], brittleness makes it unusable in

flexible devices [2], and chemistry makes it incompatible with some materials (such as

crystalline Si [3] and organic semiconductors [4]). The most popular alternatives to ITO

also have fundamental drawbacks: Al:ZnO is unstable in damp heat and soda-lime glass

environments [5-8], F:SnO2 has inferior transmissivity (T)-sheet resistance (Rsh) trade-offs

[5], Ag nanowires are highly unstable under typical operating conditions [9-16], and carbon

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
nanotubes and graphene are both too resistive for solar cells [17]. Besides TCOs, metal

grids have been deployed to conduct high currents over long distances in PV modules for

decades. For grids, smaller dimensions yield better performance [18], so the industry-

standard screen printing resolution (down to 30,000 nm [19]) has historically limited

performance. This limitation has recently been lifted, as new lithography strategies have

emerged to fabricate metal nano-grids (MNGs). MNGs can have better T and Rsh than

metal macro-grids and ITO in theory and practice [20-28], use cheap processes [27, 29,

30], use flexible substrates [31-33], and exhibit low degradation rates under mechanical

stress and damp heat [27, 34]. The MNGs are not contiguous films, so grid thickness can

be designed to reduce MNG resistance with less effect on optical transmittance, relative to

conducting films. The patterning processes lead to robust, well-formed wires and junctions,

relative to solution-synthesized Ag nanowires. MNGs can also replace opaque metal films

to enable new solar cell architectures, multiple junctions, and bifacial devices. While

quality p-type TCOs have eluded the scientific community for decades due to large hole

effective masses [35], the MNG work function can easily be tailored for n- or p-type

contacts.

A variety of processes and metals have already been used to fabricate MNGs with

better T-Rsh trade-offs than ITO (detailed below), but they have only been used as stand-

alone transparent contacts in solar cells with efficiencies up to 7.0% [36, 37], less than half

that of most commercial modules. Moreover, the direct comparison of ITO and MNG

transmittance-Rsh curves is misleading, as grids are not homogenous films. This

inhomogeneity induces more resistance in the adjacent semiconducting layer (“neighbor”),

but it also lifts the constraint of thickness being tied to transmittance. The present work

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
proposes a general method for calculating a metal grid figure of merit that includes

neighbor and grid resistance, with particular focus on MNGs and high efficiency solar

cells. Analytical calculations are compared with simulation results for neighbor resistance,

MNG transmittance, and MNG resistance. These are used to establish fundamental limits

for MNG performance, and identify the best MNG designs. The figure of merit is then used

to analyze potential MNG gains for every high efficiency PV absorber: silicon, HIT silicon,

CdTe, CIGS, GaAs, and hybrid organic-inorganic perovskites.

2. Methods

Neighbor resistance calculations and simulations are described below. Previous

analyses of neighbor series resistance considered how a two- or three-dimensional diode is

affected by a grid in detail [38-49]. In contrast, the limits of grid performance were

presently examined in terms of MNG process-relevant parameters. MNG sheet resistance

was simulated by using the finite element method (FEM) to solve for electric potential

throughout the MNG volume, and integrating it over a given current. A uniform current

source flowed in the plane of the solar cell and traversed some distance L, assumed to be

much greater than the MNG thickness. The optimal conduction direction was used for each

grid shape. To combine neighbor and MNG resistances, it was assumed that the

neighbor/MNG stack does not behave like two films in parallel [50], but is instead limited

by neighbor lateral resistance in aperture areas, and then by lateral MNG resistance in the

obscured areas (i.e. resistors in series) [51]. A resistivity of 700 Ω·cm was used for the CdS

neighbor—near the middle of the wide range reported at 1 sun for chemical bath deposited

CdS [52-55], although it was an order of magnitude greater than resistivity values used in

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
previous simulations [56-63]. The CdS thickness was set to 100 nm to avoid potential

pinholes and shunts. The MNG metal was copper with 1.68 · 10-6 Ω·cm [28] and 50 nm

thickness, unless noted. A short-circuit current density of 40 mA/cm2 [64] and open-circuit

voltage of 750 mV [65] were used. The finite-difference frequency domain (FDFD) method

was used to simulate two-dimensional electromagnetic fields. Plane waves incident on the

air/MNG/air or air/MNG/substrate structures were simulated for TE and TM polarization

of 280 – 1,360 nm wavelengths, chosen to bound the appreciable power in the air mass 1.5

spectrum. Surfaces with no roughness were assumed for all simulations. Periodic boundary

conditions were used on the sides of a single MNG unit cell. In all cases, simulation

domains and mesh sizes were optimized to reduce error.

3. Results

3.1. General approach

The optimal design for a grid or TCO minimizes the total power loss due to (1)

resistance in the neighbor layer (Rneighb), (2) diminished transmittance through the

transparent conductor, and (3) resistance in the transparent conductor. TCO films typically

cover the neighbor layer homogenously, making Rneighb negligible. On the other hand, grids

have inhomogeneity that greatly increases lateral conduction in the neighbor. The only way

to mitigate this Rneighb is to reduce the distance between grid wires—this is the reason

MNGs are advantageous. The limit to this trend is where the assumptions used in deriving

it break down, taken in this work to be a minimum wire width, w of less than 50 nm. Wires

of smaller width are more difficult to process, exhibit increased resistivity (relative to bulk

ρmetal) due to increased grain boundary, surface, and impurity scattering [66, 67], increased

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
junction resistance (similar to Ag nanowires) [1], depressed melting temperatures [12, 16],

and high surface area-to-volume ratios that lead to increased degradation rates [10, 13, 15].

The exact limitations must ultimately be identified experimentally for specific processes,

and the present work should guide the choice of those experiments while motivating them

with performance gain estimates. Below, the prospect of replacing TCOs with MNGs in

CIGS and CdTe solar cells is examined in detail, but all results can easily be scaled for

other applications.

Widely diverse grid shapes have been used in previous experiments and

simulations: gratings [68], squares [20], rectangles [69], triangles [70], hexagons [71],

truncated hexagons [72], hexagonal-packed circles [73], square-packed circles [74], multi-

sized circles [21], tetrakis squares [70], disordered circles [75], random wires [37], fractals

[76], hierarchical shapes [77], busbar with fingers [78], and various other shapes [79, 80].

The grating, triangle, hexagon, square, and hexagonal-packed circle were investigated in

the present study. Rectangles with 2x and 4x the square length were also examined, where

the widths were reduced by 2x and 4x, respectively. Three parameters completely defined

each shape, only two of which were independent: w, Taper, and space between wires (s;

where the unit cell period is the sum of w and s). An example is shown for the grating in

Fig. 1.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
Fig. 1. Schematic of a grating MNG showing aperture area, obscured area, w, s, hMNG, and

hneighb, as used in the text.

3.2. Neighbor resistance (Rneighb)

Wyeth first outlined the general procedure for calculating neighbor resistance for

any grid shape [51]. His analysis made several approximations: current generation is

uniform throughout the neighbor volume, but only in aperture areas, resistance is

dominated by lateral current flow (in the plane of the solar cell), planes dividing aperture

and obscured areas are effectively grounded, and neighbor diffusion current is negligible.

These assumptions lead to an equation for potential in the neighbor layer:

𝜕2 𝜑 𝜕2 𝜑 𝜌𝑛𝑒𝑖𝑔ℎ𝑏 ∙𝐽
+ 𝜕𝑦 2 = − = −𝑅𝑠ℎ,𝑛𝑒𝑖𝑔ℎ𝑏 ∙ 𝐽 (1)
𝜕𝑥 2 ℎ𝑛𝑒𝑖𝑔ℎ𝑏

This equation can be solved by setting potential at the aperture/obscured boundaries to zero

(𝜑(𝑥, 𝑦) = 0), or setting the potential’s normal derivative to zero at the uncontacted

boundaries (e.g., 𝜕𝜑(𝑥, 𝑦)⁄𝜕𝑦 = 0). A simple analytical solution results for the grating,

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
which can be multiplied by current density and integrated over the aperture area to find

Ohmic power dissipation:


𝑅𝑠ℎ,𝑛𝑒𝑖𝑔ℎ𝑏 ∙𝐽
𝜑= ∙ (𝑠 ∙ 𝑥 − 𝑥 2 ) (3)
2

𝑥=𝑠 𝑦=𝑠 𝑅𝑠ℎ,𝑛𝑒𝑖𝑔ℎ𝑏 ∙𝐽2 𝑅𝑠ℎ,𝑛𝑒𝑖𝑔ℎ𝑏 ∙𝐽2 ∙𝑠4


𝑃𝑛𝑒𝑖𝑔ℎ𝑏 = ∫𝑥=0 ∫𝑦=0 ∙ (𝑠 ∙ 𝑥 − 𝑥 2 ) ∙ 𝑑𝑥 ∙ 𝑑𝑦 = (4)
2 12

The voltage loss is power divided by the product of current density and aperture area

(𝑉𝑛𝑒𝑖𝑔ℎ𝑏 = 𝑃𝑛𝑒𝑖𝑔ℎ𝑏 ⁄(𝐽 ∙ 𝑠 2 )). The neighbor power loss density and series resistance are

also referred to the aperture area (𝑃𝑛𝑒𝑖𝑔ℎ𝑏 = 𝑃𝑛𝑒𝑖𝑔ℎ𝑏 ⁄𝑠 2 and 𝑅𝑛𝑒𝑖𝑔ℎ𝑏 = 𝑃𝑛𝑒𝑖𝑔ℎ𝑏 ⁄(𝐽2 ∙ 𝑠 2 )).

For more complicated shapes, the potential surface is readily solved numerically with the

FEM. The Pneighb that results from integrating that potential surface at a given Rsh,neighb, J,

and s can simply be scaled to new values of those parameters:

𝑅 𝐽 𝑠4
𝑃𝑛𝑒𝑖𝑔ℎ𝑏,𝑛𝑒𝑤 = 𝑃𝑛𝑒𝑖𝑔ℎ𝑏,𝑜𝑙𝑑 ∙ ( 𝑅𝑠ℎ,𝑛𝑒𝑖𝑔ℎ𝑏,𝑛𝑒𝑤 ) ∙ ( 𝐽𝑛𝑒𝑤 ) ∙ ( 𝑠𝑛𝑒𝑤
4 ) (5)
𝑠ℎ,𝑛𝑒𝑖𝑔ℎ𝑏,𝑜𝑙𝑑 𝑜𝑙𝑑 𝑜𝑙𝑑

For the grating, s is defined in Fig. 1 so that aperture area is s2. The above equation is valid

for any shape whose aperture area is a function of s2, so s can be set to a circle’s radius or

an equilateral triangle’s side length or height. If Taper is specified, then Pneighb and Rneighb

become functions of w, since (for the grating):


𝑠
𝑇𝑎𝑝𝑒𝑟 = 𝑠+𝑤 (6)

𝑅𝑠ℎ,𝑛𝑒𝑖𝑔ℎ𝑏 ∙𝐽2 ∙𝑇𝑎𝑝𝑒𝑟


4 ∙𝑤 4
𝑃𝑛𝑒𝑖𝑔ℎ𝑏 = 4 (7)
12∙(1−𝑇𝑎𝑝𝑒𝑟 )

Rneighb was calculated at constant Taper of 0.9 for each shape as a function of w in Fig. 2.

The circle had far more resistance, while the grating, 2x and 4x rectangle had slightly less

resistance, relative to the square, triangle, and hexagon. The circle performed worse due to

its poor tessellation, which resulted in a much smaller minimum wire width at a given Taper,

relative to the other shapes. All shapes had neighbor series resistances proportional to w2—

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
underscoring the importance of minimum wire width on performance for any grid

fabrication process. The neighbor power density dissipation axis in Fig. 2 was calculated

at an example JSC of 40 mA/cm2. At constant w of 50 nm, Taper is shown as a function of

Rneighb in Fig. 3. The Taper-Rneighb trade-off was best for the grating, followed by the 4x

rectangle, 2x rectangle, triangle, hexagon, square, and circle. The grating was only bounded

on two sides, so it led to a larger potential gradient for equal aperture area, relative to the

square. However, when the grating and square were compared at equal Taper and w, the

grating had a much smaller aperture area. The increased wire area along the square and

other high symmetry shapes’ edges did not reduce the potential well enough to compensate

for its reduction of Taper, making the grating superior. This situation was remedied by

increasing the aspect ratio of any shape, causing its potential surface to more closely

resemble a parabolic trough. This is why the 2x and 4x rectangle Taper-Rneighb curves

approach that of the grating in Fig. 3.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
Fig. 2. Neighbor series resistance versus minimum wire width at constant aperture

transmittance of 0.9 for the grating (purple), 4x rectangle (blue), 2x rectangle (green),

triangle (orange), hexagon (red), square (gray), and circle (black). Resulting power

dissipation density on the right axis was calculated at 40 mA/cm2.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
Fig. 3. Aperture transmittance versus neighbor series resistance at constant minimum wire

width of 50 nm for the grating (purple), 4x rectangle (blue), 2x rectangle (green), triangle

(orange), hexagon (red), square (gray), and circle (black). Resulting power dissipation

density on the top axis was calculated at 40 mA/cm2.

To probe the limits of Wyeth’s analytical model, FEM simulations were performed

with two additional components: current flow normal to the solar cell and in the obscured

area. The simulated current entered the neighbor from below, only in the aperture area, and

exited the neighbor at the grounded neighbor/MNG interface. As seen in Fig. 4 for the

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
grating, the analytical and simulated Rneighb were indistinguishable at large w. At small w,

the simulation asymptotically approached series resistance for current density flow normal

to the neighbor layer (𝑅𝑛𝑒𝑖𝑔ℎ𝑏 ≈ 𝜌𝑛𝑒𝑖𝑔ℎ𝑏 ∙ ℎ𝑛𝑒𝑖𝑔ℎ𝑏 ), while the analytical Rneighb was for

transverse current flow in the aperture area. The potential contour insets in Fig. 4

demonstrate these limits. The analytical and simulated curves were also indistinguishable

at high Rneighb and Taper (constant w; Fig. 5). At low Taper the analytical Rneighb was again

for transverse current flow in the aperture area. The simulated Rneighb was dominated by

current normal to the layer, which in this case was reduced by spreading to fill obscured

areas (see Fig. 5 insets). Future simulations at low Rneighb and high Taper can use the dark

resistivity of CdS in those obscured areas. Outside of this case, Rneighb is dominated by

potential gradients in aperture areas, so ρneighb at 1 sun is appropriate.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
Fig. 4. Analytical (black square) and simulated (red circle) neighbor series resistance

versus minimum wire width at constant aperture transmittance of 0.9 for the grating. Insets

are neighbor layer potential contours with shifted scales. Resulting power dissipation

density on the right axis was calculated at 40 mA/cm2.

Fig. 5. Aperture transmittance versus analytical (black square) and simulated (red circle)

neighbor series resistance at constant minimum wire width of 50 nm for the grating. Insets

are neighbor layer potential contours with shifted scales. Resulting power dissipation

density on the top axis was calculated at 40 mA/cm2.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
The analytical and simulated Rneighb only deviated at low w and low Rneighb. These

deviations stemmed from current flow through the thickness of the neighbor, which must

occur regardless of the grid design. Therefore, simulated Rneighb should only be used for

simultaneous optimization of solar cell Rsh,neighb and MNGs. While the grating had a

superior Taper-Rneighb trade-off, the differences among shapes were not large, except for the

poor-performing and poorly-packed circle. The increased connectivity of the triangle,

hexagon, square, 2x, and 4x rectangles may be preferred to the grating if a fabrication

process is prone to defects. MNG processes that produce elongated aperture shapes that are

closely packed are preferred.

3.3. MNG transmittance

The aperture transmittance for any MNG shape is the aperture area divided by the

total area. This definition is based on geometrical, or ray optics, and is only valid for

relatively large grid dimensions. As MNG thicknesses and wire widths approach the

wavelengths of visible light, wave optics become more important, and FDFD simulations

must be used to calculate transmittance [81-83]. The deviation of simulated transmittance

from aperture transmittance depends on light’s propagating modes, and their interference

with surface plasmon polariton modes [20]. Many reports have focused on simulating

transmittance for various MNG designs, and they have typically found at most 6% absolute

increases in broadband transmittance over the aperture value [28, 81, 82, 84, 85]. Results

are highly dependent on specific simulation parameters, such as substrate refractive indices

[81, 84, 86], metal refractive indices, aperture volume-filling refractive indices [84],

wavelengths, precise geometry, light polarization, and light incidence angle [20, 84]. The

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
transmittance of a Cu MNG with w and hMNG of 50 nm and Taper of 0.9 was simulated with

air and CdS substrates, different wavelengths, and with refractive indices from different

sources (Table 1). The present results were in line with previous reports: simulated

transmittance was 0.860 to 0.947, depending on the choice of substrate, wavelength, and

indices. Simulation parameters should therefore be tailored to a given MNG fabrication

process, and even then, the simulation will likely deviate from experimental transmittance

due to substrate roughness, metal roughness, imperfect MNG shapes, and deviations from

literature refractive indices. The aperture transmittance should be used during MNG

design, and after optimization, simulated transmittance and other smaller effects should be

modeled.

Table 1. Simulated transmittance of Cu grating with aperture transmittance of 0.9,

minimum wire width of 50 nm, and MNG height of 50 nm using different literature

refractive indices and wavelengths. The CdS substrate was 100 nm thick.

Tsim
Reference(s) Substrate
280-1,360 nm 400-1,220 nm 550 nm
[87] Air 0.922 0.937 0.947
[88] Air 0.919 0.935 0.935
[89] Air 0.927 0.940 0.935
[89, 90] CdS - 0.928 0.860

3.4. MNG resistance (RMNG)

For MNG resistance calculations, it was assumed that conduction occurred in one

dimension, along the most favorable direction for a given shape. Some reports on MNGs

[21, 73, 83, 84, 91, 92] have used percolation conduction models [93] to predict sheet

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
resistance. Percolation conduction can be operative in solution-synthesized Ag nanowires

[1]. However, many MNG fabrication processes form robust wires, junctions and networks,

making operation in the percolation conduction regime unlikely [94]. MNG sheet

resistances can still deviate from simple geometric filling fraction estimates, but the

magnitude of these deviations is small, relative to percolation [94]. Conduction along less

favorable directions can also be modeled [94] for specific MNG processes. Experimental

studies suggest that Cu films and individual wires down to thicknesses and widths of 50

nm exhibit at most a twofold increase in resistivity, relative to the bulk metal [67, 95-101].

These deviations from ideal grid sheet resistance and individual wire resistivity should be

carefully modelled for specific processes; they are not used in the present general model.

The grating, 4x and 2x rectangle, triangle, and square MNG sheet resistances were

given by a simple expression (for the grating, 𝑅𝑠ℎ,𝑀𝑁𝐺 = 𝜌𝑚𝑒𝑡𝑎𝑙 ∙ 𝑤 ⁄(ℎ𝑀𝑁𝐺 ∙ (𝑠 + 𝑤)))

[20]. The hexagon and circle had no linear conduction pathways, so their sheet resistances

were simulated [28]. For constant hMNG of 50 nm, Taper is shown against Rsh,MNG for each

shape in Fig. 6. The curves in Fig. 6 depend on Taper, but not w. Grating and rectangle

shapes were again superior. A grating with Taper of 0.9 had an Rsh,MNG of 3.4 Ω/sq. At low

Taper, the circle performed better than the other shapes, because its wires were significantly

wider than w throughout most of the unit cell. The Taper-Rsh,MNG curve was recalculated for

the grating at three hMNG, and compared with literature Al:ZnO and ITO data in Fig. 7. The

data demonstrate that hMNG can be increased until MNG sheet resistance is negligible for

many designs, making it less likely that RMNG will dominate, and emphasizing the roles of

Rneighb and Taper on performance. RMNG strongly depended on L, the cell length, or distance

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
current must flow to reach a macro-grid finger, busbar, or terminal (𝑅𝑀𝑁𝐺 =

𝑅𝑠ℎ,𝑀𝑁𝐺 ∙ 𝐿2 ⁄3) [102]. RMNG are included in output power calculations in the next section.

Fig. 6. Aperture transmittance versus MNG sheet resistance at constant MNG height of 50

nm for the grating (purple), 4x rectangle (blue), 2x rectangle (green), triangle (orange),

hexagon (red), square (gray), and circle (black). All calculations used an analytical model,

except for the hexagon and circle, which were simulated.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
Fig. 7. Aperture transmittance (purple) versus grating MNG sheet resistance at MNG

heights of 50 nm (solid), 100 nm (dashed), and 200 nm (dotted). Literature transmittance

versus sheet resistance for relatively high quality Al:ZnO (green dashed [103]), low quality

Al:ZnO (green solid [104]), high quality ITO (red dotted [17]), mid-range quality ITO (red

dashed [105]), and low quality ITO (red solid [105]).

3.5. Output power (Pout) and figure of merit (FFgrid)

Transparent conduction of a metal grid can be limited by Rneighb, Taper, or RMNG.

Transmittance determines photocurrent, while the photocurrent and series resistances

reduce the photovoltage through Ohmic dissipation. The MNG design can therefore be

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
expressed in terms of current density loss and voltage loss for a solar cell that would

otherwise operate at some J and V, taken here to be near the world record JSC and VOC for

a given technology.

𝐽𝑜𝑢𝑡 = 𝐽𝑆𝐶 − 𝐽𝑙𝑜𝑠𝑠 = 𝐽𝑆𝐶 ∙ 𝑇𝑎𝑝𝑒𝑟 (8)

𝑉𝑜𝑢𝑡 = 𝑉𝑂𝐶 − 𝑉𝑙𝑜𝑠𝑠 = 𝑉𝑂𝐶 − (𝑅𝑛𝑒𝑖𝑔ℎ𝑏 + 𝑅𝑀𝑁𝐺 ) ∙ 𝐽𝑆𝐶 (9)

For the grating case, in terms of Taper and w:

2
𝜌𝑛𝑒𝑖𝑔ℎ𝑏 ∙𝑇𝑎𝑝𝑒𝑟 ∙𝑤 2 𝜌𝑚𝑒𝑡𝑎𝑙 ∙(1−𝑇𝑎𝑝𝑒𝑟 )∙𝐿2
𝑉𝑜𝑢𝑡 = 𝑉𝑂𝐶 − ( 2 + ) ∙ 𝐽𝑆𝐶 (10)
12∙ℎ𝑛𝑒𝑖𝑔ℎ𝑏 ∙(1−𝑇𝑎𝑝𝑒𝑟 ) 3∙ℎ𝑀𝑁𝐺

Choosing values for JSC, VOC, ρneighb, ρmetal, hneighb, hmetal, L, and w then allows Taper to be

independently varied to calculate Jout and Vout. An example Jout-Vout curve is shown for the

grating in Fig. 8 for w of 50 nm and L of 1 cm. The product of Jout and Vout was then plotted

against Vout to yield a familiar Pout curve (right axis in Fig. 8). The Pout maximum occurs at

the optimal Taper for the chosen variables, which can be used to calculate optimal grid

dimensions. The optimal transmittance (Taper,opt) achieves the optimal current density-

voltage tradeoff (Jout,opt and Vout,opt, respectively). To generalize maximum Pout and

compare among different technologies, it was divided by JSC and VOC to yield a grid fill

factor:

2
𝐽𝑜𝑢𝑡,𝑜𝑝𝑡 ∙𝑉𝑜𝑢𝑡,𝑜𝑝𝑡 𝜌𝑛𝑒𝑖𝑔ℎ𝑏 ∙𝑇𝑎𝑝𝑒𝑟,𝑜𝑝𝑡 ∙𝑤 2 𝜌𝑚𝑒𝑡𝑎𝑙 ∙(1−𝑇𝑎𝑝𝑒𝑟,𝑜𝑝𝑡 )∙𝐿2
𝐹𝐹𝑔𝑟𝑖𝑑 = = 𝑇𝑎𝑝𝑒𝑟,𝑜𝑝𝑡 − ( 2 + )∙
𝐽𝑆𝐶 ∙𝑉𝑂𝐶 12∙ℎ𝑛𝑒𝑖𝑔ℎ𝑏 ∙(1−𝑇𝑎𝑝𝑒𝑟,𝑜𝑝𝑡 ) 3∙ℎ𝑀𝑁𝐺

𝐽𝑆𝐶 ∙𝑇𝑎𝑝𝑒𝑟,𝑜𝑝𝑡
(11)
𝑉𝑂𝐶

This equation is for a grating, where more complicated expressions can be derived for other

shapes by following the analysis in this work. FFgrid is shown for each shape as a function

of w in Fig. 9. At w of 50 nm, the shapes performed almost identically, making shape

unimportant in the very small w regime. As w increased, the performance spread among

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
shapes was magnified, making the high aspect ratio shapes like the grating and rectangles

more desirable. The circle performed the worst, so nanosphere lithography with particles

that pack more closely than spheres could lead to better tessellated grid apertures, and better

MNG performance.

40
Jout,opt 30
High Taper Taper,opt
30

Pout (mW/cm2)
Jout (mA/cm2)

L = 1 cm 20
20 w = 50 nm

10
10
Low Taper
0 0
0 200 400 600
Vout,opt
Vout (mV)

Fig. 8. Output current density (black; left axis) and output power density (blue; right axis)

versus output voltage for a grating with constant cell length of 1 cm and minimum wire

width of 50 nm. Low (left), optimal (red x), and high (right) aperture transmittances were

used to construct the curves, and the dashed lines show the maximum power, current

density, and voltage.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
Fig. 9. Grid fill factor versus minimum wire widths for the grating (purple), 4x rectangle

(blue), 2x rectangle (green), triangle (orange), hexagon (red), square (gray), and circle

(black). MNG height was 50 nm and cell length was 0.01 cm.

In Fig. 8 and 9, L was assumed to be relatively large and small, respectively. The

effect of L on FFgrid is demonstrated in Fig. 10, where L was varied for grating MNGs with

different w and hMNG, as compared to literature TCO curves (where the TCO series

resistance is 𝑅𝑇𝐶𝑂 = 𝑅𝑠ℎ,𝑇𝐶𝑂 ∙ 𝐿2 ⁄3 [102]). Rneighb dominated to the left of Fig. 10, causing

lines with equal w to converge, while RMNG dominated to the right of Fig. 10, making lines

with equal hMNG converge. For w and hMNG of 50 nm, Rsh,MNG may increase by 2x [67, 95-

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
101], or even more [94]. This increase can partially be offset with thicker MNGs, but MNG

fabrication processes have hMNG limits, so sheet resistance deviations and maximum MNG

thicknesses will be process-specific. For the range typically used in thin film PV modules

(L of 0.1 to 1.0 cm [102, 106]), the idealized MNGs were predicted to generally have better

performance than TCOs. The different slopes for MNGs and TCOs demonstrate why,

historically, metal grids have been used to conduct large currents over long distances.

Fig. 10. Grid fill factor versus cell length for grating MNGs with 50 nm wire width and 50

nm height (purple dashed) or 100 nm height (purple solid) or 800 nm wire width and 50

nm height (blue dashed) or 100 nm height (blue solid). Literature TCO fill factor for

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
Al:ZnO (green dashed [103]), relatively high quality ITO (red dotted [17]), mid-range

quality ITO (red dashed [105]), and low quality ITO (red solid [105]).

Several figures of merit have been developed and used for transparent conducting

films [17, 106-108], and they have sometimes been used to assess the merit of metal grids

[23-25, 29, 30, 34, 70, 77, 81, 109-114]. Those figures would only be informative if RMNG

dominated Rneighb, a case which may not occur in solar cell applications. Other figures of

merit have been proposed for general grid performance [94, 115]. For metal grid

performance in any solar cell, the FFgrid should be used as a figure of merit to correctly

balance photocurrent and photovoltage losses. The FFgrid derived here reduces to a figure

of merit very recently proposed for the case of a grating with 𝑤 ≪ 𝑤 + 𝑠 [116], which only

differs from Eq. (11) for very poorly performing grids.

FFgrid for a grating was calculated as a function of Rsh,neighb and w, and the contours

are shown in Fig. 11. Rsh,neighb ranges were estimated from the literature for various high

efficiency technologies (shown against record JSC in Fig. 12), and are shown to the right of

Fig. 11. The w ranges that have been achieved experimentally by various printing and

lithography techniques are listed in Table 2 and summarized at the top of Fig. 11. This plot

may be used to estimate PV performance for a given absorber and MNG process.

Homojunction silicon and GaAs have the lowest Rsh,neighb, which has historically allowed

metal macroscopic grids (w of 105 to 106 nm) to be employed in those devices with minimal

losses. However, an order of magnitude decrease in w would increase FFgrid by 5% relative.

Additionally, since Rsh,neighb is proportional to w2 (mentioned above and observable in Fig.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
11), technologies with smaller w could permit Si and GaAs solar cell designs with relaxed

Rsh,neighb constraints.

Drying cracked film


Electrospun nanofiber Laser print
EHD jet print Inkjet
Nanosphere Capillary flow
Screen print;
Nanoimprint assembly
flexographic

109 Grid fill Hybrid


perovskite
108 factor
Rsh,neighb (/sq)

CdTe front;
107 CIGS front;
106 CdTe back

105 HIT

104 CIGS back

103 GaAs back

102 Si;
GaAs front
101
102 103 104 105
w (nm)

Fig. 11. Grid fill factor contours versus neighbor sheet resistance and minimum wire width

for a grating MNG with negligible cell length. Thicker 0.999, 0.99, and 0.9 contour lines

are labeled. Literature neighbor sheet resistance estimates for high efficiency PV

technologies are summarized on the right (detailed in Fig. 12). Literature ranges for

minimum wire widths of metal grid fabrication process are summarized at the top (detailed

in Table 2). This study used CdS with Rsh,neighb of 7 · 107 Ω/sq.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
Fig. 12. Neighbor sheet resistances versus world record short-circuit current density for

homojunction Si front [18, 117-121], HIT front or back [122-127], GaAs front [119, 121,

128-133], GaAs back [119, 121, 128-131, 134], CdTe front [52-56, 58, 60, 62, 121], CdTe

back [56, 58, 60, 62, 121], CIGS front [52-55, 57-59, 61-63, 65], CIGS back [58, 59, 61-

64], and hybrid perovskite front or back [121, 135-144].

Table 2. Experimental reports of metal nano- and micro-grids with minimum wire widths

of less than 50,000 nm.

Reference(s) Process classification Min. w (nm) Max. w (nm)


Capillary force lith.;
[68, 145] 15 30
secondary sputtering
[26, 146] Mold transfer 40 20,000

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
[24, 36, 76, 147,
E-beam lith. 45 5,000
148]
Super-aligned carbon
[149] 50 50
nanotube lith.
[30, 34, 108, 150-
Nanoimprint lith. 55 1,000
155]
[22, 68, 156, 157] Nanosphere lith. 56 378
[91] Colloid lith. 60 210
Focused ion beam
[27] 70 70
milling
Electrohydrodynamic
[72, 109] 80 7,500
jet print
[77, 158] Natural pattern lith. 80 110,000
[75, 159] Reaction front lith. 90 90
[31, 70, 111-114,
Photolith. 200 180,000
160-163]
[25, 29, 32, 76, Drying cracked film
300 150,000
164-166] lith.
[74, 148, 167] Interference lith. 340 3,000
[27, 68, 115, 168, Electrospun
380 1,500
169] nanofiber lith.
Capillary flow self-
[170, 171] 2,000 20,400
assembly
[73] Laser ablation 2,000 2,000
Patterned capillary
[172] 4,000 25,000
drying
[23, 33, 173] Mold filling with ink 4,500 40,000
Evaporation-induced
[37] 5,000 6,000
self-assembly
Inkjet capillary
[174, 175] 5,000 10,000
drying
Salt crystallization
[176] 10,000 10,000
lith.
[177] Roll-offset print 20,000 80,000
[110] Gravure print 22,000 39,000

The technologies that presently employ TCOs (HIT, CIGS, CdTe, and hybrid

provskites) in laboratory solar cells and modules have far greater Rsh,neighb, and demand

smaller w for MNGs to be effective. These demands on w approach the limits of the

presently employed general model. Rsh,neighb estimates typically span several orders of

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
magnitude, reducing the certainty of theoretical results based on those figures. The largest

deviation from the present idealized grating model should be for w smaller than 100 nm.

On that scale, metal resistivities can deviate from bulk values, experimental MNGs can

deviate from idealized periodic geometric structures, and transmittance can deviate from

aperture transmittance. The exact Rsh,neighb used in this work (7 · 107 Ω/sq) is near the limit

of MNG benefits. Nevertheless, gratings with w of 50 – 100 nm were predicted to have

excellent grid fill factors for less conservative Rsh,neighb estimates (near 107 Ω/sq). In spite

of this remarkable potential, MNGs have only been employed in relatively low efficiency

solar cells—literature results are listed in Table 3. The guide for designing high efficiency

MNG solar cells in Fig. 11 should help the record MNG efficiency advance well past its

present value of 7.0% [36, 37].

Table 3. Experimental reports of solar cells with metal grid contacts and minimum wire

widths of less than 106 nm. Organic absorbers were used unless noted.

Max. 1
hMNG sun
Reference Process w (nm) Shape Metal
(nm) efficiency
(%)
Capillary
force lith.;
[68] 15 300 Grating Au 1.9
secondary
sputtering
Nanoimprint
[178] 55-130 30 Square Ag 2.8
lith.
Nanoimprint Hierarchical
[150] 70-400 40 Cu 2.1
lith. grating
Nanoimprint Hierarchical Ag; Au;
[151] 70-400 40 2.1
lith. grating Cu
[36] E-beam lith. 100 50 Square Au; Pd 7.0*
Drying
300- Random
[29] cracked film 20-60 Ag; Au 2.0
5,000 wire
lith.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
Capillary
4,500- Al; Au;
[170] flow 90 Square 4.1
20,400 Cr
assembly
Evaporation-
5,000- Random
[37] induced self- 2,000 Ag 7.0
6,000 wire
assembly
Microcontact 5,000-
[146] 30 Square Ag 3.2
print 20,000
Square;
triangle;
tetrakis
[70] Photolith. 5,000 50 Ag 1.6
square
(right
triangle)
[160] Photolith. 6,000 1,000 Grating Ti 6.2**
Capillary 20,000-
[173] 100-160 Grating Ag 1.0
mold filling 40,000
*Crystalline Si absorber

**Dye-sensitized absorber

4. Conclusions

A general procedure for quantifying a metal grid’s effect on PV performance was

outlined, with an emphasis on replacing TCOs with metal nano-grids in high efficiency

CdS-buffered solar cells. Neighbor series resistance, grid transmittance, and grid series

resistance determined this performance, and estimates for each were considered separately.

Rneighb could be estimated using Wyeth’s model, which is analytical for simple geometries.

Rneighb were also simulated with fewer approximations, but differences between analytical

and simulated results were unimportant for most MNG designs. Simulated transmittance

was ~3% absolute greater than aperture transmittance for w of 50 nm, and relatively

unimportant for MNG design. Taper-MNG sheet resistance trade-offs were shown to easily

outperform TCOs, although deviations from idealized sheet resistance were not included

due to their process-specificity. MNG thickness increases can partially offset those effects

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
in many designs, in agreement with experimental reports [23-28]. Losses in photocurrent

and photovoltage were calculated from Rneighb, Taper, and RMNG. These were used to

construct familiar J-V curves that illustrated the trade-off between optical and electrical

losses. The maximum power point was at the optimal Taper for a given w. This optimum

was normalized to JSC and VOC to calculate a grid fill factor. The grid fill factor should be

used as a figure of merit for metal grids in solar cells, where other figures of merit were

derived for films, and are uninformative for inhomogeneous grids. MNG performance

strongly depended on neighbor series resistance, making transmittance-RMNG plots in the

literature less informative for TCO comparison. The dependence of grid fill factor on

Rsh,neighb and w was used to map out the performance gains that are possible as w is

decreased in every high efficiency PV technology.

Acknowledgements

The author declares that he has no conflict of interest. This research did not receive any

specific grant from funding agencies in the public, commercial, or not-for-profit sectors.

References

[1] S. Ye, A.R. Rathmell, Z. Chen, I.E. Stewart, B.J. Wiley, Metal Nanowire Networks:
The Next Generation of Transparent Conductors, Adv. Mater., 26 (2014) 6670-6687.
[2] Z. Chen, B. Cotterell, W. Wang, E. Guenther, S.-J. Chua, A mechanical assessment of
flexible optoelectronic devices, Thin Solid Films, 394 (2001) 201-205.
[3] S.M. Goodnick, J.F. Wager, C.W. Wilmsen, Thermal degradation of indium‐tin‐
oxide/p‐silicon solar cells, J. Appl. Phys., 51 (1980) 527-531.
[4] H.-K. Kim, D.-G. Kim, K.-S. Lee, M.-S. Huh, S.H. Jeong, K.I. Kim, T.-Y. Seong,
Plasma damage-free sputtering of indium tin oxide cathode layers for top-emitting
organic light-emitting diodes, Appl. Phys. Lett., 86 (2005) 183503.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
[5] H. Liu, V. Avrutin, N. Izyumskaya, Ü. Özgür, H. Morkoç, Transparent conducting
oxides for electrode applications in light emitting and absorbing devices, Superlattices
Microstruct., 48 (2010) 458-484.
[6] J.I. Kim, W. Lee, T. Hwang, J. Kim, S.-Y. Lee, S. Kang, H. Choi, S. Hong, H.H.
Park, T. Moon, B. Park, Quantitative analyses of damp-heat-induced degradation in
transparent conducting oxides, Sol. Energy Mater. Sol. Cells, 122 (2014) 282-286.
[7] H.M. Lemire, K.A. Peterson, S. Sprawls, K. Singer, I.T. Martin, R.H. French,
Degradation of transparent conductive oxides: mechanistic insights across configurations
and exposures, in, 2013, pp. 882502-882502-882508.
[8] F.J. Pern, R. Noufi, Stability of CIGS solar cells and component materials evaluated
by a step-stress accelerated degradation test method, in, 2012, pp. 84720J-84720J-84714.
[9] H. Khaligh, I. Goldthorpe, Failure of silver nanowire transparent electrodes under
current flow, Nanoscale Res. Lett., 8 (2013) 1-6.
[10] J.L. Elechiguerra, L. Larios-Lopez, C. Liu, D. Garcia-Gutierrez, A. Camacho-
Bragado, M.J. Yacaman, Corrosion at the Nanoscale: The Case of Silver Nanowires and
Nanoparticles, Chem. Mater., 17 (2005) 6042-6052.
[11] B. Feldman, Simulating Causes of Conductivity Degradation in Nanoscale Metal
Wires, in: Physics, University of Washington, Seattle, WA, 2008, pp. 103.
[12] Q. Jiang, S. Zhang, M. Zhao, Size-dependent melting point of noble metals,
Materials Chemistry and Physics, 82 (2003) 225-227.
[13] J. Jiu, J. Wang, T. Sugahara, S. Nagao, M. Nogi, H. Koga, K. Suganuma, M. Hara,
E. Nakazawa, H. Uchida, The effect of light and humidity on the stability of silver
nanowire transparent electrodes, RSC Adv., 5 (2015) 27657-27664.
[14] Y.C.G. Kwan, Q.L. Le, C.H.A. Huan, Time to failure modeling of silver nanowire
transparent conducting electrodes and effects of a reduced graphene oxide over layer, Sol.
Energy Mater. Sol. Cells, 144 (2016) 102-108.
[15] C. Mayousse, C. Celle, A. Fraczkiewicz, J.-P. Simonato, Stability of silver nanowire
based electrodes under environmental and electrical stresses, Nanoscale, 7 (2015) 2107-
2115.
[16] M. Zhang, M.Y. Efremov, F. Schiettekatte, E.A. Olson, A.T. Kwan, S.L. Lai, T.
Wisleder, J.E. Greene, L.H. Allen, Size-dependent melting point depression of
nanostructures: Nanocalorimetric measurements, Physical Review B, 62 (2000) 10548-
10557.
[17] T.M. Barnes, M.O. Reese, J.D. Bergeson, B.A. Larsen, J.L. Blackburn, M.C. Beard,
J. Bult, J. van de Lagemaat, Comparing the Fundamental Physics and Device
Performance of Transparent, Conductive Nanostructured Networks with Conventional
Transparent Conducting Oxides, Advanced Energy Materials, 2 (2012) 353-360.
[18] M.A. Green, Solar Cells: Operating Principles, technology, and system applications,
Prentice-Hall, New Jersey, 1982.
[19] M. Ju, Y.-J. Lee, J. Lee, B. Kim, K. Ryu, K. Choi, K. Song, K. Lee, C. Han, Y. Jo, J.
Yi, Double screen printed metallization of crystalline silicon solar cells as low as 30 μm
metal line width for mass production, Sol. Energy Mater. Sol. Cells, 100 (2012) 204-208.
[20] P.B. Catrysse, S. Fan, Nanopatterned Metallic Films for Use As Transparent
Conductive Electrodes in Optoelectronic Devices, Nano Lett., 10 (2010) 2944-2949.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
[21] A. Hubarevich, M. Marus, W. Fan, A. Smirnov, X.W. Sun, H. Wang, Theoretical
comparison of optical and electronic properties of uniformly and randomly arranged
nano-porous ultra-thin layers, Opt. Express, 23 (2015) 17860-17865.
[22] W. Wu, N.G. Tassi, A broadband plasmonic enhanced transparent conductor,
Nanoscale, 6 (2014) 7811-7816.
[23] Y.S. Oh, D.Y. Choi, H.J. Sung, Direct imprinting of thermally reduced silver
nanoparticles via deformation-driven ink injection for high-performance, flexible metal
grid embedded transparent conductors, RSC Adv., 5 (2015) 64661-64668.
[24] J. van de Groep, P. Spinelli, A. Polman, Transparent Conducting Silver Nanowire
Networks, Nano Lett., 12 (2012) 3138-3144.
[25] B. Han, K. Pei, Y. Huang, X. Zhang, Q. Rong, Q. Lin, Y. Guo, T. Sun, C. Guo, D.
Carnahan, M. Giersig, Y. Wang, J. Gao, Z. Ren, K. Kempa, Uniform Self-Forming
Metallic Network as a High-Performance Transparent Conductive Electrode, Adv.
Mater., 26 (2014) 873-877.
[26] P. Kuang, J.-M. Park, W. Leung, R.C. Mahadevapuram, K.S. Nalwa, T.-G. Kim, S.
Chaudhary, K.-M. Ho, K. Constant, A New Architecture for Transparent Electrodes:
Relieving the Trade-Off Between Electrical Conductivity and Optical Transmittance,
Adv. Mater., 23 (2011) 2469-2473.
[27] H. Wu, D. Kong, Z. Ruan, P.-C. Hsu, S. Wang, Z. Yu, T.J. Carney, L. Hu, S. Fan, Y.
Cui, A transparent electrode based on a metal nanotrough network, Nat Nano, 8 (2013)
421-425.
[28] T. Gao, B. Wang, B. Ding, J.-k. Lee, P.W. Leu, Uniform and Ordered Copper
Nanomeshes by Microsphere Lithography for Transparent Electrodes, Nano Lett., 14
(2014) 2105-2110.
[29] C. Hunger, K.D.M. Rao, R. Gupta, C.R. Singh, G.U. Kulkarni, M. Thelakkat,
Transparent Metal Network with Low Haze and High Figure of Merit applied to Front
and Back Electrodes in Semitransparent ITO-free Polymer Solar Cells, Energy
Technology, 3 (2015) 638-645.
[30] J. van de Groep, D. Gupta, M.A. Verschuuren, M. M. Wienk, R.A.J. Janssen, A.
Polman, Large-area soft-imprinted nanowire networks as light trapping transparent
conductors, Sci. Rep., 5 (2015).
[31] D.-J. Kim, H.-J. Kim, K.-W. Seo, K.-H. Kim, T.-W. Kim, H.-K. Kim, Indium-free,
highly transparent, flexible Cu2O/Cu/Cu2O mesh electrodes for flexible touch screen
panels, Sci. Rep., 5 (2015) 16838.
[32] S. Kiruthika, R. Gupta, A. Anand, A. Kumar, G.U. Kulkarni, Fabrication of
Oxidation-Resistant Metal Wire Network-Based Transparent Electrodes by a Spray-Roll
Coating Process, ACS Appl. Mater. Interfaces, 7 (2015) 27215-27222.
[33] J.-S. Yu, I. Kim, J.-S. Kim, J. Jo, T.T. Larsen-Olsen, R.R. Sondergaard, M. Hosel,
D. Angmo, M. Jorgensen, F.C. Krebs, Silver front electrode grids for ITO-free all printed
polymer solar cells with embedded and raised topographies, prepared by thermal imprint,
flexographic and inkjet roll-to-roll processes, Nanoscale, 4 (2012) 6032-6040.
[34] H.-J. Kim, S.-H. Lee, J. Lee, E.-S. Lee, J.-H. Choi, J.-H. Jung, J.-Y. Jung, D.-G.
Choi*, High-Durable AgNi Nanomesh Film for a Transparent Conducting Electrode,
Small, 10 (2014) 3767-3774.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
[35] G. Hautier, A. Miglio, G. Ceder, G.-M. Rignanese, X. Gonze, Identification and
design principles of low hole effective mass p-type transparent conducting oxides, Nat.
Commun., 4 (2013) 2292-2297.
[36] S.Z. Oener, J. van de Groep, B. Macco, P.C.P. Bronsveld, W.M.M. Kessels, A.
Polman, E.C. Garnett, Metal–Insulator–Semiconductor Nanowire Network Solar Cells,
Nano Lett., 16 (2016) 3689-3695.
[37] K.-W. Seo, Y.-J. Noh, S.-I. Na, H.-K. Kim, Random mesh-like Ag networks
prepared via self-assembled Ag nanoparticles for ITO-free flexible organic solar cells,
Sol. Energy Mater. Sol. Cells, 155 (2016) 51-59.
[38] C.R. Fang, J.R. Hauser, A two dimensional analysis of sheet resistance and contact
resistance effects in solar cells, in: 13th Photovoltaic Specialists Conference, 1978, pp.
1306-1311.
[39] A. de Vos, The distributed series resistance problem in solar cells, Solar Cells, 12
(1984) 311-327.
[40] I.L. Eisgruber, J.R. Sites, Effect of thin film module geometry on solar cell current-
voltage analysis, in: Photovoltaic Energy Conversion, 1994., Conference Record of the
Twenty Fourth. IEEE Photovoltaic Specialists Conference - 1994, 1994 IEEE First World
Conference on, 1994, pp. 271-274.
[41] R.J. Handy, Theoretical analysis of the series resistance of a solar cell, Solid-State
Electron., 10 (1967) 765-775.
[42] M.W. Denhoff, N. Drolet, The effect of the front contact sheet resistance on solar
cell performance, Sol. Energy Mater. Sol. Cells, 93 (2009) 1499-1506.
[43] L.D. Nielsen, Distributed series resistance effects in solar cells, IEEE Trans.
Electron Devices, 29 (1982) 821-827.
[44] T. Miyadera, H. Ogo, T. Taima, T. Yamanari, Y. Yoshida, Analytical model for the
design principle of large-area solar cells, Sol. Energy Mater. Sol. Cells, 97 (2012) 127-
131.
[45] A.W. Haas, J.R. Wilcox, J.L. Gray, R.J. Schwartz, Numerical modeling of loss
mechanisms resulting from the distributed emitter effect in concentrator solar cells, in:
Photovoltaic Specialists Conference (PVSC), 2009 34th IEEE, 2009, pp. 002244-002249.
[46] Y. Yang, G. Xu, K. Zhang, X. Zhang, H. Shen, P.P. Altermatt, P.J. Verlinden, Z.
Feng, Analysis of series resistance of industrial crystalline silicon solar cells by
numerical simulation and analytical modeling, in: 28th European Photovoltaic Solar
Energy Conference and Exhibition, 2013, pp. 1558-1561.
[47] H. Mäckel, G. Micard, K. Varner, Analytical models for the series resistance of
selective emitters in silicon solar cells including the effect of busbars, Prog. Photovolt:
Res. Appl., 23 (2015) 135-149.
[48] A.V. Sachenko, A.P. Gorban, On the collection of photocurrent in solar cells with a
contact grid, Semiconductor Physics, Quantum Electronics and Optoelectronics, 2 (1999)
42-44.
[49] A.K. Aboul Seoud, H. Amer, Calculation of the resistance of the diffused top layer
in a photovoltaic cell, Solar Cells, 19 (1986) 1-7.
[50] V.K. Jain, A.P. Kulshreshtha, Indium-Tin-Oxide transparent conducting coatings on
silicon solar cells and their “figure of merit”, Solar Energy Materials, 4 (1981) 151-158.
[51] N.C. Wyeth, Sheet resistance component of series resistance in a solar cell as a
function of grid geometry, Solid-State Electron., 20 (1977) 629-634.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
[52] T.L. Chu, S.S. Chu, N. Schultz, C. Wang, C.Q. Wu, Solution‐Grown Cadmium
Sulfide Films for Photovoltaic Devices, Journal of The Electrochemical Society, 139
(1992) 2443-2446.
[53] J.M. Doña, J. Herrero, Chemical Bath Deposition of CdS Thin Films:
Electrochemical In Situ Kinetic Studies, Journal of The Electrochemical Society, 139
(1992) 2810-2814.
[54] H. Khallaf, I.O. Oladeji, G. Chai, L. Chow, Characterization of CdS thin films
grown by chemical bath deposition using four different cadmium sources, Thin Solid
Films, 516 (2008) 7306-7312.
[55] M. Rami, E. Benamar, M. Fahoume, F. Chraibi, A. Ennaoui, Effect of the cadmium
ion source on the structural and optical properties of chemical bath deposited CdS thin
films, Solid State Sci., 1 (1999) 179-188.
[56] N. Amin, K. Sopian, M. Konagai, Numerical modeling of CdS/CdTe and
CdS/CdTe/ZnTe solar cells as a function of CdTe thickness, Sol. Energy Mater. Sol.
Cells, 91 (2007) 1202-1208.
[57] S.J. Fonash, Solar Cell Device Physics, Academic Press, London, 2012.
[58] M. Gloeckler, A.L. Fahrenbruch, J.R. Sites, Numerical modeling of CIGS and CdTe
solar cells: setting the baseline, in: Photovoltaic Energy Conversion, 2003. Proceedings
of 3rd World Conference on, 2003, pp. 491-494.
[59] C.-H. Huang, Effects of junction parameters on Cu(In,Ga)Se2 solar cells, J. Phys.
Chem. Solids, 69 (2008) 779-783.
[60] M.A. Matin, M. Mannir Aliyu, A.H. Quadery, N. Amin, Prospects of novel front and
back contacts for high efficiency cadmium telluride thin film solar cells from numerical
analysis, Sol. Energy Mater. Sol. Cells, 94 (2010) 1496-1500.
[61] J.R. Ray, C.J. Panchal, M.S. Desai, U.B. Trivedi, Simulation of CIGS thin film solar
cells using AMPS-1D, Journal of Nano- and Electronic Physics, 3 (2011) 747-754.
[62] R. Scheer, H.-W. Schock, Chalcogenide photovoltaics: Physics, technologies, and
thin film devices, Wiley-VCH, Weinheim, Germany, 2011.
[63] J. Song, Development, Characterization, and Modeling of CuGaSe2/Cu(In,Ga)Se2
thin-film tandem solar cells, in: Electrical and Computer Engineering, University of
Florida, Gainesville, FL, 2006, pp. 142.
[64] R. Kamada, T. Yagioka, S. Adachi, A. Handa, K.F. Tai, T. Kato, H. Sugimoto, New
world record Cu(In,Ga)(Se,S)2 thin film solar cell efficiency beyond 22%, in:
Photovoltaic Specialist Conference (PVSC), 2016 IEEE 43rd, Portland, OR, 2016, pp. in
press.
[65] P. Jackson, R. Wuerz, D. Hariskos, E. Lotter, W. Witte, M. Powalla, Effects of
heavy alkali elements in Cu(In,Ga)Se2 solar cells with efficiencies up to 22.6%, physica
status solidi (RRL) – Rapid Research Letters, (2016) in press.
[66] C. Durkan, M.E. Welland, Size effects in the electrical resistivity of polycrystalline
nanowires, Physical Review B, 61 (2000) 14215-14218.
[67] S.M. Rossnagel, T.S. Kuan, Alteration of Cu conductivity in the size effect regime,
Journal of Vacuum Science & Technology B, 22 (2004) 240-247.
[68] C.J. An, S. Jang, K.M. Kang, S.J. Kim, M.L. Jin, H.-T. Jung, A combined graphene
and periodic Au nanograte structure: Fundamentals and application as a flexible
transparent conducting film in a flexible organic photovoltaic cell, Carbon, 103 (2016)
488-496.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
[69] S. Deb, B. Ghosh, Series resistance and optimum grid design for a thin film solar
cell of rectangular shape, Solar Cells, 13 (1984) 145-162.
[70] J.W. Lim, Y.T. Lee, R. Pandey, T.-H. Yoo, B.-I. Sang, B.-K. Ju, D.K. Hwang, W.K.
Choi, Effect of geometric lattice design on optical/electrical properties of transparent
silver grid for organic solar cells, Opt. Express, 22 (2014) 26891-26899.
[71] Y. Galagan, E.W.C. Coenen, S. Sabik, H.H. Gorter, M. Barink, S.C. Veenstra, J.M.
Kroon, R. Andriessen, P.W.M. Blom, Evaluation of ink-jet printed current collecting
grids and busbars for ITO-free organic solar cells, Sol. Energy Mater. Sol. Cells, 104
(2012) 32-38.
[72] J. Schneider, P. Rohner, D. Thureja, M. Schmid, P. Galliker, D. Poulikakos,
Electrohydrodynamic NanoDrip Printing of High Aspect Ratio Metal Grid Transparent
Electrodes, Adv. Funct. Mater., 26 (2016) 833-840.
[73] D. Paeng, J.-H. Yoo, J. Yeo, D. Lee, E. Kim, S.H. Ko, C.P. Grigoropoulos, Low-
Cost Facile Fabrication of Flexible Transparent Copper Electrodes by Nanosecond Laser
Ablation, Adv. Mater., 27 (2015) 2762-2767.
[74] J.W. Menezes, J. Ferreira, M.J.L. Santos, L. Cescato, A.G. Brolo, Large-Area
Fabrication of Periodic Arrays of Nanoholes in Metal Films and Their Application in
Biosensing and Plasmonic-Enhanced Photovoltaics, Adv. Funct. Mater., 20 (2010) 3918-
3924.
[75] B. Kazarkin, A.S. Mohammed, A. Stsiapanau, S. Zhuk, Y. Satskevich, A. Smirnov,
Transparent conductor based on aluminum nanomesh, J. Phys. Conf. Ser., 541 (2014)
012027.
[76] K.D.M. Rao, R. Gupta, G.U. Kulkarni, Fabrication of Large Area, High-
Performance, Transparent Conducting Electrodes Using a Spontaneously Formed Crackle
Network as Template, Advanced Materials Interfaces, 1 (2014) 1400090.
[77] B. Han, Y. Huang, R. Li, Q. Peng, J. Luo, K. Pei, A. Herczynski, K. Kempa, Z. Ren,
J. Gao, Bio-inspired networks for optoelectronic applications, Nat. Commun., 5 (2014)
5674.
[78] K.W. Heizer, T.L. Chu, Solar cell conducting grid structure, Solid-State Electron.,
19 (1976) 471-472.
[79] G.A. Landis, Optimization of tapered busses for solar cell contacts, Sol. Energy, 22
(1979) 401-402.
[80] H.B. Serreze, Optimizing solar cell performance by simultaneous consideration of
grid pattern design and interconnect configuration, in: 13th Photovoltaic Specialists
Conference, 1978, pp. 609-614.
[81] K. Lee, J. Ahn, Substrate effects on the transmittance of 1D metal grid transparent
electrodes, Opt. Express, 22 (2014) 19021-19028.
[82] K. Lee, S.H. Song, J. Ahn, FDTD simulation of transmittance characteristics of one-
dimensional conducting electrodes, Opt. Express, 22 (2014) 6269-6275.
[83] M. Marus, A. Hubarevich, H. Wang, Y. Mukha, A. Smirnov, H. Huang, X.W. Sun,
W. Fan, Towards understanding the difference of optoelectronic performance between
micro- and nanoscale metallic layers, Opt. Mater. Express, 6 (2016) 2655-2661.
[84] Q. Guo Du, K. Sathiyamoorthy, L. Ping Zhang, H. Volkan Demir, C. Hin Kam, X.
Wei Sun, A two-dimensional nanopatterned thin metallic transparent conductor with high
transparency from the ultraviolet to the infrared, Appl. Phys. Lett., 101 (2012) 181112.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
[85] G. Yang, B. Liu, K. Cheng, Z. Du, Modulation of optical transmittance and
conductivity by the period, linewidth and height of Au square mesh electrodes, Opt.
Express, 23 (2015) A62-A70.
[86] D. Qing Guo, R. Hengjiang, W. Lin, B. Ping, P. Ching Eng, S. Xiao Wei, K. Chan
Hin, C.M.d. Sterke, Light absorption mechanism in organic solar cells with hexagonal
lattice nanohole aluminum transparent electrodes, J. Opt., 17 (2015) 085901.
[87] A.D. Rakić, A.B. Djurišić, J.M. Elazar, M.L. Majewski, Optical properties of
metallic films for vertical-cavity optoelectronic devices, Appl. Opt., 37 (1998) 5271-
5283.
[88] CRC Handbook of Chemistry and Physics, 96th Edition, CRC Press/Taylor and
Francis, Boca Raton, FL, 2016.
[89] S. Adachi, The Handbook on Optical Constants of Metals: In Tables and Figures,
World Scientific, Singapore, 2012.
[90] A.M. Salem, Structure, refractive-index dispersion and the optical absorption edge
of chemically deposited ZnxCd(1-x)S thin films, Appl. Phys. A, 74 (2002) 205-211.
[91] T.H. Reilly, R.C. Tenent, T.M. Barnes, K.L. Rowlen, J. van de Lagemaat,
Controlling the Optical Properties of Plasmonic Disordered Nanohole Silver Films, ACS
Nano, 4 (2010) 615-624.
[92] M. Marus, A. Hubarevich, H. Wang, A. Stsiapanau, A. Smirnov, X.W. Sun, W. Fan,
Comparative analysis of opto-electronic performance of aluminium and silver nano-
porous and nano-wired layers, Opt. Express, 23 (2015) 26794-26799.
[93] M. Weber, M.R. Kamal, Estimation of the volume resistivity of electrically
conductive composites, Polym. Compos., 18 (1997) 711-725.
[94] A. Kumar, G.U. Kulkarni, Evaluating conducting network based transparent
electrodes from geometrical considerations, J. Appl. Phys., 119 (2016) 015102.
[95] J.M. Camacho, A.I. Oliva, Morphology and electrical resistivity of metallic
nanostructures, Microelectron. J., 36 (2005) 555-558.
[96] C. Huang, Y. Feng, X. Zhang, J. Li, G. Wang, Electron mean free path model for
rectangular nanowire, nanofilm and nanoparticle, Physica B, 438 (2014) 17-21.
[97] F. Lacy, Developing a theoretical relationship between electrical resistivity,
temperature, and film thickness for conductors, Nanoscale Res. Lett., 6 (2011) 636-614.
[98] H.D. Liu, Y.P. Zhao, G. Ramanath, S.P. Murarka, G.C. Wang, Thickness dependent
electrical resistivity of ultrathin (<40 nm) Cu films, Thin Solid Films, 384 (2001) 151-
156.
[99] H. Marom, J. Mullin, M. Eizenberg, Size-dependent resistivity of nanometric copper
wires, Physical Review B, 74 (2006) 045411.
[100] A.E. Yarimbiyik, H.A. Schafft, R.A. Allen, M.E. Zaghloul, D.L. Blackburn,
Modeling and simulation of resistivity of nanometer scale copper, Microelectron. Reliab.,
46 (2006) 1050-1057.
[101] W. Zhang, S.H. Brongersma, O. Richard, B. Brijs, R. Palmans, L. Froyen, K.
Maex, Influence of the electron mean free path on the resistivity of thin metal films,
Microelectronic Engineering, 76 (2004) 146-152.
[102] G.T. Koishiyev, J.R. Sites, Impact of sheet resistance on 2-D modeling of thin-film
solar cells, Sol. Energy Mater. Sol. Cells, 93 (2009) 350-354.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
[103] J. Wennerberg, J. Kessler, J. Hedström, L. Stolt, B. Karlsson, M. Rönnelid, Thin
film PV modules for low-concentrating systems, Sol. Energy, 69, Supplement 6 (2001)
243-255.
[104] J. Kessler, S. Wiedeman, L. Russell, J. Fogleboch, S. Skibo, R. Arya, D. Carlson,
Front contact optimization for Cu(In,Ga)Se2 (sub)modules, in: Photovoltaic Specialists
Conference, 1996., Conference Record of the Twenty Fifth IEEE, 1996, pp. 885-888.
[105] K. Ellmer, Past achievements and future challenges in the development of optically
transparent electrodes, Nat Photon, 6 (2012) 809-817.
[106] M.W. Rowell, M.D. McGehee, Transparent electrode requirements for thin film
solar cell modules, Energy Environ. Sci., 4 (2011) 131-134.
[107] G. Haacke, New figure of merit for transparent conductors, J. Appl. Phys., 47
(1976) 4086-4089.
[108] B. Sciacca, J. van de Groep, A. Polman, E.C. Garnett, Solution-Grown Silver
Nanowire Ordered Arrays as Transparent Electrodes, Adv. Mater., 28 (2016) 905-909.
[109] J. Yonghee, K. Jihoon, B. Doyoung, Invisible metal-grid transparent electrode
prepared by electrohydrodynamic (EHD) jet printing, J. Phys. D: Appl. Phys., 46 (2013)
155103.
[110] S. Jung, S. Lee, M. Song, D.-G. Kim, D.S. You, J.-K. Kim, C.S. Kim, T.-M. Kim,
K.-H. Kim, J.-J. Kim, J.-W. Kang, Extremely Flexible Transparent Conducting
Electrodes for Organic Devices, Advanced Energy Materials, 4 (2014) n/a-n/a.
[111] A. Khan, S. Lee, T. Jang, Z. Xiong, C. Zhang, J. Tang, L.J. Guo, W.-D. Li, High-
Performance Flexible Transparent Electrode with an Embedded Metal Mesh Fabricated
by Cost-Effective Solution Process, Small, 12 (2016) 3021-3030.
[112] W.-K. Kim, S. Lee, D. Hee Lee, I. Hee Park, J. Seong Bae, T. Woo Lee, J.-Y. Kim,
J. Hun Park, Y. Chan Cho, C. Ryong Cho, S.-Y. Jeong, Cu Mesh for Flexible Transparent
Conductive Electrodes, Sci. Rep., 5 (2015) 10715.
[113] P. Kumar, P.V. Reddy, B. Choudhury, P. Chowdhury, H.C. Barshilia, Transparent
conductive Ta/Al/Ta-grid electrode for optoelectronic and electromagnetic interference
shielding applications, Thin Solid Films, 612 (2016) 350-357.
[114] C.-T. Wang, C.-C. Ting, P.-C. Kao, S.-R. Li, S.-Y. Chu, Investigation of surface
energy, polarity, and electrical and optical characteristics of silver grids deposited via
thermal evaporation method, Appl. Surf. Sci., 360, Part A (2016) 349-352.
[115] C. Bao, J. Yang, H. Gao, F. Li, Y. Yao, B. Yang, G. Fu, X. Zhou, T. Yu, Y. Qin, J.
Liu, Z. Zou, In Situ Fabrication of Highly Conductive Metal Nanowire Networks with
High Transmittance from Deep-Ultraviolet to Near-Infrared, ACS Nano, 9 (2015) 2502-
2509.
[116] D.A. Jacobs, K.R. Catchpole, F.J. Beck, T.P. White, A re-evaluation of transparent
conductor requirements for thin-film solar cells, Journal of Materials Chemistry A, 4
(2016) 4490-4496.
[117] I. Tobias, C.d. Canizo, J. Alonso, Handbook of Photovoltaic Science and
Engineering, 2nd ed., Wiley, West Sussex, UK, 2011.
[118] S.W. Glunz, High-Efficiency Crystalline Silicon Solar Cells, Advances in
OptoElectronics, 2007 (2007) 97370-97315.
[119] J. Nelson, The Physics of Solar Cells, Imperial College Press, London, 2003.
[120] A.L. Fahrenbruch, R.H. Bube, Fundamentals of Solar Cells: Photovoltaic Solar
Energy Conversion, Academic Press, New York, 1983.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
[121] M.A. Green, K. Emery, Y. Hishikawa, W. Warta, E.D. Dunlop, Solar cell
efficiency tables (version 48), Prog. Photovolt: Res. Appl., 24 (2016) 905-913.
[122] N. Hernández-Como, A. Morales-Acevedo, Simulation of hetero-junction silicon
solar cells with AMPS-1D, Sol. Energy Mater. Sol. Cells, 94 (2010) 62-67.
[123] J. Peter Seif, A. Descoeudres, M. Filipič, F. Smole, M. Topič, Z. Charles Holman,
S. De Wolf, C. Ballif, Amorphous silicon oxide window layers for high-efficiency silicon
heterojunction solar cells, J. Appl. Phys., 115 (2014) 024502-024508.
[124] N. Dwivedi, S. Kumar, A. Bisht, K. Patel, S. Sudhakar, Simulation approach for
optimization of device structure and thickness of HIT solar cells to achieve ~27%
efficiency, Sol. Energy, 88 (2013) 31-41.
[125] M. Filipič, Z.C. Holman, F. Smole, S. De Wolf, C. Ballif, M. Topič, Analysis of
lateral transport through the inversion layer in amorphous silicon/crystalline silicon
heterojunction solar cells, J. Appl. Phys., 114 (2013) 074504-074507.
[126] X. Wen, X. Zeng, W. Liao, Q. Lei, S. Yin, An approach for improving the carriers
transport properties of a-Si:H/c-Si heterojunction solar cells with efficiency of more than
27%, Sol. Energy, 96 (2013) 168-176.
[127] K. Masuko, M. Shigematsu, T. Hashiguchi, D. Fujishima, M. Kai, N. Yoshimura,
T. Yamaguchi, Y. Ichihashi, T. Mishima, N. Matsubara, T. Yamanishi, T. Takahama, M.
Taguchi, E. Maruyama, S. Okamoto, Achievement of More Than 25% Conversion
Efficiency With Crystalline Silicon Heterojunction Solar Cell, IEEE J. Photovoltaics, 4
(2014) 1433-1435.
[128] P.D. DeMoulin, M.S. Lundstrom, Projections of GaAs solar-cell performance
limits based on two-dimensional numerical simulation, IEEE Trans. Electron Devices, 36
(1989) 897-905.
[129] K.L. Schulte, W.L. Rance, R.C. Reedy, A.J. Ptak, D.L. Young, T.F. Kuech,
Controlled formation of GaAs pn junctions during hydride vapor phase epitaxy of GaAs,
Journal of Crystal Growth, 352 (2012) 253-257.
[130] K.L. Schulte, J. Simon, A. Roy, R.C. Reedy, D.L. Young, T.F. Kuech, A.J. Ptak,
Computational fluid dynamics-aided analysis of a hydride vapor phase epitaxy reactor,
Journal of Crystal Growth, 434 (2016) 138-147.
[131] J. Simon, K.L. Schulte, D.L. Young, N.M. Haegel, A.J. Ptak, GaAs Solar Cells
Grown by Hydride Vapor-Phase Epitaxy and the Development of GaInP Cladding
Layers, IEEE J. Photovoltaics, 6 (2016) 191-195.
[132] J. Simon, K.L. Schulte, D.L. Young, A.J. Ptak, Low cost GaAs solar cells grown by
hydride vapor phase epitaxy and the development of GaInP cladding layers, in:
Photovoltaic Specialist Conference (PVSC), 2015 IEEE 42nd, 2015, pp. 1-4.
[133] J. Simon, D. Young, A. Ptak, Low-cost III-V solar cells grown by hydride vapor-
phase epitaxy, in: 2014 IEEE 40th Photovoltaic Specialist Conference (PVSC), 2014, pp.
0538-0541.
[134] G. Létay, M. Hermle, A.W. Bett, Simulating single-junction GaAs solar cells
including photon recycling, Prog. Photovolt: Res. Appl., 14 (2006) 683-696.
[135] A. Abate, T. Leijtens, S. Pathak, J. Teuscher, R. Avolio, M.E. Errico, J. Kirkpatrik,
J.M. Ball, P. Docampo, I. McPherson, H.J. Snaith, Lithium salts as "redox active" p-type
dopants for organic semiconductors and their impact in solid-state dye-sensitized solar
cells, Phys. Chem. Chem. Phys., 15 (2013) 2572-2579.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
[136] A. Baktash, O. Amiri, A. Sasani, Improve efficiency of perovskite solar cells by
using Magnesium doped ZnO and TiO2 compact layers, Superlattices Microstruct., 93
(2016) 128-137.
[137] C.-H. Chiang, C.-G. Wu, Bulk heterojunction perovskite–PCBM solar cells with
high fill factor, Nat Photon, 10 (2016) 196-200.
[138] J.H. Heo, H.J. Han, D. Kim, T.K. Ahn, S.H. Im, Hysteresis-less inverted
CH3NH3PbI3 planar perovskite hybrid solar cells with 18.1% power conversion
efficiency, Energy Environ. Sci., 8 (2015) 1602-1608.
[139] B.S. Jeong, D.P. Norton, J.D. Budai, Conductivity in transparent anatase TiO2 films
epitaxially grown by reactive sputtering deposition, Solid-State Electron., 47 (2003)
2275-2278.
[140] T. Leijtens, J. Lim, J. Teuscher, T. Park, H.J. Snaith, Charge Density Dependent
Mobility of Organic Hole-Transporters and Mesoporous TiO2 Determined by Transient
Mobility Spectroscopy: Implications to Dye-Sensitized and Organic Solar Cells, Adv.
Mater., 25 (2013) 3227-3233.
[141] H.-H. Wang, Q. Chen, H. Zhou, L. Song, Z.S. Louis, N.D. Marco, Y. Fang, P. Sun,
T.-B. Song, H. Chen, Y. Yang, Improving the TiO2 electron transport layer in perovskite
solar cells using acetylacetonate-based additives, Journal of Materials Chemistry A, 3
(2015) 9108-9115.
[142] K. Wojciechowski, M. Saliba, T. Leijtens, A. Abate, H.J. Snaith, Sub-150 °C
processed meso-superstructured perovskite solar cells with enhanced efficiency, Energy
Environ. Sci., 7 (2014) 1142-1147.
[143] M.-C. Wu, S.-H. Chan, M.-H. Jao, W.-F. Su, Enhanced short-circuit current density
of perovskite solar cells using Zn-doped TiO2 as electron transport layer, Sol. Energy
Mater. Sol. Cells, 157 (2016) 447-453.
[144] H. Zhou, Q. Chen, G. Li, S. Luo, T.-b. Song, H.-S. Duan, Z. Hong, J. You, Y. Liu,
Y. Yang, Interface engineering of highly efficient perovskite solar cells, Science, 345
(2014) 542-546.
[145] S. Jang, W.-B. Jung, C. Kim, P. Won, S.-G. Lee, K.M. Cho, M.L. Jin, C.J. An, H.-
J. Jeon, S.H. Ko, T.-S. Kim, H.-T. Jung, A three-dimensional metal grid mesh as a
practical alternative to ITO, Nanoscale, 8 (2016) 14257-14263.
[146] J. Zou, H.-L. Yip, S.K. Hau, A.K.-Y. Jen, Metal grid/conducting polymer hybrid
transparent electrode for inverted polymer solar cells, Appl. Phys. Lett., 96 (2010)
203301.
[147] Q.G. Du, W. Yue, Z. Wang, W.T. Lau, H. Ren, E.-P. Li, High optical transmittance
of aluminum ultrathin film with hexagonal nanohole arrays as transparent electrode, Opt.
Express, 24 (2016) 4680-4688.
[148] E.S. Román, A. Vitrey, J. Buencuerpo, I. Fernández, I. Prieto, B. Alén, A. García-
Martín, J.M. Llorens, S.R.J. Brueck, J.M. Ripalda, High transmission nanowire contact
arrays with subwavelength spacing, physica status solidi (RRL) – Rapid Research
Letters, 10 (2016) 164-167.
[149] Y. Jin, Q. Li, M. Chen, G. Li, Y. Zhao, X. Xiao, J. Wang, K. Jiang, S. Fan, Large
area nanoscale metal meshes for use as transparent conductive layers, Nanoscale, 7
(2015) 16508-16515.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
[150] M.-G. Kang, H. Joon Park, S. Hyun Ahn, L. Jay Guo, Transparent Cu nanowire
mesh electrode on flexible substrates fabricated by transfer printing and its application in
organic solar cells, Sol. Energy Mater. Sol. Cells, 94 (2010) 1179-1184.
[151] M.-G. Kang, M.-S. Kim, J. Kim, L.J. Guo, Organic Solar Cells Using
Nanoimprinted Transparent Metal Electrodes, Adv. Mater., 20 (2008) 4408-4413.
[152] M.G. Kang, L.J. Guo, Nanoimprinted Semitransparent Metal Electrodes and Their
Application in Organic Light-Emitting Diodes, Adv. Mater., 19 (2007) 1391-1396.
[153] T. Nakanishi, A. Fujimoto, R. Kitagawa, K. Masunaga, K. Nakamura, E. Tsutsumi,
K. Asakawa, Electrical resistivity of transparent metal nanomesh electrodes, physica
status solidi (a), 210 (2013) 327-334.
[154] J. Kang, C.-G. Park, S.-H. Lee, C. Cho, D.-G. Choi, J.-Y. Lee, Fabrication of high
aspect ratio nanogrid transparent electrodes via capillary assembly of Ag nanoparticles,
Nanoscale, 8 (2016) 11217-11223.
[155] H.-J. Choi, S.-W. Ryu, J. Jun, S. Moon, D. Huh, Y.D. Kim, H. Lee, Fabrication of a
transparent conducting Ni-nanomesh-embedded film using template-assisted Ni
electrodeposition and hot transfer process, RSC Adv., 6 (2016) 81814-81817.
[156] J.F. Zhu, B.Q. Zeng, Z. Wu, Enhanced Broadband Optical Transmission Through
Ultrathin Metallic Nanomesh, Journal of Electromagnetic Waves and Applications, 26
(2012) 342-352.
[157] N. Tsutomu, T. Eishi, M. Kumi, F. Akira, A. Koji, Transparent Aluminum
Nanomesh Electrode Fabricated by Nanopatterning Using Self-Assembled Nanoparticles,
Applied Physics Express, 4 (2011) 025201.
[158] B. Han, Q. Peng, R. Li, Q. Rong, Y. Ding, E.M. Akinoglu, X. Wu, X. Wang, X.
Lu, Q. Wang, G. Zhou, J.-M. Liu, Z. Ren, M. Giersig, A. Herczynski, K. Kempa, J. Gao,
Optimization of hierarchical structure and nanoscale-enabled plasmonic refraction for
window electrodes in photovoltaics, Nat. Commun., 7 (2016) 12825.
[159] C.F. Guo, T. Sun, Q. Liu, Z. Suo, Z. Ren, Highly stretchable and transparent
nanomesh electrodes made by grain boundary lithography, Nat. Commun., 5 (2014)
3121.
[160] J. Chua, N. Mathews, J.R. Jennings, G. Yang, Q. Wang, S.G. Mhaisalkar, Patterned
3-dimensional metal grid electrodes as alternative electron collectors in dye-sensitized
solar cells, Phys. Chem. Chem. Phys., 13 (2011) 19314-19317.
[161] Y. Galagan, B. Zimmermann, E.W.C. Coenen, M. Jørgensen, D.M. Tanenbaum,
F.C. Krebs, H. Gorter, S. Sabik, L.H. Slooff, S.C. Veenstra, J.M. Kroon, R. Andriessen,
Current Collecting Grids for ITO-Free Solar Cells, Advanced Energy Materials, 2 (2012)
103-110.
[162] D.S. Ghosh, T.L. Chen, V. Pruneri, High figure-of-merit ultrathin metal transparent
electrodes incorporating a conductive grid, Appl. Phys. Lett., 96 (2010) 041109.
[163] L. Qi, J. Li, C. Zhu, Y. Yang, S. Zhao, W. Song, Realization of a flexible and
mechanically robust Ag mesh transparent electrode and its application in a PDLC device,
RSC Adv., 6 (2016) 13531-13536.
[164] S. Kiruthika, K.D.M. Rao, K. Ankush, G. Ritu, G.U. Kulkarni, Metal wire network
based transparent conducting electrodes fabricated using interconnected crackled layer as
template, Materials Research Express, 1 (2014) 026301.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.
[165] K. Pei, Z. Wang, X. Ren, Z. Zhang, B. Peng, P.K.L. Chan, Fully transparent
organic transistors with junction-free metallic network electrodes, Appl. Phys. Lett., 107
(2015) 033302.
[166] K.D.M. Rao, C. Hunger, R. Gupta, G.U. Kulkarni, M. Thelakkat, A cracked
polymer templated metal network as a transparent conducting electrode for ITO-free
organic solar cells, Phys. Chem. Chem. Phys., 16 (2014) 15107-15110.
[167] J.-M. Park, T.-G. Kim, K. Constant, K.-M. Ho, Fabrication of submicron metallic
grids with interference and phase-mask holography, MOEMS, 10 (2011) 013011-
013011-013015.
[168] K. Azuma, K. Sakajiri, H. Matsumoto, S. Kang, J. Watanabe, M. Tokita, Facile
fabrication of transparent and conductive nanowire networks by wet chemical etching
with an electrospun nanofiber mask template, Mater. Lett., 115 (2014) 187-189.
[169] F. Yiin-Kuen, L. Li-Chih, Pattern transfer of aligned metal nano/microwires as
flexible transparent electrodes using an electrospun nanofiber template, Nanotechnology,
24 (2013) 055301.
[170] J.H. Park, D.Y. Lee, Y.-H. Kim, J.K. Kim, J.H. Lee, J.H. Park, T.-W. Lee, J.H.
Cho, Flexible and Transparent Metallic Grid Electrodes Prepared by Evaporative
Assembly, ACS Appl. Mater. Interfaces, 6 (2014) 12380-12387.
[171] J.H. Park, D.Y. Lee, W. Seung, Q. Sun, S.-W. Kim, J.H. Cho, Metallic Grid
Electrode Fabricated via Flow Coating for High-Performance Flexible Piezoelectric
Nanogenerators, The Journal of Physical Chemistry C, 119 (2015) 7802-7808.
[172] M. Layani, S. Magdassi, Flexible transparent conductive coatings by combining
self-assembly with sintering of silver nanoparticles performed at room temperature,
Journal of Materials Chemistry, 21 (2011) 15378-15382.
[173] K. Tvingstedt, O. Inganäs, Electrode Grids for ITO Free Organic Photovoltaic
Devices, Adv. Mater., 19 (2007) 2893-2897.
[174] Z. Zhang, X. Zhang, Z. Xin, M. Deng, Y. Wen, Y. Song, Controlled Inkjetting of a
Conductive Pattern of Silver Nanoparticles Based on the Coffee-Ring Effect, Adv.
Mater., 25 (2013) 6714-6718.
[175] M. Mohl, A. Dombovari, R. Vajtai, P.M. Ajayan, K. Kordas, Self-assembled large
scale metal alloy grid patterns as flexible transparent conductive layers, Sci. Rep., 5
(2015) 13710.
[176] D.-E. Lee, S. Go, G. Hwang, B.D. Chin, D.H. Lee, Two-Dimensional
Micropatterns via Crystal Growth of Na2CO3 for Fabrication of Transparent Electrodes,
Langmuir, 29 (2013) 12259-12265.
[177] I. Kim, S.-W. Kwak, Y. Ju, G.-Y. Park, T.-M. Lee, Y. Jang, Y.-M. Choi, D. Kang,
Roll-offset printed transparent conducting electrode for organic solar cells, Thin Solid
Films, 580 (2015) 21-28.
[178] J. van de Groep, D. Gupta, M.A. Verschuuren, M. M. Wienk, R.A.J. Janssen, A.
Polman, Large-area soft-imprinted nanowire networks as light trapping transparent
conductors, Sci. Rep., 5 (2015) 11414.

Pursuant to the DOE Public Access Plan, this document represents the authors' peer-reviewed, accepted manuscript.
The published version of the article is available from the relevant publisher.

You might also like