169
FABRICATION AND TRANSFER ASSEMBLY OF MICROSCALE, SOLID-STATE LIGHT EMITTING DIODES AND SOLAR CELLS FOR TRANSPARENT AND FLEXIBLE ELECTRONICS APPLICATIONS BY ERIC P. BRUECKNER DISSERTATION Submitted in partial fulfillment of the requirements for the degree of Doctor of Philosophy in Chemistry in the Graduate College of the University of Illinois at Urbana-Champaign, 2013 Urbana, Illinois Doctoral Committee: Professor Ralph G. Nuzzo, Chair Professor John A. Rogers Professor Placid M. Ferreira Professor Catherine J. Murphy

fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

FABRICATION AND TRANSFER ASSEMBLY OF MICROSCALE, SOLID-STATE LIGHT

EMITTING DIODES AND SOLAR CELLS FOR TRANSPARENT AND FLEXIBLE

ELECTRONICS APPLICATIONS

BY

ERIC P. BRUECKNER

DISSERTATION

Submitted in partial fulfillment of the requirements

for the degree of Doctor of Philosophy in Chemistry

in the Graduate College of the

University of Illinois at Urbana-Champaign, 2013

Urbana, Illinois

Doctoral Committee:

Professor Ralph G. Nuzzo, Chair

Professor John A. Rogers

Professor Placid M. Ferreira

Professor Catherine J. Murphy

Page 2: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

ii

Abstract

Efficiency metrics for some solid-state electronic materials systems have progressed to

the point where theoretical limits are being approached. Gallium nitride-based light-

emitting diodes and silicon solar cells, for example, have achieved such extraordinarily

high performance metrics that only incremental improvements upon them are expected in

the next decade of intense research. This pseudo-plateau in performance development

means concentrated effort can now be placed on strategic implementation of these

materials into platforms that fill a growing demand for high-performance consumer

products. Such products have traditionally relied upon large-scale materials, but

possibilities now exist for manipulating micro-scale, wafer-based devices in ways that

promote improvements in areas of electrical current spreading, light absorption and

extraction, and thermal management. To this end, my research has focused on routes to

fabricating and assembling solid-state light-emitting diodes and solar cells of indium

gallium nitride and single-crystalline silicon, respectively, in configurations which

optimize characteristics of their performance. Specifically, I have worked, in

collaboration with others, to achieve a processing strategy that creates dense arrays of

indium gallium nitride light-emitting diodes on a silicon wafer of (111) orientation and

assemble them onto transparent and flexible substrates. This work produced novel form

factors for solid-state lighting where small, light-emitting devices were spatially

distributed and integrated with color-converting phosphors in ways that controllably

tuned their chromaticity. We also demonstrated that incredible passive heat dissipation

with these micro-scale elements stemming naturally from their small size and integration

Page 3: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

iii

with metal films serving dually as an electrically interconnecting medium. The cell

design and etching strategies used were then transferred to a single-crystalline silicon

system where small, ribbon-like solar cells were fabricated. This work improved upon

previous studies creating similar devices by increasing critical solar cell performance

metrics. The developed solar cell structure utilizes a highly robust manufacturing layer

of thermally-grown silicon dioxide which naturally doubles as an anti-reflection and

passivation layer. Other improvements to previous performance metrics comes from

optimized cell assembly onto structures that recycle and redistribute incident irradiation.

Page 4: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

iv

Courtney: loving, understanding, and a wonderful mother

LOVE

Page 5: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

v

Acknowledgements

My graduate work at the University of Illinois has been the product of countless

contributions from a host of individuals. Such a means of acknowledgement inherently

implies an order of rank or importance, when, in reality, all contributions were equally

indispensible.

To Professor Ralph Nuzzo, your guidance and insight to problems big or small is

something I will always remember. You have the uncanny ability to look at the big

picture and develop a unique plan of action. Your ideas and suggestions have

immeasurably aided my work at the University of Illinois.

To Professor John Rogers, I have the utmost respect for the work that you and

your group has and will continue to generate. Thank you for your constant help during

my time at UIUC.

To the Nuzzo Group members, the vast and disparate knowledge base that comes

from this group is truly unique and something that has served me well during my time

here. Thank you to all group members, including Matthew Smalls, Dr. Sergio Sanchez,

Michael Cason, Jason Goldman, Dr. Evan Erickson, Enes Oruc, Dr. Lucas Thompson,

Dr. Chris Anderton, Dr. An-Phong Le, Dr. Huaibin Zhang, and others, for being a

constant source of camaraderie, idea generation, and problem solving. Thank you to

those whom I have had the distinct pleasure of collaborating with or otherwise

developing similar projects with: Yuan Yao, Dr. Hoon-sik Kim, Dr. Lanfang Li, Chris

Corcoran, Dr. Michael Motala, and Dr. Audrey Bowen, and others.

Page 6: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

vi

To my family, you never always understood why I would continue to be a student

for so long. Thank you for your constant and unwavering love and support throughout

my entire academic career. Having this source of unconditional love is something that I

will always cherish.

And to Courtney, words cannot begin to give due justice to the love and support

you have given me throughout my graduate career. Thank you for all of your personal

and professional sacrifices you have made to make this accomplishment possible. These

past five years have not always been easy, but your constant commitment to me and to

our relationship is something that I will treasure forever.

Page 7: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

vii

Table of Contents

Chapter 1 Introduction to Micro-Transfer Printing ............................................................ 1

1.1 Overview of Dissertation .......................................................................................... 1

1.2 Introduction to Rate-Dependent Control of Micro-Transfer Printing....................... 1

1.3 Overview of Stamp Fabrication ................................................................................ 3

1.4 Stamp Geometry ....................................................................................................... 5

1.5 Chapter 1 Figures ...................................................................................................... 9

1.6 Chapter 1 References .............................................................................................. 12

Chapter 2 Release Strategies for Microscale, Solid-State Semiconductor Elements ....... 15

2.1 Overview ................................................................................................................. 15

2.2 Isotropic and Anisotropic Etching Systems ............................................................ 16

2.3 Design Considerations for Isotropically Undercut Devices .................................... 19

2.4 Design Considerations for Anisotropically Undercut Devices ............................... 22

2.5 Discussion on Wafer Utilization Calculations ........................................................ 29

2.6 Chapter 2 Figures .................................................................................................... 33

2.7 Chapter 2 References .............................................................................................. 39

Chapter 3 Unusual Strategies for Using InGaN Grown on Silicon (111) for Solid State

Lighting ............................................................................................................................. 43

3.1 Motivation for Microscale Light Emitting Diodes ................................................. 43

3.2 Overview of Microscale, InGaN Light Emitting Diode Fabrication ...................... 45

3.3 Large Area, White-Light Panels ............................................................................. 50

Page 8: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

viii

3.4 Thermal Dissipation in μ-ILEDs ............................................................................ 52

3.5 Detailed Processing Protocol for InGaN μ-ILEDs and Interconnected μ-ILED

Arrays ............................................................................................................................ 56

3.6 Analytical Model of Printed μ-ILEDs on a Glass Substrate ................................... 66

3.7 Discussion on Device Characterization .................................................................. 72

3.8 Chapter 3 Figures .................................................................................................... 75

3.9 Chapter 3 References .............................................................................................. 93

Chapter 4 Fabrication of High Performance Silicon Solar Microcells Integrating

Passivation and Antireflection Coatings ........................................................................... 95

4.1 Introduction ............................................................................................................. 95

4.2 Fabrication of Silicon Solar Microcells .................................................................. 98

4.3 Silicon Solar μ-Cell Characterization ................................................................... 100

4.4 Thermal Oxide Passivation and Anti-Reflection Coating .................................... 100

4.5 Packaging Silicon Solar μ-Cell on Secondary Substrates ..................................... 103

4.6 Optical Enhancements to Packaged μ-Cells ......................................................... 104

4.7 Silicon Solar μ-Cell Module ................................................................................. 107

4.8 Conclusion ............................................................................................................ 108

4.9 Detailed Processing Protocol for Silicon Solar Microcells................................... 109

4.10 Future Directions for Silicon Solar μ-Cell Research .......................................... 117

4.11 Chapter 4 Figures ................................................................................................ 122

4.12 Chapter 4 References .......................................................................................... 140

Chapter 5 Nanotextured and Anti-Reflective PDMS...................................................... 143

Page 9: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

ix

5.1 Introduction ........................................................................................................... 143

5.2 Experimental ......................................................................................................... 147

5.3 Nano-Scale Columnar Texture on Plasma-Etched PDMS.................................... 149

5.4 Future Work on Nanotextured PDMS .................................................................. 151

5.5 Chapter 5 Figures .................................................................................................. 155

5.6 Chapter 5 References ............................................................................................ 159

Page 10: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

1

Chapter 1

Introduction to Micro-Transfer Printing

1.1 Overview of Dissertation

This dissertation starts by speaking broadly on an area of science which has

gained traction in the past half-decade; techniques for creating assembled arrays of

microscale (μ-scale), high performance, solid-state semiconductor elements on non-

traditional substrates. My research at the University of Illinois has been intimately

related to this field. Within this dissertation I describe a tool-set whereby μ-scale

electronic devices can be fabricated in dense arrays on and lifted off of a donor wafer

then transferred to a secondary substrate. It is my impression that a comprehensive

overview of design criteria for the entire process has not yet been clearly defined.

Therefore, I present guiding principles which are intended to aid the researcher in

developing a suitable fabrication protocol for all aspects of this field of μ-scale device

fabrication and assembly.

1.2 Introduction to Rate-Dependent Control of Micro-Transfer Printing

Transfer printing [1] was developed as a means to assemble μ-scale devices from

a donor wafer onto secondary substrates. In most embodiments of transfer printing, a

soft, elastomeric stamp is brought into conformal contact with a device which is weakly

Page 11: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

2

adhered to the host wafer by means discussed in Chapter 2. Upon conformal contact of

the elastomeric stamp, typically polydimethylsiloxane (PDMS, Dow Corning Sylgard

184), van der Waals interactions [2] at the stamp/device interface serve as the basis for

adhesion to remove the device from the donor wafer. Studies have been performed by

Feng., et al. concerning the rate dependent nature of competing fracture modes in the

device/PDMS/substrate system. Interfacial fracture at any of these interfaces is a

manifestation of the critical value of energy release rate G. In general, the guiding

expression for device pick-up (referred to here as the PDMS stamp retrieving the device

from the donor wafer) is

/ /stamp device device substrate

c cG G

1.1

and device printing (referred to here as the transfer of the device to the secondary

substrate) is governed by

/ /stamp device object substrate

c cG G

1.2

They observed that pick-up rates were maximized when the peel velocity was high and

that the printing rates were maximized when the peel velocity was low (illustrated

schematically in Figure 1.1). This infers that there is a critical peel velocity v, above

which promotes pick-up and below promotes printing. They found that Gc for a 1 mm

thick PDMS stamp on a 100 nm thick Au film is

Page 12: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

3

0.65

/

13( ) 2.16 1

1.81 10

stamp device tc

a vG v

1.3

Where aT is the temperature-shift factor and is approximated as:

g

g

TTT

TTa

526.17log10

1.4

for most polymers within 50K of their glass transition temperature, Tg [3].

1.3 Overview of Stamp Fabrication

As mentioned previously, most transfer printing protocols call for the use of

PDMS as the stamp material, as shown in references [1, 4-37], just to name a few. There

are primarily two modes whereby μ-scale devices are lifted off of the donor wafer and

transferred to the secondary substrate. The first is one-to-one printing of the device array.

In this simplified route of transfer printing the devices are printed en masse, without

selective retrieval of certain devices. Therefore, the registration of devices to one another

on the receiving substrate can be “pre-programmed” by appropriately designing the donor

wafer to contain the required layout of features. This protocol is particularly useful when

high-precision stages are not available to accurately align the stamp to micron-size

devices.

Page 13: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

4

A far more powerful embodiment of the transfer printing process is that of

selective retrieval and “areal expansion” of the devices from the donor wafer to the

secondary, receiving substrate. This technique allows for extremely high donor wafer

utilization and minimal costs associated with wasted active material due to wafer dicing.

For example, commercial laser dicing systems can achieve kerf widths of ~50µm [38]. If

this technique is used to dice 100µm x 100µm devices, 56% of the wafer is being lost to

the dicing process. Transfer printing approaches typically use microfabricated devices

which are delineated with sub-micron precision photolithography patterning and etching

tools. With this particular tool-set, extremely narrow kerf widths, ≤2µm, can be

achieved. This geometry gives ≥96% wafer utilization, a value unmatched by

conventional wafer-dicing techniques. Additionally, wafer-processing tools, such as

reactive ion etching, are a massively parallel process whereby all devices are being

delineated simultaneously. A traditional wafer-dicing technique is a serial process,

meaning only one die can be diced at a given time.

Enabling such capabilities is the process of replica molding [39] PDMS to a

photo-resist patterned silicon wafer to create relief features which will selectively contact

devices on the donor wafer, avoiding mass pick-up of all devices. As shown in Figure

1.2, the process starts with a clean Si wafer. The photo-resist (SU8, Microchem) is spun

on the wafer in any desired thickness. SU8 is an epoxy based photo-resist and is cross-

linked upon exposure to ultraviolet light, rendering it insoluble in developer solvent. To

selectively pattern areas for subsequent relief features, an appropriately designed

photolithography mask is pressed into contact with the SU8 layer and exposed with i-line

Page 14: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

5

UV radiation. A brief hard bake further increases the cross-linking density and the

unexposed regions are ready for development. Using a solvent-based developer,

unexposed regions are dissolved and removed, leaving only the exposed, cross-linked

regions of the photo-resist behind. These features serve as the mold for subsequent

replications of the “master” mold. This SU8 master is then exposed to tridecafluoro-

1,1,2,2-tetrahydrooctyl trichlorosilane to render the surface hydrophobic and prevent

irreversible bonding of the PDMS to the wafer or SU8 layer. Finally, uncured PDMS is

cast over the mold and cured. The PDMS mold is peeled away giving a negative

topographical image of the mold (i.e., valleys in the master form peaks in the SU8).

1.4 Stamp Geometry

The PDMS stamp has been used in a number of different configurations

depending on the level of production or adhesion required for a particular application. In

its simplest form, a single post is used to contact a single device for transfer. This gives

the highest level of flexibility for creating abstract arrays of devices on the secondary

substrate. For instance, single GaN LED pixel elements were printed on a glass slide

substrate in reference [31]. Repeating this process 356 times spells the word

“ILLINOIS”. For higher throughput manufacturing, multiple devices can be contacted

and printed simultaneously. This particular mode of printing is demonstrated in reference

[22] where multiple devices are contacted by a PDMS stamp with multiple relief features

matching the spatial layout of devices on the donor wafer. When printing multiple

devices, considerations should be given to the desired center-to-center spacing of devices

Page 15: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

6

on the secondary substrate. It is this pitch, or whole multiples thereof, that will dictate

the device layout on the donor wafer and also determine how the stamp should be

designed (i.e., the center-to-center pitch of printed devices will be whole-number

multiples of the center-to-center pitch of devices on the donor wafer). A schematic

representation can be found in Figure 1.3. This example helps demonstrate the massively

parallel production of sparse arrays of devices while maintaining an extremely tight

packing density of fabricated devices on-wafer.

When considering a device/stamp/substrate system, one needs to consider the

relative forces of adhesion between all systems. For low device/stamp adhesion,

ultraviolet ozone (UVO) or oxygen plasma “activation” of the stamp will increase

adhesion at the device/stamp interface. When the surface of the device is comprised of

SiO2, silicon nitride, Si, PDMS, polyethylene, polystyrene, or glassy carbon, such a

treatment can result in covalent, Si-O-Si bonding across the interface [40]. X-ray

photoelectron spectroscopy and contact angle measurements have shown that UVO

and/or oxygen plasma oxidation of PDMS hydroxylates the surface (-SiOH). Upon

contact to a complementary surface, condensation reactions covalently bind the two

interfaces [40-42]. This interface is largely useless for transfer-printing techniques, as

cohesive failure into the bulk of the PDMS stamp is the preferred mode of failure [43]

when sufficient force is exerted to remove the device from the stamp.

A much higher demand exists for improving the ability to transfer devices from

the stamp to a wide range of materials. Two general approaches can be taken to aid the

transferability of the device from the stamp to substrate: increase device/substrate

Page 16: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

7

adhesion or decrease device/stamp adhesion. Many examples exist of the former being

used and is considered the simpler of the two methods discussed. This is as achieved by

spin-coating a thin film polymeric material such as PDMS [22], SU8 [14, 35], polyimide

[16], NOA [19, 33], and others [31, 36]. Using a thin-film adhesive layer provides a

simple approach to transfer printing when active interactions between the device and

substrate are not needed [44]. Most applications using transfer printing do not require

such active interactions; the substrate merely serves as a vehicle to achieve transparent,

stretchable, and/or transparent modules.

Three dimensional, heterogeneous integration is gaining traction as a route to

create multifunctional chips with individually optimized components for added

functionality [45]. This technique is said to help extend Moore’s Law by allowing more

powerful materials systems to be optimized and integrated into a highly functional

package. However, 3D integration relies on pick-and-placement of a processed die and

thermocompression to bond the two layers together [46, 47]. But thermocompression

relies on elevated temperatures and pressures, adding energy and time costs as well as

potentially damaging fragile device layers. The ability to transfer discrete device

components with relatively minimal processing might be considered the holy grail of 3D

integration making transfer printing active ink materials to dissimilar substrates an

attractive technology. However, fulfilling the fundamental criterion of Eqn.1.2 is very

difficult due to extremely small interfacial adhesion between the ink and substrate, most

often a solid, low surface energy wafer material. Extensive research has been done to

reduce the amount of ink/stamp adhesion, usually through dynamic control of adhesion

Page 17: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

8

from the lift-off to the transfer process through reduction of the stamp’s contact area to

the ink by mechanical means [48], laser-driven delamination of the ink material [49], or

viscoelastic restoration of a structured stamp [26]. The latter provides a particularly

interesting approach as it relies solely on the restorative forces within the PDMS stamp to

retract the collapsed roof of the stamp during lift-off to provide an ultra-low contact area

(~0.07%) of the stamp to the ink. This allows Eqns. 1.1 and 1.2 to be realized, even for

solid-to-solid printing.

Page 18: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

9

1.5 Chapter 1 Figures

Donor wafer

“Inks”

PDMS Stamp

Donor wafer

“Inks”

PDMS Stamp

Donor wafer

“Inks”

“Inked” Stamp

“Inks”

“Inked” Stamp

“Inks”

“Inked” Stamp

Receiving Substrate

Printed “Inks”

PDMS Stamp

Receiving Substrate

Receiving Substrate

Contact “Inks”

Fast Retraction Contact Receiving Substrate

Slow Retraction

Figure 1.1: Schematic representation of the transfer printing process.

Page 19: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

10

Spin-coat photoresist

UV exposeDevelop

Pour PDMS pre-polymer

“No Stick”treatment

Peel away PDMS mold

Si wafer

Si wafer

SU8

Si wafer

SU8

Si wafer

SU8“No Stick”

Si wafer

SU8“No Stick”

PDMS

PDMS

Figure 1.2: Schematic representation of the process flow for creating selective area

stamps for transfer printing.

Page 20: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

11

Figure 1.3: Schematic representing areal expansion of a) sparsely arrayed devices on a

secondary substrate from b) a densely packed array on-wafer. Using a stamp configured

in c), devices are picked up in multiple steps (identically colored devices coming from

the same print step) and spread out over areas much larger than those obtainable on small,

expensive donor wafers.

Page 21: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

12

1.6 Chapter 1 References

1. Meitl, M.A., et al., Transfer Printing by Kinetic Control of Adhesion to an

Elastomeric Stamp. Nat. Mater., 2006. 5: p. 33-38.

2. Feng, X., et al., Competing Fracture in Kinetically Controlled Transfer Printing.

Langmuir, 2007. 23: p. 12555-12560.

3. Williams, M.L., R.F. Landel, and J.D. Ferry, The Temperature Dependence of

Relaxation Mechanisms in Amorphous Polymers and Other Glass-forming

Liquids. J. Am. Chem. Soc., 1955. 77: p. 3701-3707.

4. Jackman, R., J. Wilbur, and G. Whitesides, Fabrication of Submicrometer

Features on Curved Substrates by Microcontact Printing. Science, 1995.

269(5224): p. 664-666.

5. Xia, Y., et al., Microcontact Printing of Octadecylsiloxane on the Surface of

Silicon Dioxide and Its Application in Microfabrication. Journal of the American

Chemical Society, 1995. 117(37): p. 9576-9577.

6. James, C.D., et al., Patterned Protein Layers on Solid Substrates by Thin Stamp

Microcontact Printing. Langmuir, 1998. 14(4): p. 741-744.

7. Childs, W.R. and R.G. Nuzzo, Decal Transfer Microlithography: A New Soft-

Lithographic Patterning Method. J. Am. Chem. Soc., 2002. 124: p. 13583-13596.

8. Loo, Y.-L., et al., Interfacial Chemistries for Nanoscale Transfer Printing.

Journal of the American Chemical Society, 2002. 124(26): p. 7654-7655.

9. Childs, W.R. and R.G. Nuzzo, Patterning of Thin-Film Microstructures on Non-

Planar Substrate Surfaces Using Decal Transfer Lithography. Adv. Mater., 2004.

16: p. 1323-1327.

10. Jeon, S., et al., Three-Dimensional Nanofabrication with Rubber Stamps and

Conformable Photomasks. Adv. Mater., 2004. 16: p. 1369-1373.

11. Sun, Y. and J.A. Rogers, Fabricating Semiconductor Nano/Microwires and

Transfer Printing Ordered Arrays of Them onto Plastic Substrates. Nano Letters,

2004. 4(10): p. 1953-1959.

12. Lee, K.J., et al., Large-area, Selective Transfer of Microstructured Silicon: A

Printing-based Approach to High-performance Thin-film Transistors Supported

on Flexible Substrates. Adv. Mater., 2005. 17: p. 2332-2336.

13. Ahn, J.-H., et al., Heterogeneous Three-Dimensional Electronics by Use of

Printed Semiconductor Nanomaterials. Science, 2006. 314: p. 1754-1757.

14. Mack, S., et al., Mechanically Flexible Thin-Film Transistors that use Ultrathin

Ribbons of Silicon Derived from Bulk Wafers. Applied Physics Letters, 2006.

88(21): p. 213101-3.

Page 22: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

13

15. Sun, Y., et al., Controlled Buckling of Semiconductor Nanoribbons for

Stretchable Electronics. Nat Nano, 2006. 1(3): p. 201-207.

16. Kim, D.-H., et al., Materials and Noncoplanar Mesh Designs for Integrated

Circuits with Linear Elastic Responses to Extreme Mechanical Deformations. P.

Natl. Acad. Sci. USA, 2008. 105: p. 18675-18680.

17. Kim, D.-H., et al., Stretchable and Foldable Silicon Integrated Circuits. Science,

2008. 320(5875): p. 507-511.

18. Ko, H.C., et al., A Hemispherical Electronic Eye Camera Based on Compressible

Silicon Optoelectronics. Nature, 2008. 454(7205): p. 748-753.

19. Yoon, J., et al., Ultrathin Silicon Solar Microcells for Semitransparent,

Mechanically Flexible and Microconcentrator Module Designs. Nat. Mater.,

2008. 7: p. 907-915.

20. Kim, D.-H., et al., Ultrathin Silicon Circuits With Strain-Isolation Layers and

Mesh Layouts for High-Performance Electronics on Fabric, Vinyl, Leather, and

Paper. Advanced Materials, 2009. 21(36): p. 3703-3707.

21. Kim, K.S., et al., Large-Scale Pattern Growth of Graphene Films for Stretchable

Transparent Electrodes. Nature, 2009. 457(7230): p. 706-710.

22. Park, S.-I., et al., Printed Assemblies of Inorganic Light-Emitting Diodes for

Deformable and Semitransparent Displays. Science, 2009. 325: p. 977-981.

23. Baca, A.J., et al., Compact Monocrystalline Silicon Solar Modules With High

Voltage Outputs and Mechanically Flexible Designs. Energy Environ. Sci., 2010.

3: p. 208-211.

24. Kim, D.-H., et al., Dissolvable Films of Silk Fibroin for Ultrathin Conformal Bio-

Integrated Electronics. Nat Mater, 2010. 9(6): p. 511-517.

25. Kim, R.-H., et al., Waterproof AlInGaP Optoelectronics on Stretchable Substrates

with Applications in Biomedicine and Robotics. Nat. Mater., 2010. 9: p. 929-937.

26. Kim, S., et al., Microstructured Elastomeric Surfaces with Reversible Adhesion

and Examples of Their Use in Deterministic Assembly by Transfer Printing. P.

Natl. Acad. Sci. USA, 2010. 107: p. 17095-17100.

27. Moutanabbir, O. and U. Gosele, Heterogeneous Integration of Compound

Semiconductors. Annu. Rev. Mater. Res., 2010. 40: p. 469-500.

28. Park, K.-I., et al., Bendable and Transparent Barium Titanate Capacitors on

Plastic Substrates for High Performance Flexible Ferroelectric Devices.

Electrochem. Solid. St., 2010. 13: p. G57-G59.

29. Yoon, J., et al., GaAs Photovoltaics and Optoelectronics using Releasable

Multilayer Epitaxial Assemblies. Nature, 2010. 465: p. 329-334.

30. Kim, R.-H., et al., Stretchable, Transparent Graphene Interconnects for Arrays of

Microscale Inorganic Light Emitting Diodes on Rubber Substrates. Nano Lett.,

2011. 11: p. 3881-3886.

31. Kim, H.-s., et al., Unusual Strategies for using Indium Gallium Nitride Grown on

Silicon (111) for Solid-State Lighting. P. Natl. Acad. Sci. USA, 2011. 108: p.

10072-10077.

32. Chanda, D., et al., Large-area Flexible 3D Optical Negative Index Metamaterial

Formed by Nanotransfer Printing. Nat. Nanotechnol., 2011. 6: p. 402-407.

Page 23: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

14

33. Yoon, J., et al., Flexible Concentrator Photovoltaics Based on Microscale Silicon

Solar Cells Embedded in Luminescent Waveguides. Nat. Commun., 2011. 2: p.

343.

34. Kim, S., et al., Imbricate Scales as a Design Construct for Microsystem

Technologies. Small, 2012. 8(6): p. 901-906.

35. Lee, J., et al., Stretchable Semiconductor Technologies with High Areal

Coverages and Strain-Limiting Behavior: Demonstration in High-Efficiency

Dual-Junction GaInP/GaAs Photovoltaics. Small, 2012. 8(12): p. 1851-1856.

36. Kim, T.-i., et al., High-Efficiency, Microscale GaN Light-Emitting Diodes and

Their Thermal Properties on Unusual Substrates. Small, 2012. 8(11): p. 1643-

1649.

37. Park, K.-I., et al., Piezoelectric BaTiO3 Thin Film Nanogenerator on Plastic

Substrates. Nano Lett., 2010. 10: p. 4939-4943.

38. www.wafer-dicing.com. 2012.

39. McDonald, J.C., et al., Fabrication of Microfluidic Systems in

Poly(dimethylsiloxane). Electrophoresis, 2000. 21: p. 27-40.

40. Duffy, D.C., et al., Rapid Prototyping of Microfluidic Systems in

Poly(dimethylsiloxane). Anal. Chem., 1998. 70: p. 4974-4984.

41. Chaudhury, M.K. and G.M. Whitesides, Direct Measurement of Interfacial

Interactions between Semispherical Lenses and Flat Sheets of

Poly(dimethylsiloxane) and Their Chemical Derivatives. Langmuir, 1991. 7: p.

1013-1025.

42. Morra, M., et al., On the Aging of Oxygen Plasma-Treated Polydimethylsiloxane

Surfaces. J. Colloid Interf. Sci., 1989. 137: p. 11-24.

43. Childs, W.R., et al., Masterless Soft Lithography: Patterning UV/Ozone-Induced

Adhesion on Poly(dimethylsiloxane) Surfaces. Langmuir, 2005. 21: p. 10096-

10105.

44. Oh, D.-W., et al., Interfacial Thermal Conductance of Transfer-Printed Metal

Films. Advanced Materials, 2011. 23(43): p. 5028-5033.

45. Roelkens, G., et al., III-V/Si Photonics by Die-to-Wafer Bonding. Materials

Today, 2007. 10(7–8): p. 36-43.

46. Loh, G.H., 3D-Stacked Memory Architectures for Multi-Core Processors. Int.

Sym. Comp. Archit., 2008: p. 453-464.

47. Schmidt, M.A., Wafer-to-Wafer Bonding for Microstructure Formation.

Proceedings of the IEEE, 1998. 86(8): p. 1575-1585.

48. Carlson, A., et al., Active, Programmable Elastomeric Surfaces with Tunable

Adhesion for Deterministic Assembly by Transfer Printing. Advanced Functional

Materials, 2012: p. n/a-n/a.

49. Li, R., et al., Thermo-Mechanical Modeling of Laser-Driven Non-Contact

Transfer Printing: Two-Dimensional Analysis. Soft Matter, 2012. 8(27): p. 3122-

3127.

Page 24: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

15

Chapter 2

Release Strategies for Microscale, Solid-State Semiconductor

Elements

2.1 Overview

Transfer printing of solid inks inherently relies on the ability for the ink material

to be removed from the wafer on which the device was fabricated. At its root, this

requires selective etching of a sacrificial layer beneath the active device layer, thereby

delineating the active layer from the donor wafer. Multiple systems have emerged as

viable materials systems for undercut etching; those based on isotropic [1-11] and

anisotropic [12-18] etching of similar [12, 13, 15-18] or dissimilar [1-11, 14] materials.

The broad range of materials have lead to an equally large set of implementation

strategies that best exploit the etching characteristics of the system of interest. Such

strategies include a range of etchants [19, 20], masking protocols [5, 14, 16, 21, 22], and

device geometries [9, 12, 14]. All of these strategies must be considered when designing

a system for undercut etching so that critical device layers and components are properly

retained during the undercut process and that device performance is retained from the on-

wafer state. This chapter will direct attention to useful strategies and design

considerations for undercut etching solid-state semiconductor elements.

Page 25: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

16

2.2 Isotropic and Anisotropic Etching Systems

The growing demand for highly functional electronics has driven research into

exotic, epitaxially grown materials, such as Group III-V solar cells [23], lasers [24] and

light-emitting diodes (LED) [24], to the point where performance metrics are achieving

extremely high efficiencies and lifetimes [25]. Such systems benefit from the mature

silicon-based technology by utilizing processing tools and techniques adopted from

silicon processing, while expanding on the limited functionality provided by silicon-

based electronics (e.g., low electron mobility, low band-gap, etc.). New possibilities for

assembled, μ-scale electronics have arisen with this increased academic interest in exotic

materials. The range of materials and their chemical etching properties span a vast range.

Therefore, a powerful tool set must be developed to appropriately design protocols that

successfully release the active layers from their host wafer. These protocols vary from

simple immersion into an etching bath, to careful crystallographic orientation of the

devices, from highly selective etchants, to non-selective etchants requiring intricate

masking layers. This section will cover these criteria, giving design consideration

covering broad classes of materials.

The recent literature has given multiple examples of exemplary systems whereby

devices are fabricated on a donor wafer and successfully undercut and transferred to

secondary substrates. Isotropic etching (equal etch rates in all directions) of a sacrificial

layer has had a prominent role in transfer-printed devices. The most common and easily

processable materials system for undercut etching is that of silicon-on-insulator (SOI)

wafers, composed of a layer of silicon on a SiO2 layer bonded to a silicon handle wafer.

Page 26: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

17

Such a system provides a natural route to creating silicon-based devices; etching through

the silicon device layer then selectively removing the buried oxide layer with a

hydrofluoric acid (HF) based etchant. This general technique has been used to create

ultrathin membranes with unique electrical [26, 27], optical [28], and thermal [29, 30]

properties, ultra-thin transistors [31], and relatively thick devices for various

microelectromechanical systems (MEMS) [32].

A notable example of non-silicon based microelectronic fabrication and assembly

comes from Park, et al. [5] whereby GaAs-based LEDs were created and assembled on

substrates of glass, plastic, and PDMS. The essential technique for creating printable

devices was the ability to selectively etch a sacrificial layer of AlAs beneath the active

device layers without deleteriously effecting the other device layers and components

(e.g., quantum well). This work highlights critical process-design considerations; high

etch selectivity towards the sacrificial layer. It has long been known that an extremely

high selectivity exists in the Si/SiO2 system when exposed to HF, and this technology has

enabled the aforementioned SOI-based devices. To realize success in III-V materials, a

highly selective sacrificial layer needed to be discovered. This discovery was made in

1978 when high Al-containing alloys of AlGaAs were found to etch faster in HF than the

GaAs active layer [33]. This was further explored in 1987 by Yablonovitch, et al.,

finding that AlAs based alloys with 40-50% Al concentration exhibits a >107 higher etch

rate than the Al0.4Ga0.6As active layer [34]. Similar processing techniques were used to

fabricate and undercut BaTiO3 capacitors [3]. Barium titanate was sputtered on a

Pt/Ti/SiO2/Si(100) substrate and subsequently annealed which oxidized the Ti layer

Page 27: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

18

forming TiO2. After a dry etching step to define the device geometry, the TiO2 layer was

selectively removed in a buffered oxide etchant (BOE) creating a suspended, undercut

structure ready for transfer printing.

The second general route to undercut etching relies on highly directionally

selective etching of a single crystal material. Anisotropically etched systems have very

deep roots within the MEMS community, dating back to late 1970’s when Si was realized

to be a good candidate for mechanical systems [35, 36]. The unique etching properties of

Si accelerated this movement; Si could be etched with a high degree of selectivity and

reproducibility in certain crystallographic directions. This paved the road for detailed

studies on the etching characteristics and chemistries [37-39] involved in anisotropically

etching Si as well as novel devices to exploit this property. Consequential of the

orientation dependency of the etch, different classes of devices can be fabricated on

wafers of different crystallographic orientation, such as V-groove structures on Si(100)

[40, 41], high aspect-ratio trenches on Si(110) [40, 42], and suspended structures on

Si(111) [22]. Interesting approaches have been taken to create suspended structures by

selectively removing embedded material, but most of them are limited in the structure’s

lateral dimensions—because the undercut process is via isotropic etching of the surface

[43]—or doping profile which relies on embedded etch stop barriers composed of highly

doped layers [44], respectively, or relies on isotropic removal of a dissimilar material, as

discussed above. A more useful approach is to have almost no geometrical constraints

(lateral or vertical) as was developed by Lee, et al. [22] and further explored by others

[12, 16, 17, 21] where the vertical dimensions are pre-defined by a plasma etching step

Page 28: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

19

then a protective barrier is deposited and patterned on the sidewalls to permit lateral

etching in selected areas while blocking deleterious etching of the critical device layer.

2.3 Design Considerations for Isotropically Undercut Devices

Isotropically undercut devices, by convention, etch at equal rates in all directions

and, therefore, must be designed carefully to prevent full delamination of the active

device layer into the etchant bath. This is accomplished by using specially engineered

structures which are designed to tether the device to the wafer. Some methods of

undercutting devices have come in the form of complete removal of the device layer [6].

But such a process is not amenable to the massively parallel and repeatable processing of

transfer printing which inherently relies on knowing the relative locations of all ink

objects. To this end, a regarded necessity of creating devices for transfer printing is the

ability to retain the lithographically defined locations of the fabricated devices. I will

discuss herein some strategies used for tethering undercut devices to the donor wafer.

Two broad strategies exist for “anchoring” or tethering undercut devices in their

lithographically defined locations to the donor wafer: Homogeneous (portions of the

wafer material) and heterogeneous (using foreign material) anchors. In early examples of

homogeneous anchoring of isotropically undercut etched devices, SOI wafers were

vertically etched to expose the buried oxide layer then immersed in HF or BOE to

laterally remove the oxide layer. The devices were kept in their lithographically defined

location by careful design of the device geometry and tight control of the undercut etch

procedure such that undercut was incomplete at the two bulbous ends, tethering the

Page 29: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

20

device to the wafer, yet completely undercut in the remaining, narrow portion of the

“dumbbell”-type structure [45]. This approach is unfavorable because the processing

window for creating non-undercut SiO2 pillars at the bulbous ends which are sufficiently

small to allow for efficient lift-off while avoiding complete undercut, is very small. A

more favorable approach is to design structures similar to that illustrated in Figure 2.1,

showing a representative device to be undercut etched, where L1 and L2 represent the

lengths of the two orthogonal sides and WAnch represents the width of the anchor bar. The

time, t, required for full undercut etching can be approximated by assuming that the etch

proceeds equally from both sides of the device and is governed by the narrowest width of

device, LMin:

R

Lt Min5.0

2.1

where R is the etch rate of the material. When designing devices to be isotropically

undercut etched, the relative dimensions of the device should fulfill the minimum

requirement of:

MinAnch LW 5.1

2.2

Such a design criterion is needed because undercut etching will proceed at equal rates in

all directions, including beneath the anchor bar which serves to tether the devices to the

donor wafer. Over-etching, in the context of undercut etching, occurs when the sample is

Page 30: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

21

continually exposed to the etchant after full undercut of the device. In most practical

applications, a short (~15%) over-etch is appropriate to ensure that all devices are equally

undercut. Equation 2.2 highlights the low wafer utilization associated with homogeneous

anchoring systems for isotropically undercut devices. As is shown in Figure 2.2, wafer

utilization is relatively low with such an anchoring scheme. Please refer to Chapter 2.5

for discussion on assumptions made in calculating theoretical wafer utilization curves.

To alleviate the limitations of poor wafer utilization, a heterogeneous anchoring

scheme was developed for isotropically undercut-etched devices [5]. In this design,

foreign materials (usually polymeric materials such as photoresist) are patterned to cover

the corners of devices. The photoresist serves as the anchoring structure, eliminating the

need to use precious wafer material as the anchor, thereby dramatically improving wafer

utilization. Most applications call for an HF-based etchant to undercut the devices. In

these systems, photoresist is an excellent anchor material because positive photoresists

exhibit zero or near-zero etch rates in HF [20]. This allows for relaxed control for timing

the undercut etch.

Heterogeneous anchoring on an isotropically undercut-etched wafer follows a

similar fabrication protocol to that of a homogeneous anchoring scheme. A mesa etch

(plasma or wet etch) through the sacrificial layer is performed to isolate the devices and

expose the sidewalls of the sacrificial layer to be undercut. An additional

photolithography step is needed to pattern the photoresist anchors. A good representation

of the anchor geometry can be found in reference [5]. The devices are ready for undercut

etching and immersed in the etchant (usually HF-based). Unlike homogeneous anchoring

Page 31: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

22

schemes on isotropically undercut-etched wafers, heterogeneous anchors do not have

strict limitations on the undercut protocol (e.g., length of etch time) and design criteria.

Since photoresist exhibits such low (zero in most cases) etch rates in HF, there is little

concern for anchor failure due to extended etchant exposure. Additionally, the absence

of anchor bars further relaxes the processing conditions as there is no concern for

undercutting the tethering structure and releasing all of the devices from their

lithographically defined locations.

As with all micromachining processes, a single processing step might limit the

processes that can be utilized for future processing or deleteriously affect layers of the

fabricated device. Photoresist is a polymeric material and has a relatively low melting

temperature, making it incompatible with post-undercut high temperature processes such

as doping, thermal oxidation, and plasma enhanced chemical vapor deposition.

Additionally, HF (the most common etchant for isotropically undercut devices) is an

aggressive etchant that attacks a wide range of materials. Therefore, choosing a

compatible passivation and anti-reflective layer can be difficult as most are ceramic based

materials that etch quickly in HF such as SiO2, SiN, and Al2O3 [41, 46-48].

2.4 Design Considerations for Anisotropically Undercut Devices

Certain wafer/etchant systems display the unusual and potentially powerful

attribute of anisotropic etching (high etch selectivity towards certain crystallographic

planes). The most notable of these systems is silicon and potassium hydroxide. This

particular system has an enormously high anisotropy (~600x selectivity to Si(110) vs.

Page 32: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

23

Si(111) [40]) and has garnered most of the attention in microelectromechanical systems.

Other materials/etchants for anisotropic etching exist [49-51] albeit with far less

academic attention. For the purposes of this dissertation, I will restrict conversation to

the silicon wafer system.

The very large anisotropy realized in a crystalline silicon wafer originates from

the relative differences in packing density for the crystallographic planes [40]. Figure 2.3

shows a crystallographic projection of the Si(110) and Si(111) planes. Clearly the atomic

density of the (110) plane is less than that of the (111) plane. The (110) plane has a

higher number of dangling bonds making the surface more thermodynamically unstable

and more amenable to etchant attack. The disparate packing densities alone are not

enough to account for large differences in relative etch rates. Adsorption of H20 on the

etched surface is thought to play a role in further increasing the anisotropy of the etch by

frustrating etchant interaction with the Si surface. Higher atomic densities of the (111)

plane increase the efficacy of this process [36].

Seidel, et al. [39] proposed a mechanism for Si etching based on experimental

correlations between etch rates and activation energies of different crystallographic

planes. His model says that a surface Si atom with two dangling bonds reacts with OH-

to form –Si(OH)2. The electronegativity of the O atom weakens the Si-Si backbonds and

allows thermal excitation of two bonding electrons to the conduction band. This oxidized

silicon-hydroxide species reacts with two more OH- species forming orthosilicic acid,

Si(OH)4. Si(OH)4 leaves the surface by diffusion but is unstable at high pH and forms

SiO2(OH)2--, which was observed by in situ Raman spectroscopy in reference [38].

Page 33: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

24

While KOH has been the most researched etchant for Si, there remains a strong

drive to use tetramethyl ammonium hydroxide (TMAH) for its high etch-mask selectivity

as some devices or processing conditions require high selectivity between Si and the

masking material. For instance, through silicon vias require extremely long etch times

and subsequently very thick masking layers. In other devices (e.g., those containing

shallow p-n junctions), the extended thermal oxidation times required to grow a

sufficiently thick mask layer may have deleterious effects on the doping profile. Biswas

and Kal [52] showed that the highest selectivity between SiO2 and Si for KOH and

TMAH is found for 10 wt% and 20 wt%, respectively. At an etchant solution

temperature of 80°C (which was found to give the fastest Si etch rate for both etchants),

the Si/SiO2 etch rate ratio is ~400 for KOH and ~5000 for TMAH (Si: ~70 µm/hr, SiO2:

40 nm/hr for KOH and Si: ~37 µm/hr, SiO2: ~7 nm/hr for TMAH). While clearly the

duration of a given etch will be shorter for KOH, the length of thermal oxidation required

to provide a sufficiently thick SiO2 layer will be dramatically decreased if TMAH is

chosen over KOH as

2

4 ( )1 1

2ox

A B tx

A

2.3

Where t, τ, A, and B are

( )ox oxx x A

tB

,

Page 34: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

25

( )i ix x A

B

,

2

2 exp

BA

bC

T

, and

11 exp

bB C

T

,

respectively, and xox, t, and τ are the oxide thickness, time at temperature, T, and time

required to grow initial oxide thickness, xi. C1, C2, b1, and b2 are experimentally

determined coefficients dependant upon the oxidation conditions (i.e., dry, wet, or steam

oxidation) [53, 54].

Etch mask selection is a critical parameter when designing a system for wet

etching. Factors such as cost, ease of deposition, effects on existing components, and

efficacy in performing its designed purpose all need to be considered. In the context of

anisotropic Si etching with KOH or TMAH, two main classes of masking materials exist,

SiO2 and Si3N4, and several routes to deposit them, thermal oxidation, plasma enhanced

chemical vapor deposition (PECVD) and low-pressure chemical vapor deposition

(LPCVD).

Wet etch masks of PECVD-grade SiO2 or Si3N4 have been used, but it is my

personal experience that they do not work very well, with the PECVD layer allowing

etchant solution through discrete holes. Figure 2.4 shows an SEM image of a μ-bar

fabricated on a Si(111) wafer using a process similar to reference [22]. The image shows

a large number of defects associated with KOH etchant attacking the Si surfaces through

pinhole defects. A higher quality etch mask was created by depositing PECVD Si3N4

Page 35: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

26

followed by a short annealing step at 1050°C in an O2 atmosphere. Figure 2.5 shows an

SEM image of a sample using such a masking configuration that successfully protected

the sidewalls of 99.7% of the devices on the wafer. PECVD-grade masking layers can

potentially provide sufficient masking for KOH and TMAH, especially when used in

conjunction with noble metals [16, 55]. Without subsequent processing, such as

annealing or noble metal deposition, PECVD of masking layers is not a reliable means to

KOH or TMAH masking.

Thermally grown oxide serves as a good mask for silicon anisotropic etchants.

Oxidizing a wafer can be as simple as loading samples into a small tube furnace with O2

and/or steam flow. Small tube furnaces as well as small quartz tubes are generally cheap

and provide an accurate and reproducible means to oxidize Si wafers. Thermally grown

oxide films are very dense with few pinholes for etchant attack and provide an ideal

surface for wet etch masking while providing the added benefit of surface passivation

[56] and/or anti-reflection coatings. But thermal oxidation is generally carried out at high

temperature (>900°C), so careful consideration must be taken to ensure previous layers

and doping profiles are not altered deleteriously. Grove, et al. [57] developed a model to

show how dopants redistribute within the Si wafer during thermal oxidation. Dopants

with large diffusivity values in SiO2 can easily diffuse through and to the surface of the

SiO2 layer. Thermal volatization of the dopant into the gaseous phase acts as a sink thus

rapidly depleting dopant concentration in Si. Dopants with large segregation coefficients

m will also be rapidly depleted from Si, where

Page 36: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

27

equilibrium concentration of impurity in Si

equilibrium concentration of impurity in oxidem .

His model suggests that gallium, boron, and indium dopants are depleted rapidly from the

Si wafer, while phosphorous, antimony, and arsenic “pile up” at the Si/SiO2 interface

during thermal oxidation, meaning the impurity concentration is high at this interface

with gradual decay in concentration into the depth of the Si wafer away from the SiO2

interface. These redistribution phenomena should be carefully considered when

developing fabrication protocols requiring thermal oxidation of doped Si wafers.

Low pressure chemical vapor deposition-grade Si3N4 represents an interesting

alternative to the aforementioned masking layers as it provides extremely high selectivity

in KOH, TMAH, and even HF. Williams, et al. [20] reports etch rates for LPCVD Si3N4

in KOH and 10:1 HF to be 0 nm/min and 1.1 nm/min, respectively, while Tabata, et al.

[58] reports zero etch rate in TMAH. Clearly such small etch rates will cause very little

concerns for subsequent processing meaning relatively (compared to thermal oxide or

PECVD-grade oxide or nitride) thin layers can be used to adequately mask devices. In

addition, the deposition temperature (~850°C) will have little effect on the doping profile

as diffusivity, 0

1expD

T

, but will affect metal layers, etc. LPCVD systems are

complex and the gaseous precursors (dichlorosilane for Si3N4) are generally very toxic.

Certain design criteria exist for anisotropically undercut devices which dictate

anchor bar structure and anchor location. One criterion which specifically affects the

required anchor bar width are the differences in selectivity between the Si(110) vs.

Si(111) planes for KOH and TMAH etchants (>100:1 and 50:1, respectively) [59].

Page 37: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

28

Under ideal conditions for KOH-etched systems, the anchor bar width, WAnch, can be

defined as

MinAnchKOH LW 01.0

2.4

due to the >100:1 selectivity between the (110) and (111) planes, but limitations on

photolithography resolution make

MinAnchKOH LW 1.0

2.5

more realistic. Slight misalignments to the (110) plane will cause anchor bar undercut,

even for the highly selective KOH etching scheme. For that reason, a slight buffer of

MinAnchKOH LW 3.0

2.6

sufficiently allows for slight processing errors. As stated previously, selectivity between

the (110) and (111) planes is lower when etched in TMAH. Therefore, a wider anchor

bar of

MinAnchTMAH LW 5.0 2.7

provides a good compromise between wafer utilization and processing error buffer.

Transfer printing of microfabricated devices by undercut etching of a sacrificial

layer is facilitated by conformal contact of the bottom-side of the created device with the

Page 38: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

29

secondary substrate. Reports have been made of roughened Si surfaces following etching

in KOH and TMAH [12, 60-62]. A more dramatic source of roughened surfaces comes

from inherent crystallographic misalignments in the wafer manufacturing. Si(111)

wafers are sawed with a small (~1º) offset to the (111) plane. Following deep etch and

undercut for fabricating devices on a Si(111) wafer, there is a corresponding angular

offset between (110) planes. When undercut etching in TMAH, the bottom-side of the

device is severely scalloped and will require longer etching times to polish [37]. Figure

2.6a and b is an optical image and profilometry line scan, respectively, of Si(111) μ-bars

with Si(111) ±1° fully undercut with TMAH and picked up with a PDMS stamp to show

the bottom-side of the devices and the scalloping associated with the offset of the wafer

face from the (111) plane. Figure 2.6c and d shows similar devices undercut with ±0.1°

offset. These devices are very smooth and facilitate intimate contact of the device to the

receiving substrate.

2.5 Discussion on Wafer Utilization Calculations

This brief section will discuss the assumptions made and derivation of the Wafer

Utilization Plot shown in Figure 2.2.

The term “wafer utilization” is used to describe the amount of active device area,

ADev, on a given wafer, AWafer. The calculation is based on the assumption of a single unit

cell and extrapolated to a full wafer,

Wafer

Dev

A

AnUtilizatio

2.8

Page 39: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

30

where

GapAnchDevWafer AAAA

2.9

and AAnch and AGap are the areas associated with the anchor and etched regions,

respectively. This assumption does not account for losses on wafer edges. The device,

anchor, and gap areas will be treated as:

2LADev

2.10

)2( LWA AnchAnch

2.11

86 LAGap

2.12

The calculations for this plot assume a square device, L1 = L2. The plot is meant

to show the maximum, realistically achievable utilization with the given equipment

available at the Frederick Seitz Materials Research Laboratory at the University of

Illinois at Urbana-Champaign, including photolithography-related tools, plasmas etchers,

photoresists, etc. To that end, the assumed gap between devices is set to 2µm (an

achievable, yet difficult value to realize with the equipment) on all sides of the device.

Page 40: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

31

All systems (homogeneous anchor/isotropically etched, heterogeneous

anchor/isotropically etched, and homogeneous anchor/anisotropically etched) have

slightly different assumptions assigned to them. For a homogeneous anchoring scheme

with an isotropically undercut etched device,

LWAnch 25.1

2.13

A scaling factor of 1.25 was chosen as it represents the lowest, most realistically

achievable value when utilizing a 15% over-etch as discussed in Chapter 2.3, even

though Eqn. 2.2 states that the anchor should be 1.5 times wider than the device (I believe

1.5 is a more practical value, but 1.25 represents the lowest, most realistically achievable

value). For a heterogeneous anchoring scheme with an isotropically undercut etched

device,

0AnchW

2.14

and

44 LAGap

2.15

These equations account for the fact that no wafer material is needed for anchoring and

an extra gap is not needed for patterning the anchor bar, respectively. Finally, for a

homogeneous anchoring scheme with an anisotropically undercut etched device,

Page 41: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

32

LWAnch 1.0

2.16

Again, a scaling factor of 0.1L represents the lowest, most realistically achievable value

while 0.2-0.3 is a more practical value, as discussed in Chapter 2.4.

Page 42: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

33

2.6 Chapter 2 Figures

Figure 2.1: Representative device to be undercut. Design considerations of relative device

and anchor bar widths need to be taken for isotropic and anisotropic undercut etching

systems.

Page 43: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

34

0 200 400 600 800 1000

0

10

20

30

40

50

60

70

80

90

100

Wafe

r U

tili

zati

on

(%

)

Device Length (m)

Isotropic, Homogeneous

Isotropic, Heterogeneous

Anisotropic, Homogeneous

Figure 2.2: Plot showing theoretical wafer utilization for different anchoring strategies in

isotropic and anisotropically undercut etched wafer systems. Assumes 2µm gap and

square devices (L1 = L2).

Page 44: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

35

a) b)

Figure 2.3: Crystallographic projection of a) Si(110) and b) Si(111) planes.

Page 45: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

36

Figure 2.4: SEM image of μ-bars fabricated on a Si(111) wafer using PECVD-grade

Si3N4 as a TMAH etch mask. The image shows a large number of defects associated

with pinholes in the masking layer.

Page 46: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

37

Figure 2.5: SEM image of defect-free μ-bars fabricated on Si(111) wafer using PECVD-

grade Si3N4 followed by short anneal in O2 environment at 1050°C for 5 min as a

sidewall etch mask in TMAH.

Page 47: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

38

8

6

4

2

0

-2

-4

-6

-80 200 400 600 800 1000 1200 1400

Distance (μm)

Tip

Dis

pla

cem

ent (μ

m)

8

6

4

2

0

-2

-4

-6

-80 200 400 600 800 1000 1200 1400

Distance (μm)

Tip

Dis

pla

cem

ent (μ

m)

a) c)

b) d)

Line scan Line scan

Figure 2.6: a) Optical image and b) profilometry line scan of the backside of μ-bars

fabricated from Si(111)±1° wafers after lift-off with a PDMS stamp. c) Optical image

and d) profilometry line scan of similar devices on Si(111) ±0.1°.

Page 48: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

39

2.7 Chapter 2 References

1. Kim, R.-H., et al., Stretchable, Transparent Graphene Interconnects for Arrays of

Microscale Inorganic Light Emitting Diodes on Rubber Substrates. Nano Lett.,

2011. 11: p. 3881-3886.

2. Kim, D.-H., et al., Ultrathin Silicon Circuits With Strain-Isolation Layers and

Mesh Layouts for High-Performance Electronics on Fabric, Vinyl, Leather, and

Paper. Advanced Materials, 2009. 21(36): p. 3703-3707.

3. Park, K.-I., et al., Bendable and Transparent Barium Titanate Capacitors on

Plastic Substrates for High Performance Flexible Ferroelectric Devices.

Electrochem. Solid. St., 2010. 13: p. G57-G59.

4. Park, K.-I., et al., Piezoelectric BaTiO3 Thin Film Nanogenerator on Plastic

Substrates. Nano Lett., 2010. 10: p. 4939-4943.

5. Park, S.-I., et al., Printed Assemblies of Inorganic Light-Emitting Diodes for

Deformable and Semitransparent Displays. Science, 2009. 325: p. 977-981.

6. Yoon, J., et al., GaAs Photovoltaics and Optoelectronics using Releasable

Multilayer Epitaxial Assemblies. Nature, 2010. 465: p. 329-334.

7. Kim, R.-H., et al., Waterproof AlInGaP Optoelectronics on Stretchable Substrates

with Applications in Biomedicine and Robotics. Nat. Mater., 2010. 9: p. 929-937.

8. Lee, J., et al., Stretchable Semiconductor Technologies with High Areal

Coverages and Strain-Limiting Behavior: Demonstration in High-Efficiency

Dual-Junction GaInP/GaAs Photovoltaics. Small, 2012. 8(12): p. 1851-1856.

9. Ahn, J.-H., et al., Bendable Integrated Circuits on Plastic Substrates by use of

Printed Ribbons of Single-Crystalline Silicon. Appl. Phys. Lett., 2007. 90(21): p.

213501-3.

10. Kim, D.-H., et al., Complementary Metal Oxide Silicon Integrated Circuits

Incorporating Monolithically Integrated Stretchable Wavy Interconnects. Appl.

Phys. Lett., 2008. 93(4): p. 044102-3.

11. Ko, H.C., et al., Curvilinear Electronics Formed Using Silicon Membrane

Circuits and Elastomeric Transfer Elements. Small, 2009. 5(23): p. 2703-2709.

12. Baca, A.J., et al., Printable Single-Crystal Silicon Micro/Nanoscale Ribbons,

Platelets and Bars Generated from Bulk Wafers. Adv. Funct. Mater., 2007.

17(16): p. 3051-3062.

13. Baca, A.J., et al., Compact Monocrystalline Silicon Solar Modules With High

Voltage Outputs and Mechanically Flexible Designs. Energy Environ. Sci., 2010.

3: p. 208-211.

Page 49: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

40

14. Kim, H.-s., et al., Unusual Strategies for using Indium Gallium Nitride Grown on

Silicon (111) for Solid-State Lighting. P. Natl. Acad. Sci. USA, 2011. 108: p.

10072-10077.

15. Mack, S., et al., Mechanically Flexible Thin-Film Transistors that use Ultrathin

Ribbons of Silicon Derived from Bulk Wafers. Applied Physics Letters, 2006.

88(21): p. 213101-3.

16. Yoon, J., et al., Ultrathin Silicon Solar Microcells for Semitransparent,

Mechanically Flexible and Microconcentrator Module Designs. Nat. Mater.,

2008. 7: p. 907-915.

17. Ko, H.C., A.J. Baca, and J.A. Rogers, Bulk Quantities of Single-Crystal Silicon

Micro-/Nanoribbons Generated from Bulk Wafers. Nano Lett., 2006. 6(10): p.

2318-2324.

18. Shir, D., et al., Performance of Ultrathin Silicon Solar Microcells with

Nanostructures of Relief Formed by Soft Imprint Lithography for Broad Band

Absorption Enhancement. Nano Lett., 2010. 10(8): p. 3041-3046.

19. Williams, K.R. and R.S. Muller, Etch Rates for Micromachining Processing. J.

Microelectromech. S., 1996. 5(4): p. 256-269.

20. Williams, K.R., K. Gupta, and M. Wasilik, Etch Rates for Micromachining

Processing-Part II. J. Microelectromech. S., 2003. 12: p. 761-778.

21. Goldman, J.L., et al., Strain Anisotropies and Self-Limiting Capacities in Single-

Crystalline 3D Silicon Microstructures: Models for High Energy Density Lithium-

Ion Battery Anodes. Adv. Funct. Mater., 2011. 21(13): p. 2412-2422.

22. Lee, S., S. Park, and D.-i. Cho, The Surface/Bulk Micromachining (SBM)

Process: A New Method for Fabricating Released MEMS in Single Crystal

Silicon. J. Microelectromech. S., 1999. 8: p. 409-416.

23. King, R.R., et al., 40% Efficient Metamorphic GaInP/GaInAs/Ge Multijunction

Solar Cells. Appl. Phys. Lett., 2007. 90: p. 183516.

24. Nakamura, S., S.J. Pearton, and G. Fasol, The Blue Laser Diode: The Complete

Story. 2000 Springer.

25. Green, M.A., et al., Solar Cell Efficiency Tables (Version 39). Prog. Photovoltaics

Res. Appl., 2012. 20(1): p. 12-20.

26. Chen, F., et al., Quantum Confinement, Surface Roughness, and the Conduction

Band Structure of Ultrathin Silicon Membranes. ACS Nano, 2010. 4(4): p. 2466-

2474.

27. Zhang, P., et al., Electronic Transport in Nanometre-Scale Silicon-on-Insulator

Membranes. Nature, 2006. 439(7077): p. 703-706.

28. Fujita, M., et al., Simultaneous Inhibition and Redistribution of Spontaneous Light

Emission in Photonic Crystals. Science, 2005. 308(5726): p. 1296-1298.

29. Yu, J.-K., et al., Reduction of Thermal Conductivity in Phononic Nanomesh

Structures. Nat. Nano., 2010. 5(10): p. 718-721.

30. Tang, J., et al., Holey Silicon as an Efficient Thermoelectric Material. Nano Lett.,

2010. 10(10): p. 4279-4283.

31. Kim, D.-H., et al., Stretchable and Foldable Silicon Integrated Circuits. Science,

2008. 320(5875): p. 507-511.

Page 50: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

41

32. Noell, W., et al., Applications of SOI-based Optical MEMS. Selected Topics in

Quantum Electronics, IEEE Journal of, 2002. 8(1): p. 148-154.

33. Konagai, M., S. Sugimoto, and K. Takahashi, High Efficiency GaAs Thin Film

Solar Cells by Peeled Film Technology. J. Cryst. Growth, 1978. 45: p. 277-280.

34. Yablonovitch, E., et al., Extreme Selectivity in the Liftoff of Epitaxial GaAs Films.

Appl. Phys. Lett., 1987. 51: p. 2222-2224.

35. Petersen, K.E., Dynamic Micromechanics on Silicon: Techniques and Devices.

IEEE T. Electron Dev., 1978. 25(10): p. 1241-1250.

36. Petersen, K.E., Silicon as a Mechanical Material. P. IEEE, 1982. 70: p. 420-456.

37. Oosterbroek, R.E., et al., Etching Methodologies in <111>-Oriented Silicon

Wafers. J. Microelectromech. S., 2000. 9(3): p. 390-398.

38. Palik, E.D., H.F. Gray, and P.B. Klein, A Raman Study of Etching Silicon in

Aqueous KOH. J. Electrochem. Soc., 1983. 130: p. 956-959.

39. Seidel, H., et al., Anisotropic Etching of Crystalline Silicon in Alkaline Solutions.

J. Electrochem. Soc., 1990. 137: p. 3612-3626.

40. Bean, K.E., Anisotropic Etching of Silicon. IEEE Trans. Electron Devices, 1978.

ED-25: p. 1185-1193.

41. Zhao, J., et al., 24% Efficient PERL Silicon Solar Cell: Recent Improvements in

High Efficiency Silicon Cell Research. Sol. Energ. Mat. Sol. C., 1996. 41–42(0):

p. 87-99.

42. Kendall, D.L., Vertical Etching of Silicon at Very High Aspect Ratios. Ann. Rev.

Mater. Sci., 1979. 9: p. 373-403.

43. Shaw, K.A., Z.L. Zhang, and N.C. MacDonald, SCREAM I: A Single Mask,

Single-Crystal Silicon, Reactive Ion Etching Process for Microelectromechanical

Structures. Sensor Actuat. A-Phys., 1994. 40(1): p. 63-70.

44. Li, X.Y., et al. Fabrication of a Single Crystalline Silicon Capacitive Lateral

Accelerometer Using Micromachining Based on Single Step Plasma Etching. in

Micro Electro Mechanical Systems, 1995, MEMS '95, Proceedings. IEEE. 1995.

45. Lee, K.J., et al., Large-Area, Selective Transfer of Microstructured Silicon: A

Printing- Based Approach to High-Performance Thin-Film Transistors Supported

on Flexible Substrates. Adv. Mater., 2005. 17(19): p. 2332-2336.

46. Holt, J.K., et al., Hot-Wire Chemical Vapor Deposition of High Hydrogen

Content Silicon Nitride for Solar Cell Passivation and Anti-Reflection Coating

Applications. Thin Solid Films, 2003. 430(1–2): p. 37-40.

47. Poodt, P., et al., High-Speed Spatial Atomic-Layer Deposition of Aluminum Oxide

Layers for Solar Cell Passivation. Adv. Mater., 2010. 22(32): p. 3564-3567.

48. Aberle, A.G. and R. Hezel, Progress in Low-temperature Surface Passivation of

Silicon Solar Cells using Remote-plasma Silicon Nitride. Prog. Photovoltaics Res.

Appl., 1997. 5(1): p. 29-50.

49. Chahoud, M., H.-H. Wehmann, and A. Schlachetzki, Anisotropic-etching

Simulation of InP and Si. Sensor Actuat. A-Phys., 1997. 63: p. 141-146.

50. Hemenway, B.R., J.E. Bowers, and B.I. Miller, Anisotropic Undercutting ing

(100) Indium Phosphide. Electron. Lett., 1983. 19(1049-1051).

Page 51: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

42

51. Leancu, R., et al., Anisotropic Etching of Germanium. Sensor Actuat. A-Phys.,

1995. 46-47: p. 35-37.

52. Biswas, K. and S. Kal, Etch Characteristics of KOH, TMAH, and Dual Doped

TMAH for Bulk Micromachining of Silicon. Microelectron. J., 2006. 37: p. 519-

525.

53. Deal, B.E. and A.S. Grove, General Relationship for the Thermal Oxidation of

Silicon. J. Appl. Phys., 1965. 36: p. 3770-3778.

54. Anner, G., Planar Processing Primer. 1990: Springer.

55. Madou, M.J., Fundamentals of Microfabrication: The Science of Miniturization.

2002: CRC Press LLC.

56. Mark, J.K. and C. Andres, Very Low Bulk and Surface Recombination in Oxidized

Silicon Wafers. Semicond. Sci. Tech., 2002. 17(1): p. 35.

57. Grove, A.S., O. Leistiko, and C.T. Sah, Redistribution of Acceptor and Donor

Impurities during Thermal Oxidation. J. Appl. Phys., 1964. 35: p. 2695-2701.

58. Tabata, O., et al., Anisotropic Etching of Silicon in TMAH Solutions. Sensor

Actuat. A-Phys., 1992. 34: p. 51-57.

59. Sato, K., et al., Anisotropic Etching Rates of Single-Crystal Silicon for TMAH

Water Solution as a Function of Crystallographic Orientation. Sensor Actuat. A-

Phys., 1999. 73(1–2): p. 131-137.

60. Palik, E.D., et al., Etching Roughness for (100) Silicon Surfaces in Aqueous KOH.

J. Appl. Phys., 1991. 70(6): p. 3291-3300.

61. Zubel, I. and M. Kramkowska, The Effect of Isopropyl Alcohol on Etching Rate

and Roughness of (1 0 0) Si Surface Etched in KOH and TMAH Solutions. Sensor

Actuat. A-Phys., 2001. 93(2): p. 138-147.

62. Divan, R., N. Moldovan, and H. Camon, Roughning and Smoothing Dynamics

During KOH Silicon Etching. Sensor Actuat. A-Phys., 1999. 74(1–3): p. 18-23.

Page 52: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

43

Chapter 3

Unusual Strategies for Using InGaN Grown on Silicon (111) for

Solid State Lighting

The majority of the text and figures included in this chapter is reproduced with

permission from the previously published paper: H.-S. Kim, E. Brueckner, J. Song, Y. Li,

S. Kim, C. Lu, J. Sulkin, K. Choquette, Y. Huang, R. G. Nuzzo, J. A. Rogers, (2011).

"Unusual Strategies for using Indium Gallium Nitride Grown on Silicon (111) for Solid-

State Lighting." P. Natl. Acad. Sci. USA 108: 10072-10077.

3.1 Motivation for Microscale Light Emitting Diodes

The demand for green technologies has pushed solid-state lighting to the forefront

as a cost-saving alternative to low efficiency incandescent light bulbs, with indium

gallium nitride (InGaN) based, blue LEDs holding the dominant position [1, 2]. The

materials and designs for the active components of these devices are increasingly well

developed due to widespread research focus on these aspects over the last one and a half

decades. Internal and external quantum efficiencies of greater than 70% [3] and 60% [3],

respectively, with luminous efficacies larger than 200 lm/W [4] and lifetimes of >50,000

hours [5] are now possible. High luminous efficacy of these LEDs (i.e., 249 lm/W),

compared to that of fluorescent lamps with color-converting, tri-phosphor systems (i.e.,

90 lm/W), is an exciting solution for high efficiency lighting systems [4]. Solid-state

Page 53: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

44

lighting could reduce lighting-related energy consumption in half [1]. Opportunities exist

for further improvements in materials properties to continually drive internal and external

quantum efficiencies, white light efficacy, and heat-sink structures to efficiently manage

generated heat. But their emergence as a ubiquitous alternative for efficient lighting will

rely critically on cost effective methods of integrating all the active components into

packages, interconnecting them into modules, designing efficient heat-sink structures to

effectively manage generated heat during operation, and spatially homogenizing their

light output with desired chromaticity. Current commercial methods for LED die dicing,

assembly, and wiring rely on conceptually old techniques of wafer dicing by laser or

mechanical means in a slow, serial process, robotic systems to assemble the die from the

source wafer, using collections of bulk wires, lenses, and heat sinks in millimeter-scale

packages, on a device-by-device basis, with separate steps to assemble these packages

into large-scale lighting modules [6]. These intrinsically tedious processes are cost

prohibitive and have slowed the breakthrough of LEDs into the common marketplace.

Additionally, such processes are completely unpractical to realize the most appealing

features of LEDs for lighting, such as ultra-small devices, which efficiently and passively

manage smaller amounts of generated heat, which are distributed sparsely but uniformly

over large areas. Such modules could serve as direct replacements for troffers currently

used in fluorescent building lights. Similar techniques of solutions suspensions of LEDs

have been reported [7] but ultimately failed due to incompatibilities with reliable

transferring large numbers of devices reliably.

Page 54: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

45

This section will discuss a set of procedures developed to address some of the

limitations associated with existing technologies of LED fabrication, delineation,

assembly, interconnection, thermal management, and white-light creation. The

procedures developed build off of recent work in flexible electronics [8], information

display [9], and photovoltaics [10, 11], and extend it into the area of solid-state lighting.

New etching strategies, interconnection methods, thermal management techniques, and

schemes for color-conversion are introduced and applied to existing materials for blue

light generation in an effort to create large collections of sparsely assembled arrays of

LEDs.

3.2 Overview of Microscale, InGaN Light Emitting Diode Fabrication

In brief, the process begins with full device fabrication on an epitaxially grown

InGaN-based layer structure on a Si(111) wafer using photolithographic techniques to

pattern areas for ohmic contacts and to define the device geometry suitable for

anisotropic wet etching processes discussed in Chapter 2.4. These devices are amenable

to the aforementioned transfer printing techniques [12] which allow high-throughput

assembly and manipulation of devices that have geometries orders of magnitude smaller

than those compatible with traditional robotic pick-and-place procedures. Large

collections of printed devices are planarized and encapsulated by a remarkably simple,

self-aligned backside exposure technique which exploits the large bandgap of GaN. Such

processing creates distributed sources of illumination over large areas that passively

Page 55: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

46

manage heat generation within the device and are receptive to patterned layers of

phosphor and film-type optical diffusers.

The work focuses on model multilayer InGaN epitaxial stacks grown on Si(111)

wafers [13], due to the cost and throughput advantages that are expected to result from

this materials technology when optimized to offer levels of quality (e.g., threading

dislocation densities < 109 cm

-2) [13] currently available from material grown on

conventional substrates such as sapphire or SiC. The layer configurations appear in

Figure 3.1. As illustrated in Figure 3.2, lithographically patterned n-type ohmic contacts

[14] result from electron beam (e-beam) evaporation and rapid thermal annealing of the

metal stack on regions of n-GaN exposed by inductively coupled plasma reactive ion

etching (ICP-RIE). Similar procedures yield partially transparent p-type ohmic contacts

to the top p-GaN layer, the ohmic characteristics of which are shown in Figure 3.3.

Opaque pads e-beam evaporated on top of the p- and n- contacts enable single-step, self-

aligned planarization, encapsulation, and via formation, using procedures outlined

subsequently. Etching by ICP-RIE (i.e., mesa etch) defines the lateral dimensions of

individual devices, in densely packed, arrayed layouts. Etching proceeds through the

entire thickness of the InGaN material, and to a controlled depth (~1 µm) into the Si for

purposes of release described next. A representative array of such devices appears in

graphic illustration in Figure 3.4a, and in a corresponding SEM image in Figure 3.4c.

Since the active InGaN layers are epitaxially grown on a Si(111) wafer, they can

be released from the host wafer according to procedures outlined in Chapter 2.4. To

properly leverage the anisotropy of the Si/KOH system, the arrays are configured such

Page 56: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

47

that two sides of each device lie perpendicular to <110>. The devices are tightly packed

in this direction (i.e., spacing of 10 µm for this example, but with values as small as 2 µm

are achievable as discussed in Chapter 2.4), and somewhat less so in the orthogonal

direction (i.e., 40 µm shown here). Immersion in hot KOH rapidly removes Si along the

<110> direction which is composed of (110) planes exposed by the mesa etch, thereby

undercutting the devices without etching into the depth of the Si wafer. Because the

etching proceeds only along <110>, relief structures of Si remain in the orthogonal

(<111>) direction between devices. A pair of small supporting structures (i.e., anchors)

of GaN material, also defined during the mesa etch, connects each of the devices to the Si

anchor bars to yield freely suspended configurations after the KOH etching self-

terminates on the (111) planes. A graphical illustration and corresponding SEM image

are shown in Figure 3.4b and Figure 3.4d, respectively. Figure 3.4e and Figure 3.4f show

magnified views of the anchor regions before and after anisotropic Si etching,

respectively. At this stage, the devices can be removed, in a non-destructive, high-speed

parallel operation, using soft stamps and the techniques of transfer printing outlined in

Chapter 1.2 [12]. In this way, assembly into arrayed layouts on glass, plastic, or other

classes of substrate can be achieved at room temperature, with throughputs of millions of

devices per hour and micron-scale positioning accuracy, in deterministic and adjustable

ranges of pitch (Figure 3.5) [9] over areas that can be much larger than those defined by

the devices on the source wafer. The SEM images of Figure 3.4 g-i show a progression

of a representative device from delineation on a donor substrate, to removal, and delivery

onto a receiving substrate, respectively. Figure 3.5 shows a representative array of μ-

Page 57: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

48

ILEDs printed in this fashion, highlighting the capability to transform a densely packed

array of devices configurations on a secondary substrate which are arbitrarily sparse. The

LEDs formed in this manner have emission areas and thicknesses that can be up to 1600x

and 100x smaller, respectively, than conventional devices (i.e., 1 x 1 mm2). For these

reasons, these devices are referred to as μ-scale inorganic light emitting diodes (µ-

ILEDs), following previous reports on different materials systems [9, 15].

The thin µ-ILEDs make them amenable to interconnect based on thin film

metallization, providing a high-speed, parallel alternative to traditional wire bonds.

Large area modules for lighting applications (e.g., troffer-scale) present practical

challenges due to requirements on overlay and registration. Fortunately, the properties of

GaN devices allow a remarkably simple approach to precisely register contact vias

without the need for lithographic alignment. In a process called “back-side exposure”

(BSE), planarization and via formation occur simultaneously in a single-step, self-aligned

process (Figure 3.6). Here, the device structures themselves serve as a mask for photo-

induced cross-linking of a polymer overcoat. Figure 3.7a shows an SEM image of a

single 100 x 100 µm2 µ-ILED printed on a glass substrate. Spin-coating a photosensitive

polymer (Dow Chemical, Benzocyclobutene (BCB), Cyclotene 4024-40 Resin) fully

encapsulates the device (Figure 3.7b). H-line radiation incident on the backside of the

structure passes through the transparent (e.g., glass or plastic) and the GaN (band gap ≈

3.4 eV), to flood expose the polymer in all regions except those shadowed by the opaque

contact pads, shown in colorized gold in Figure 3.7c. Developing the unexposed regions

leaves a pattern of polymer with openings at the contacts, and with positively sloped

Page 58: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

49

sidewalls for conformal deposition of interconnect metal (Figure 3.7d). Due to the

encapsulating nature of the polymer overcoat, requirements on registration for the

interconnects are greatly relaxed compared to those for the contact pads themselves. In

particular, the relevant length scale for registration decreases from roughly the size of a

contact pad to the size of an entire device. This improvement corresponds to a factor of

four for the case considered here with 25 x 25 µm2 contact pads, but could be as large as

a factor of 20 with 5 x 5 µm2 contacts pads. As shown in Figure 3.7d, overly wide leads

(which was found to easily accommodate small misalignments in the printed location of

devices) were used to interconnect arrays by edge-over metallization, photolithographic

patterning, and subsequent metal etching. This method is amenable to interconnecting

large numbers of µ-ILEDs over large area arrays (e.g., 396 µ-ILEDs over ~12 cm2 in

Figure 3.7g), shown here for arrays integrated on PET (Figure 3.7e) and on glass (Figure

3.7e and g) substrates, and for exceptionally small devices. As an example of the latter

capability, ~5 x 5 µm2 vias were formed on devices wither lateral dimensions as small as

25 x 25 µm2 (Figure 3.6d).

To illustrate the versatility, Figure 3.8 a-d show SEM images of exemplary µ-

ILEDs with various sizes from a) 25 x 25 µm2, b) 50 x 50 µm

2, c) 75 x 75 µm

2, and d)

150 x 150 µm2. The sizes of the smallest and largest devices are limited by the resolution

in device processing (i.e., photolithography and mesa etching) and by degradation of

etch-resist layers during Si etching, respectively. The current density-voltage (J-V)

characteristics of these µ-ILEDs show a noticeable increase in J as the size of µ-ILEDs

decreases (Figure 3.8e). This behavior might be attributed to superior current spreading

Page 59: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

50

in small devices [16]. The properties are unaltered by the processing, as shown in Figure

3.8f. The small, thin geometries also provide enhanced mechanical bendability (Figure

3.9) [17] and dramatically improved rates for passive thermal spreading. Both of these

qualities facilitate integration with flexible sheets of plastic, as shown inFigure 3.7e.

3.3 Large Area, White-Light Panels

Gallium nitride-based LEDs have long been used in creating white light, but their

materials’ volume precludes their utility in flexible applications such as those described

herein. In addition, high materials costs makes them ill-suited for applications where

high pixel densities are desired to achieve a uniformly illuminating device, such as back-

lit displays. To demonstrate integrated sources of white light that exploit the unique

capabilities of µ-ILEDs, interconnected µ-ILEDs in spatially significant geometries are

integrated with color-converting phosphors, patterned into small tiles, and thin-film

optical diffusers to create a comparatively larger area of illumination using the same

amount of material of a 1 x 1 mm2 LED.

Fabricating a white light, µ-ILED array follows two parallel routes: (1) µ-ILED

array assembly and interconnection as shown in Figure 3.2 on a thin, plastic substrate,

and (2) generating a separate, patterned phosphor array matching the spatial geometry of

the printed devices. The design of this second sub-module is important because it allows

the use of phosphor only where required, i.e., directly above each of the µ-ILEDs in the

array. These two components are manually aligned such that the µ-ILED substrate (PET

in the example shown here) separates the µ-ILED and phosphor, i.e., phosphor is

Page 60: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

51

illuminated via bottom-side emission from the µ-ILED (interconnects were patterned

such that 90% of the device is covered by reflective aluminum, where the remaining 10%

comprises the separation of p- and n- leads). A schematic representation for processing

patterned phosphor films is shown in Figure 3.10a. Uncured PDMS is cast against an

SU8 master with patterned relief features of controlled height to create “wells” in the

PDMS film. Commercially available cerium-doped yttrium aluminum garnet phosphor

material (Intematix, NYAG-1) is mixed with uncured PDMS to create a uniformly

distributed phosphor slurry (Figure 3.11a). A doctor-blade method is used to “screen

print” fill the relief wells in the PDMS mold, creating the patterned phosphor tiles.

Controlling the chromaticity of emitted light is a necessary component of a

phosphor-integrated system. This study focused on tuning the white light emission

spectra by carefully controlling the phosphor layer thickness at a constant phosphor-in-

PDMS weight loading (37.35 wt%). Color chromaticity data of different phosphor

thicknesses are plotted on a CIE 1931 color space chromaticity diagram in Figure 3.10b.

The data shows that chromaticity can be tuned linearly between the limits of the blue

emission of the µ-ILED and yellow emission of the phosphor with increasing thickness,

achieving CIE coordinates of x=0.321 and y=0.376 with a phosphor thickness of 80 µm,

corresponding to a good quality white light which can be further improved by carefully

selecting LED + phosphor combinations.

A functional white-light, µ-ILED array was created using a printed array of 100 x

100 µm2 devices in a hexagonal pattern in an effort to achieve uniform light emission

over the entire sample area. In this example, 100 µ-ILEDs, which corresponds to the

Page 61: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

52

active material required for a 1 x 1 mm2 LED die, were printed over and area of ~300

mm2 (inter-device spacing of 2 mm), corresponding to an aerial coverage of ~0.3%.

Figure 3.12 highlights the excellent uniformity in electrical performance between

individual rows of printed μ-ILEDs. Figure 3.10 c-e shows a series optical images of the

µ-ILED array on PET without the patterned phosphor film, with the phosphor film soft

laminated to the PET substrate, and a thin, optical diffuser film to achieve uniform light

emission, respectively. By using a sparse array of printing µ-ILEDs, a 100x increase in

effective illuminated area can be realized compared to the area of a traditional LED die,

in a way that has strong optical and thermal benefits.

3.4 Thermal Dissipation in μ-ILEDs

The thermal benefits of the type of layout in Figure 3.10 are critically important,

due to the adverse effects of excessive heating that can occur in devices with

conventional sizes (e.g., 1 x 1 mm2) in the absence of bulk, or miniature, heat sinking

structures [18, 19]. Quantitative study shows that for the sparse, μ-ILED designs, the

electrical interconnects serve simultaneously as effective heat sinks. Both analytical

treatments and rigorous finite element methods (FEM) simulations are used to analyze

the system. For the former, the approximately axi-symmetric nature of the system allows

a precise analytical study of the thermal transport properties. Figure 3.13a shows the

thicknesses H, thermal conductivity k, (the subscripts denote metal interconnect, BCB, μ-

ILED, and glass, respectively), and surrounding temperature T∞. The heat source is

Page 62: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

53

modeled as a disk with a radius r0, and total heat generation Q, which is approximately

equal to the input power to the μ-ILED that does not result in light emission [20]. The

temperature distribution is obtained from the steady-state heat transfer governing

equation 2 2

2 2

10

T T T

r r r z

in cylindrical coordinates (r, z). The boundary conditions

include the free convection d

dm

Tk h T T

z at the top (air-interconnect) surface, and

constant temperature T T at the bottom glass surface, where h is the coefficient of

natural convection. The continuity of temperature and heat flux across interconnect-BCB

interface requires 0T and 0T

kz

, where [] stands for the discontinuity between

two adjacent layers. The above continuity conditions also hold at other interfaces. Heat

generation requires 2

0

T Qk

z r

( 0r r ) across the top and bottom surfaces of μ-ILED.

The interconnect surface temperature is obtained as (see Chapter 3.6 for details)

2

1 2 1 0 00

0 b

e e d2

b mbH HH m

surface

m

kQT r T C C J r J r

r k k h

3.1

where

Page 63: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

54

2

1 b m b b1 1 1 e 1L gH H

g gC k k k k k k

,

2

2 b m b b1 1 1 e 1L gH H

g gC k k k k k k

,

2

b b b b

1

1 1 1 1 e L gH H

g g g gk k k k k k k k

,

2 2 2b b b b

m m m m

1 1 e e 1 e 1m b mH H Hm m

m m

k k h k k h k k

k k h k k h k k

, with

0J and 1J being the Bessel functions of order 0 and 1, respectively. The operating μ-

ILED temperature is given by

2 2

1 02 20b 0

21 e dL gH H

LED

QT T J r

k r

3.2

This analytical treatment agrees well with full three-dimensional FEM simulations as

shown in Figure 3.18. The differences between temperatures in Eqn. 3.1, 3.2, and FEM

simulations are less than 3% for -ILED sizes from 10µm to 100 µm with a 1000 nm-

thick interconnect at a power density 400 W/cm2. The coefficient of natural convection

is h = 25 W/m2/°C [21]. Other conditions in experiments include the surrounding

temperature T = 50°C, thickness and thermal conductivity Hb =1 µm, kb = 0.3 W/m/°C

for BCB [22]; Hg=800 µm, kg = 1.1 W/m/°C for glass [23]; HL = 5 µm for μ-ILED. The

thermal conductivity for Al interconnects is thickness dependent , and is taken as 70

W/m/°C and 160 W/m/°C for 300 nm-thick and 1000 nm-thick interconnects,

Page 64: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

55

respectively. The radius of the disk heat source is r0 = 56 µm to yield the same area as

the square µ-ILED with dimensions of 100 x 100 µm2.

The left and right frames of Figure 3.13 b-g show a set of experiments using

infrared thermal imaging of temperature distributions (QFI Infra-Scope Micro-Thermal

Imager) and analytical predictions, respectively. These experiments compare surface

temperatures for cases of Al interconnects with thicknesses of 300 nm and 1000 nm

(Figure 3.13 b-d for 300 nm and Figure 3.13 e-g for 1000 nm), for input power ranging

from 7.8 mW to 43.2 mW (i.e., power density ranging from 78 W/cm2 to 432 W/cm

2).

Figure 3.13h presents surface temperatures as a function of power, where analytical

model results (lines) agree very well with the experimental measurements (symbols) for

devices with these two interconnect thicknesses.

The results of Figure 3.13 b-h clearly show pronounced decreases in the

temperatures with thicker Al interconnects, thereby demonstrating that the interconnects

themselves serve a dual role as efficient heat sinks by accelerating the rates of lateral

thermal diffusion. These effects can be attributed predominantly to the significant

thermal mass of the interconnects compared to the μ-ILEDs, and to their higher thermal

conductivities. As a consequence, both the thickness of the interconnects and the size of

the devices are important. A theoretical parametric study, summarized in Figure 3.13i,

shows the surface temperatures at a constant heat flux density of 400 W/cm2, as a

function of these two variables. Clearly, the temperature can be greatly reduced by

decreasing the sizes of the LEDs and by increasing the thicknesses of the interconnects.

As a particular example, consider a conventional, macro-size LED (i.e., 1 x 1 mm2) and

Page 65: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

56

an array of 100 μ-ILEDs (i.e., 100 x 100 μm2) at a spacing of 2 mm on otherwise

identical platforms, both at total input power densities of 400 W/cm2. The method of

superposition is used to determine the temperature of μ-ILED arrays based on the

solution for a single LED, i.e., , ,array i

i

T r z T T r z T , where ,iT r z is the

temperature distribution due to ith µ-ILED. The surface temperature distributions for a

macro-size LED and μ-ILED array with spacing 2 mm are shown in Figure 3.14 a-b,

respectively. The maximum temperature occurs at the center of the array and it decreases

with increasing spacing (Figure 3.14c). The conventional LED would reach a

temperature of over 1000°C whereas the array of μ-ILEDs would operate at ~100°C

(Figure 3.18). In real devices, the conventional LED would be completely destroyed

under these conditions, thereby motivating the requirement for advanced heat sinking

structures of the type that are presently in use commercially. By contrast, the μ-ILEDs

experience temperatures that enable stable operation, without any additional components.

3.5 Detailed Processing Protocol for InGaN μ-ILEDs and Interconnected μ-ILED

Arrays

Figure 3.1 shows a schematic illustration of the epitaxial semiconductor

multilayer stack of InGaN MQW LED on a Si(111) wafer. Active layers consist of a Si-

doped n-GaN layer with a thickness of 1700 nm, 5 layers of multi-quantum well (MQW)

of 3 nm InGaN and 10 nm of Si-doped GaN capped with Mg-doped p-GaN layer with a

thickness of 110 nm. This wafer was purchased from Azzurro Semiconductor in

Page 66: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

57

Germany. Figure 3.2 shows a schematic overview of the fabrication process. The

process starts with InGaN epitaxial layers grown on Si (111) wafer, as illustrated in

Figure 3.1. For photolithography, photo-resist AZ5214 was used as both a positive tone

and negative tone resist. The steps for photolithography with this material appear below.

3.5.1 Photolithography using AZ5214-E as a positive tone resist

a. Spin-coat at 4000 rpm for 30 seconds

b. Pre-bake at 110°C for 60 seconds

c. Exposure dose of 78.5 mJ/cm2 at 365nm.

d. Develop in MIF 327 for 35 seconds.

e. Hard bake at 130°C for 180 seconds.

f. O2 descum for 45 seconds in 250 mTorr, 20 sccm of O2 under 50 W.

3.5.2 Photolithography using AZ5214-E as a negative tone resist (Image Reversal)

a. Spin-coat at 5000 rpm for 30 seconds

b. Pre-bake at 110°C for 60 seconds

c. Exposure dose of 110 mJ/cm2 at 320 nm

d. Post-exposure bake (PEB) at 110°C for 65 seconds

e. Flood UV exposure of 400 mJ/cm2

f. Develop in MIF 327 for 35 seconds. More negatively sloped sidewalls can be

achieved for easy lift-off if developed in MIF327 for longer time (i.e.

additional 10~15 seconds).

g. O2 descum for 45 seconds in 250 mTorr, 20 sccm of O2 under 50 W.

Page 67: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

58

3.5.3 Ohmic Contact Fabrication

P-GaN and MQW layers must be etched in the region where n-ohmic contacts are

to be formed. First, n-ohmic contact regions are photo-lithographically defined using

AZ5214-E positive-tone process. Etching the GaN can be achieved using ICP-RIE with

BCl3 and Cl2 gases, with pressures of 3 mTorr and temperatures of 25°C. A two-step

etching process was employed. The first step consisted of 15 sccm of BCl3 with RF

power of 300 W and parallel plate DC voltage of 100 V for 90 seconds. The second step

consisted of 15 sccm of Cl2 gas with RF power of 300 W and parallel plate DC voltage of

100 V for an additional 120 seconds. An etch depth of 350 nm to 400 nm can be

achieved using this recipe. After the ICP-RIE etching of GaN, the photo-resist was

removed using acetone in an ultrasonic bath for about 120 seconds. The total etching

depth was about 350 nm to 400 nm, as measured using profilometry.

Image Reversal of AZ5214-E and lift-off process were used to define n-ohmic

contact metal. The native oxide on the surface n-GaN was removed using BOE at a 10:1

mixing ratio for 120 seconds prior to metal deposition.

(Ti:15nm)/(Al:60nm)/(Mo:20nm)/(Au:100nm) were evaporated at base pressures of 8 x

10-7

Torr as ohmic contacts to the n-GaN. An AG Assoc. Heatpulse 610 RTP was used for

rapid thermal annealing at 860°C for 30 seconds under N2 environment. Ohmic contact

characteristics of Ti/Al/Mo/Au on n-GaN surface are described elsewhere [14].

Image reversal with AZ5214-E was used to define the p-ohmic contact regions.

Immersion of p-GaN in HCl:DI=3:1 for 5 mins removed the native oxide. Ni (10nm)/Au

(10nm) layers were deposited in an e-beam evaporator at a base pressure of <5x10-7

Torr

Page 68: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

59

at a relatively slow rate (approx. 0.5 A/s). After deposition, PR was removed using

acetone in an ultrasonic bath for 120 seconds, and then Ni/Au layers were annealed in a

furnace at 500°C for 10 minutes in air (80% N2 + 20% O2) to improve the ohmic

properties. Ohmic contact characteristics are depicted in Figure 3.3.

Image reversal with AZ5214-E was used to define the opaque contact pad regions

on both p-ohmic contact region and n-ohmic contact region. Opaque contact pads served

not only as contact electrodes, but also as mask patterns for the self-aligned planarization

and encapsulation process, as illustrated in Figure 3.6. Opaque contact pads of

Ti(10nm)/Au(120nm) were deposited using an e-beam evaporator. After deposition, PR

was removed using acetone in an ultrasonic bath for 120 seconds.

3.5.4 Deep Etching

SiN, which served as an etch mask during the KOH undercut process, was

deposited using an STS Multiplex PECVD system. 300 nm of SiN film was deposited at

a pressure of 650 mTorr, temperature of 300oC, and gas flow rates of 1960 sccm (N2) +

40 sccm (SiH4) + 35 sccm (NH3). Mixed frequency RF power of 20 W, with frequencies

of 13.5 6MHz for 6 seconds and 380 KHz for 2 seconds was used.

On top of the SiN film, AZ5214-E was used in an image reversal mode to define

the lateral dimensions of the µ-ILEDs and the geometries of the anchors. Ti(50

nm)/Ni(450 nm) was deposited using an e-beam evaporator at relatively high deposition

rate of approximately 6 Å/sec to minimize the thermal stress caused by the heating inside

the chamber. The SiN film also served as a means to compensate for high tensile stress

within the Ni film. The conditions for SiN deposition are highly optimized to give films

Page 69: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

60

which do peel. If the stress of the Ni film is not properly compensated, the Ni will peel

away from the device surface, destroying the underlying device layers (i.e., ohmic

contacts and SiN layer) resulting in total device failure. Figure 3.15 shows an SEM

image of an array of devices after Ni deposition where the stress was not properly

compensated in the SiN film. After the deposition, PR was removed using acetone in an

ultrasonic bath for 60 seconds.

SiN was dry-etched using a parallel plate RIE (Unaxis/Plasma Therm) with 40

sccm of SF6, 35 mTorr pressure, and 100 W RF power, for an etch rate of SiN of ~100

nm/min. Upon the removal of SiN, GaN/InGaN/AlN/AlGaN epi-layers were all etched

with a gas combination of BCl3/Cl2/Ar in ICP-RIE (Plasma Therm SLR770). Two

etching steps were incorporated for the mesa etch of the GaN-based epitaxial layers. The

first step used 10 sccm BCl3 + 16 sccm Cl2 + 4 sccm Ar at a chamber pressure and

temperature of 5 mTorr and 25ºC, respectively, with ICP coil power of 500 W and

parallel plate voltage of 300 V for 1 minute. The second step used 20 sccm Cl2 + 4 sccm

Ar at a chamber pressure and temperature of 5 mTorr and 25ºC, respectively, with ICP

coil power of 500 W and parallel plate voltage of 260 V for an additional 8 minutes. The

second step etched through the entire thickness of the GaN layer and ~1 µm into the

underlying Si layer.

3.5.5 Undercutting

Following the mesa etch, the sacrificial Si layer is exposed for anisotropic etch

using KOH or TMAH. In this study a KOH-based etchant (Transene PSE-200) was used

to undercut the devices. The etchant was heated in a Pyrex dish at a hot plate temperature

Page 70: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

61

of 100ºC and etching time was ~45 minutes for 100 x 100 µm2 devices. Since GaN has a

relatively high band gap, the device layer of the epitaxial stack is optically transparent

and allows for careful examination of the etch front during undercut. It was discovered

that the location of the anchor greatly affects the undercut profile and the time required

for full undercut. This is thought to originate from exposure of different, slower etching

facets based on the presence of complex or concave corners. Figure 3.16 shows optical

images of devices, highlighting the undercut etch profile, with different anchor locations

on the device. In general, concave corners at the corner of the device result in an

undercut etch profile that exhibit a diagonal, unetched Si structure composed on an

unknown, slow-etching Si plane. However, when the anchors are off-set from the corner

by some distance, the etch front proceeds perpendicular to the side of the device. The

geometry giving the fastest undercut is shown in Figure 3.16a, where the anchor is

centered on the device. This geometry requires ~2x shorter etch times for full undercut,

compared to geometries where both anchors are located on the same side (Figure 3.16b)

or on opposite corners (Figure 3.16c).

Following undercut, the Ti and Ni layers are removed by immersion in Ni etchant

(Transene TFB) which has an etch rate of 3 nm/sec at 25ºC. The SiN layer is etched in

an RIE (Unaxis/Plasma Therm) with 40 sccm of SF6, 35 mTorr pressure, and 100 W RF

power, for an etch rate of ~100 nm/min. This completes fabrication of the donor wafer, a

dense array of undercut devices have been fabricated that are ready for transfer assembly.

Page 71: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

62

3.5.6 Transfer Printing μ-ILEDs

Transfer-printing of µ-ILEDs was carried out onto either glass or PET substrates.

Glass substrates were prepared by cleaving a slide into appropriate dimensions. PET

substrates were prepared by spinning uncured PDMS (10:1 mixture of base to curing

agent) on a glass slide cleaved to appropriate dimensions at 2000 rpm for 30 seconds.

The PET film (Dura-Lar, Grafix) was laminated to the uncured PDMS and the entire

substrate was cured at 70°C for 3 hours. A thin-film adhesive was spin-coated onto the

secondary substrate (glass or PET) after O2 plasma (150 mTorr, 50W, 20 sccm O2 for 1

min) at 3000 rpm for 30 seconds and soft-baked at 110°C for 10 min. Transfer printing

of µ-ILEDs was carried out in automated printer system using PDMS as a stamp. Step

and repeat printing allowed formation of arrays with arbitrary configurations. The thin-

film adhesive was cured under UV light for 10 minutes.

3.5.7 Planarization and Interconnection

To planarize the printed devices and pattern interconnect vias, a single-step, self-

aligned process was developed that exploits the intentional differences in transparency of

the µ-ILED layers (i.e., transparency of GaN at 405 nm, semi-transparency of the thin p-

GaN contacts, and opaque contact pads), called backside exposure (BSE). This process,

while amenable to any class of printed devices which are transparent or non-transparent,

is specially suited to material transparent to wavelengths for photo-induced cross-linking

of the photoresist. BSE through a non-transparent material is still an attractive alternative

Page 72: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

63

as the device will be planarized; a process that can be difficult and time consuming due to

small registration errors of the printed devices during the printing process.

In this embodiment, an ideal photoresist would meet the criteria of a) optically

transparent, b) mechanically flexible, c) electrically insulating, d) strong adhesion to

underlying layers, and e) low processing temperature. Benzocyclobutene (BCB,

Cyclotene 4024-40 Resin) from Dow Chemical was chosen at it meets criterion a-d. The

BSE process is as follows:

a. Adhesion promoter (AP3000) is spin-coated at 2000 rpm for 30 seconds.

b. Soft-baking at 80°C for 30 seconds.

c. BCB (Cyclotene 4024-40, Dow) is spin-coated at 2000 rpm for 60 seconds.

d. Pre-baking at 80°C for 120 seconds.

e. Flood exposure dose of 123 mJ/cm2 at 405 nm from the back side of substrate.

f. Post-exposure baking (PEB) at 70°C for 30 seconds.

g. Develop in DS2100 for 70 seconds.

h. Curing of BCB is carried out in oxygen-free environment at 210°C for 60

minutes.

i. Descum process using RIE at the pressure of 200 mTorr with 18 sccm of O2

with 2 sccm of CF4 with 150 W RF power for 30 seconds.

The high, 210ºC, curing temperature of BCB is difficult due to the lower, ~150ºC,

melting temperature of PET. It was observed experimentally that deformation of the PET

film could be prevented by keeping it laminated to the glass slide/PDMS substrate during

the curing step.

Page 73: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

64

Sputtered or e-beam evaporated Al was used for reflective interconnection.

Aluminum was deposited and patterned photo-lithographically using AZ5214 and an

etch-back process (Type A, Transene). Fully interconnected arrays of µ-ILED resulted

from this metallization process.

3.5.8 Phosphor Integration

Phosphors must be dispersed uniformly to generate uniform white light. Figure

3.11a shows optical microscope images of relief features filled with PDMS/phosphor

slurry (left column) and filled with the phosphor powder only (right column, referred to

as “dry filling”). Compared to the dry filling method, the PDMS/phosphor slurry

provides excellent dispersion and uniformity of phosphor in the PDMS matrix. Emission

spectra of white μ-ILEDs are shown in Figure 3.11b. Below is a detailed protocol for

creating patterned phosphor tiles and their integration onto a sparsely-printed array of μ-

ILEDs.

Fabricating SU-8 master for phosphor-tile mold

a. Spin coat SU8-5 on Si(100) wafer 1800 rpm for 30 sec

b. Bake at 95°C for 5min

c. Flood exposure dose of 216 mJ/cm2 at 365 nm

d. Spin SU8-50 and expose

a. For 60 μm film: Spin 1800 rpm for 30 sec, exposure dose of 432

mJ/cm2 at 365 nm

b. For 80 μm fill: Spin 1600 rpm for 30 sec, exposure dose of 513

mJ/cm2 at 365 nm

Page 74: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

65

c. For 105 μm film: Spin 1250 rpm for 30 sec, exposure dose of 583

mJ/cm2 at 365 nm

e. Bake at 65°C for 1 min then ramp to 95°C, total bake time 11 min

f. Develop in SU8 Developer 12 min

g. Bake 180°C for 10 min

h. UVO treatment for 2 min

i. Treat with tridecafluoro-1,1,2,2-tetrahydrooctyl trichlorosilane for 2 hr in air-

tight container

Creating and Integrating Phosphor Tiles to μ-ILEDs

a. Cast 10:1 mixture (base to curing agent) of uncured PDMS over SU8 master

b. Cure at 70°C for 3 hr

c. Create phosphor/PDMS slurry: Mix 37.35 wt% phosphor in 10:1 PDMS with

glass stir rod

d. Drip small amount of slurry on PDMS mold

e. With PDMS-coated razor blade, squeegee slurry into relief features of PDMS

mold

f. Repeat step e in orthogonal direction

g. Cure at 70°C for 3 hr

h. Phosphor tiles are manually aligned to the back side of a functional μ-ILED

array

Page 75: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

66

3.6 Analytical Model of Printed μ-ILEDs on a Glass Substrate

A half space with built-in disk heat source is used to model the present problem.

The cylindrical coordinate system is set such that the origin is coincident with the center

of the heat source. The steady-state axisymmetric heat conduction in cylindrical

coordinates is

2 2

2 2

10

T T T

r r r z

3.3

Setting T T , where T is the remote temperature, the above equation is equivalent

to

2 2

2 2

10

r r r z

3.4

Boundary and continuity conditions are as follows:

(1) 1g Lz H H h (Glass bottom surface):

1

0g z h

3.5

(2) 0z (BCB-glass interface):

Downward heat flux

Page 76: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

67

1 0

1

0,int 00

, 0

,b

b

z

q r rk G r

q r rz

3.6

Upward heat flux

2 0

g 2

b,int 00

, 0

,

g

z

q r rk G r

q r rz

3.7

Here the heat flux satisfy the following conditions

2 2

1 0 2 0q r q r Q

3.8

00g b zz

,

0 0

b g

0, 0,

gb

z r r z r r

k kz z

3.9

where gk and bk are the thermal conductivities of glass and BCB, 0r is the equivalent

radius of LED and Q is the total heat generated in the LED.

(3) 2bz H h (BCB-interconnect interface):

2 2

b mz h z h

,

2 2

b mb m

z h z h

k kz z

3.10

Page 77: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

68

where mk is the thermal conductivity of the metal interconnect.

(4) b m 3z H H h (Interconnect-air interface):

3

3

m

d

d

mm z h

z h

k hz

3.11

where h is the coefficient of convection at the lower surface of a plate.

Equation 3.4 is solved via the Hankel transform, for which the following

transform pair of the first kind is used,

0

0 )(),(),( drJzzr

3.12

00

, , dz r z J r r r

3.13

where ,r z is the original function and , z is the transform.

The Hankel transform of Eqn. 3.4

2

2

2

d0

dz

3.14

for which the solution is obtained as

Page 78: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

69

e ez zA B

3.15

and the heat flux is

e ez zk k A Bz

3.16

where A and B are two unknown functions to be determined according to boundary and

continuity conditions. The temperature and heat flux are obtained as

00

e e dz zA B J r

3.17

00

e e dz z

zq k k A B J rz

3.18

The boundary and continuity conditions can also be expressed in Hankel transform.

Using Eqn. 3.15-3.18, the two unknowns A and B for each layer can be solved. For

glass,

1

1 0

2

1 2 b 0

11

e

C

g h

J rQA

k r

3.19

Page 79: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

70

12e

h

g gB A

3.20

For BCB,

1 0

b

b b b 0

1 1 11 1

2 2 2

g g

g g

k k J rQA A B

k k k r

3.21

1 0

b

b b b 0

1 1 11 1

2 2 2

g g

g g

k k J rQB A B

k k k r

3.22

For interconnect,

2 22 2b bm b b

m m

11 e 1 e

2

h hk kA A B

k k

3.23

22b bm b b

m m

11 e 1

2

hk kB A B

k k

3.24

where

Page 80: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

71

1

b b

1 1g g

C

k k

k k

,2

b b

1 1g g

C

k k

k k

3.25

3 3

3 3

2 2b b

m m

2 2b b

m m

1 e e 1

e 1 e 1

h hm

m

C

h hm

m

k k h k

k k h k

k h k k

k h k k

3.26

The temperature in each layer can be obtained by Eqn. 3.17. For example, the

temperature in interconnect is given by

00

, e e dz z

m m mT r z T A B J r

3.27

The interconnect surface temperature is then obtained by setting z = h3. The LED

temperature can be approximately by its average value over the entire active region as

12

1 00

0

21 e d

h

LED gT T A J rr

3.28

The thermal conductivity of Al decreases as the film thickness decreases as shown

in Figure 3.17, which is extracted from several references. For the model, the thermal

conductivity of Al is used as a fitting parameter, but with constraints to approximate the

Page 81: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

72

literature values. In the case of 300nm and 1000nm Al interconnects, the thermal

conductivities of 70 W/m/k and 160 W/m/k, respectively, are used in the model. These

values were compared with reported values from the literature to make sure they are

within the reasonable range as depicted in Figure 3.17.

A 3D finite element model is established to study the temperature distribution in

the LED system and validate the analytical model. Eight-node, hexahedral brick

elements in the finite element software ABAQUS are used to discretize the geometry. A

volume heat source is applied on the LED. The thermal convection boundary is applied

at the air-interconnect interface and a constant temperature is applied at the bottom of the

glass substrate. For LED arrays, a ¼ unit cell is used to take advantage of symmetry and

periodic boundaries are applied. The finite element simulations agree well with

analytical modeling as shown in Figure 3.18.

3.7 Discussion on Device Characterization

Electrical measurements were performed with a semiconductor parameter

analyzer (4155C, Agilent or 2400 Sourcemeter, Keithley). Optical measurements of the

emission spectra were performed with a high resolution spectrometer (HR4000, Ocean

Optics). Color chromaticity was determined using SpectraSuite (Ocean Optics) with a

radiometric calibration source (HL-2000, Mikropack) and an Ocean Optics spectrometer

optical fiber in a fixed location, ~1 mm, above the sample. Bending measurements

involved determining the forward voltage needed to produce 10 mA current with the

sample mounted on cylindrical tubes with various radii, ranging from 5.9 mm to 65.3

Page 82: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

73

mm. Fatigue measurements were performed by repeatedly bending the specimen from a

flat state to the bent state with a bending radius of 5.9 mm.

Figure 3.3 shows the ohmic contact characteristics of Ni(10 nm)/Au(10 nm) to p-

GaN using transmission line model measurements (TLM). Contact pads to p-GaN were

patterned according to the process described in Chapter 3.5 with spacings of 2.5 µm, 7.25

µm, 12 µm, and 17 µm under different annealing conditions. Figure 3.3a shows the

current-voltage characteristics of the Ni/Au contact to p-GaN with a TLM pad spacing of

21 µm with four annealing conditions: as deposited, 5 min annealing, 10 min annealing,

and 15 min annealing. Figure 3.3b shows a plot of total resistance at four different pad

spacings, as mentioned above. Due to the large sheet resistance of associated with the

highly resistive p-GaN layer, the specific contact resistance could not be calculated. It is,

however, qualitatively shown that better ohmic characteristics are realized at 10 min

annealing at 500°C over versus 5 min or 15 min annealing at 500°C.

Thermal measurements, as shown in Figure 3.13 and Figure 3.14, of the surface

temperature of µ-ILEDs were performed using MWIR-based InSb thermal imager

(InfraScope, QFI) with the base temperature of 50°C. The printed -ILED is placed on a

heated chuck with a base temperature of 50°C, and pixel-by-pixel calibration is

performed to yield a reference irradiance image of an unpowered sample in order to

account for the emissivity differences on the sample surface. In some cases, however,

when the material has emissivity <0.1, such as Al, temperature measurement could be

inaccurate due to very low thermal emission. A surface ink or polymer that emits as a

blackbody can be placed on top of the sample and to eliminate variation in emissivity

Page 83: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

74

[24]. This procedure was not used because of the destructive nature of this material to

electrical device. As a result, quantitative values were extracted for the temperature only

at the open areas between Al interconnects.

Page 84: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

75

3.8 Chapter 3 Figures

P-GaN (GaN:Mg, 110nm)5x MQW (InGaN/GaN:Si, 3nm/10nm)

N-GaN (GaN:Si, 1700nm)

Multiple LT (AlN:Si/GaN:Si, 1900nm)

(GaN, 750nm)

Seed & Buffer (AlN/AlGaN, 300nm)

Silicon Substrate <111>

Ohmic Contact to P-GaNOhmic Contact to N-GaN

Figure 3.1: Schematic illustration of epitaxial stack of InGaN MQW LED on Si (111)

wafer.

Page 85: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

76

ICP-RIE recessing

n-contact deposition RTA 860ºC 30sec N2

HCl:DI 3:1 surface treatmentp-ohmic contact deposition

RTA 500ºC 10 min

(80% N2 20% O2)

Opaque contact pads deposition

Si3N4 depositionTi/Ni deposition

ICP-RIE deep etch

Lift-off with PDMS stamp

Apply adhesive layerand Transfer-Printing

Step & Repeat

Back-Side Exposure Encapsulation/Planarization

Interconnect metaldeposition

Figure 3.2: Schematic overview of the process for fabricating μ-ILEDs of InGaN on a

Si(111) wafer then transferring and interconnecting them on a secondary substrate.

Page 86: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

77

B

-4 -2 2 40

-100

100

0

Cu

rren

t (A

)

Voltage (V)

As deposited

5 mins

10 mins

15 mins

50

-50

0 5 15 2010

Spacing m

Tota

l Re

sist

an

ce (W

)

10k

20k

30k

40k

50k

60k15 mins

10 mins

0k

a) b)

Figure 3.3: Ohmic contact characteristics of Ni(10 nm)/Au(10 nm) to p-GaN. (a)

Current-voltage characteristics of Ni/Au contact to p-GaN with TLM pad spacing of 21

µm in three different annealing conditions (i.e. As deposited, 5 min, 10 min, and 15 min

annealing). (b) Plot of total resistance at four different pad spacing of 2.5 µm, 7.25 µm,

12 µm, and 17 µm.

Page 87: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

78

50 µm 50 µm 50 µm

200 µm200 µm

50 µm50 µm

4 µm 4 µm

a) b)

c) d)

e) f)

i)h)g)

Figure 3.4: Schematic illustration of arrays of InGaN µ-ILED arrays (a) before and (b)

after anisotropic etching of the near-interfacial region of a supporting Si (111) wafer.

The colors correspond to the InGaN (light blue), the contact pads (gold), and a thin

current spreading layer (red). SEM images of a dense array of µ-ILEDs on a Si (111)

wafer (c) before and (d) after this type of anisotropic etching process. The insets provide

magnified views (colorized using a scheme similar to that in a). SEM images of the

region of the -ILED structure that connects to the underlying silicon wafer (e) before

and (f) after etching. Break-away anchor serve as fracture point during retrieval of µ-

ILEDs from Si (111) wafer. SEM images of a representative µ-ILED, shown in

sequence, (g) after undercut, (h) after removal from the Si wafer, and (i) after assembly

onto a receiving substrate (colorized for ease of viewing).

Page 88: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

79

After KOH Undercut

After Transfer Printing (Donor Substrate)

After Transfer Printing(Receiving Substrate)

150m

150m

400m

a)

c)

b)

Figure 3.5: SEM images of -ILEDs on (a) donor substrate after KOH undercut process,

(b) donor substrate after transfer-printing process, and (c) receiving substrate (i.e. glass)

after the transfer-printing process. -ILEDs are transfer-printed onto a glass substrate

with varying pitches ranging from 25m to 500m.

Page 89: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

80

2m

4m

6m

100 µm

Polymer thickness ~1µm

40m

10 µm

40m

Transparent Substrate

GaN

Transparent Substrate

GaN

Transparent Substrate

GaN

Spin-cast Photosensitive Polymer

Backside UV exposure+

Development

UV

a) b)

c)

d)

e)

Figure 3.6: (a) Schematic illustration of Back-Side Exposure (BSE) process for self-

aligned passivation and via formation. Scanning electron microscopy (SEM) image of a

100 x 100 m2 printed -ILED (b) before, and (c) after BSE process. (d) SEM image of

25 x 25 m2 -ILED after BSE process (colorized for easy of viewing). (e) Cross-

sectional profile of -ILEDs after BSE process.

Page 90: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

81

50 µm

50 µm

50 µm

10mm

10mm

10mm

50 µm

a) e)

b)

f)c)

d)

g)

Figure 3.7: SEM images of the interconnection process for a representative InGaN -

ILED, shown in sequence, (a) after assembly onto a optically transparent substrate (e.g.

glass or plastic), (b) after spin-coating a photo-sensitive polymer, (c) after self-aligned via

formation using a back-side exposure process, and (d) after deposition and patterning of a

metallic interconnect layer The colorized regions correspond to the contact pads (gold), a

thin current spreading layer (red), and Al interconnects (green). Optical images of

various lighting modules based on arrays of µ-ILEDs (e) plastic and (f,g) glass substrates.

Page 91: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

82

2 4 60

— On wafer

— After transfer

Cu

rre

nt

De

nsi

ty (

A/c

m2)

Voltage (V)

0

20

40

60

80

100

400Inte

nsi

ty (

a.u

)Wavelength (nm)

600

50 µm 50 µm

50 µm

2 4 60Voltage (V)

Cu

rre

nt

De

nsi

ty (

A/c

m2)

0

200

400

600

800

1000

— 25m

— 50m

— 75m

— 150m

0

200

400

10-6 10-5 10-4 10-3

LED Area (cm2)

J (A

/cm

2)

50 µm

a) b)

c) d)

e) f)

Figure 3.8: SEM images of arrays of released InGaN µ-ILEDs with dimensions from (a)

25 x 25 µm2, (b) 50 x 50 µm

2, (c) 75 x 75 µm

2 to (d) 150 x 150 µm

2. The colorized

regions correspond to the contact pads (gold), and thin current spreading layers (red). (e)

Corresponding current density-voltage (J-V) characteristics for µ-ILEDs with the

dimensions shown in (a). The inset provides a plot of current density as a function of µ-

ILED area, measured at 6V. (f) Current density-voltage (J-V) characteristics and

emission spectrum (inset) of a representative device before undercut etching on the Si

wafer, and after assembly onto a glass substrate.

Page 92: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

83

20 4

400

800∞— R =

— R = 5.9mm— R = 6.9mm— R = 7.2mm— R = 20.4mm— R = 28.4mm— R = 34.3mm— R = 50.0mm— R = 65.3mm

Voltage (V)

Cu

rre

nt

De

nsi

ty

(mA

/cm

2)

Bending Radius (mm)20 40 600

0

1

2

3

4

Forw

ard

Vo

ltag

e (

V)

0

Bending Cycle

20 4

400

800— 0— 10— 30— 50— 100— 200— 500— 1000C

urr

en

t D

en

sity

(m

A/c

m2)

0

Voltage (V)

200 600 100000

Forw

ard

Vo

ltag

e (

V)

1

2

3

4

800400

a) b)

Figure 3.9: (a) Forward voltage at 10 mA of current and corresponding current-voltage

characteristics (inset) for representative µ-ILEDs printed on a PET substrate measured for

varying bending radii. (b) Forward voltage at 10 mA of current and corresponding

current-voltage characteristics (inset) for representative µ-ILEDs printed on a PET

substrate measured for repetitive cycles.

Page 93: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

84

60 µm80 µm

105 µm

0.4

0.6

0.8

0.2 0.4 0.6

520

500

490

4800.2

0.00.80.0

10mm

1mm

-ILED

PDMS mold

10mm

10mmx

y

a) c)

d)

e)

b)

Figure 3.10: (a) Schematic illustration of the process for fabricating flexible, white

lighting modules, achieved by integrating patterned, encapsulated tiles of YAG:Ce

phosphor islands with arrays of InGaN -ILEDs. (b) Color chromaticity plotted on a CIE

1931 color space diagram for -ILEDs integrated with phosphors with thicknesses of 60

µm, 80 µm, and 105 µm. Optical images of a fully interconnected array of µ-ILEDs (c)

without phosphor, (d) with a laminated film of encapsulated YAG:Ce phosphor islands

(500 x 500 µm2), and (e) with a laminated diffuser film.

Page 94: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

85

500 μm500 μm

99

µm

500 μm500 μm

88

µm

Phosphor/PDMS Slurry

“Dry” Filling

60 µm

80 µm

105 µm

Re

lati

ve In

ten

sity

(a.

u.)

Wavelength (nm)

400 800700600500

a) b)

Figure 3.11: (a) Optical microscopy images of relief features filled with PDMS/phosphor

slurry (left column) and filled with the phosphor powder only (right column). (b)

Emission spectra of white -ILEDs with phosphor layer thickness of 60 m, 80 m, and

105 m.

Page 95: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

86

20

Cu

rre

nt

(mA

)

Voltage per LED (V)

1

2

3

40

1 3

Row 1Row 2Row 3Row 4Row 5Row 6Row 7Row 8Row 9

5 5mm

Row 1

Row 2Row 3

Row 4Row 5Row 6Row 7Row 8Row 9

a) b)

Figure 3.12: (a) Current-voltage (I-V) characteristics of 100 -ILEDs from an array. (b)

Optical image of an array consists of 100 -ILEDs.

Page 96: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

87

0Input Power (mW)

Surf

ace

Te

mp

. (°C

)

20

60

100

140

180

10 20 30 40

Surf

ace

Te

mp

. (°C

)

600

400

200

500°C

0°C

250°C

140°C

120°C

100°C

80°C

60°C

40°C

Glass Hg (kg)

InterconnectHm (km)

LED

BCB Hb

(kb)

2r0

T∞

HL

z

ra)

b)

c)

d)

e)

f)

g)

h) i)

Figure 3.13: (a) Schematic illustration of the device geometry and parameters used in the

analytical model of heat flow. (b-g) Temperature distributions for isolated InGaN µ-

ILEDs with Al interconnects (300 nm and 1000 nm thick for (b-d) and (e-g),

respectively) at input powers of (b) 7.8mW, (c) 16.4mW, (d) 25.2mW, (e) 8.4mW, (f)

18.0mW, (g) 27.6mW captured using a QFI Infra-Scope Micro-Thermal Imager (left) and

calculated by analytical models (right). (h) Surface temperature for -ILEDs with Al

interconnect thicknesses of 300 nm (black) and 1000 nm (red) extracted from

experiments (dots) and computed using the analytical model (lines) as a function of input

power. (i) 3D plot of the surface temperature as function of device size and interconnect

thickness, at a constant heat flux of 400 W/cm2.

Page 97: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

88

100°C80°C60°C-15 -10 -5 0 5 10 15

-15

-10

-5

0

5

10

15

100

200

300

400

500

600

700

800

900

1000

-15 -10 -5 0 5 10 15-15

-10

-5

0

5

10

15

55

60

65

70

75

80

85

90

95

100

105

900°C500°C100°C

5mm5mm

Tem

pe

ratu

re (°C

)

Spacing(m)0 16001200800400

200

400

600

800

1000

1200

20000

a) b) c)

Figure 3.14: Simulated temperature distribution for a a) macrosize LED (i.e., 1 x 1 mm2),

and b) an array of 100 μ-ILEDs (i.e., 100 x 100 μm2) at a spacing of 2 mm. c) μ-ILEDs

surface temperature vs. spacing for an array of 100 μ-ILEDs.

Page 98: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

89

a) b)

Figure 3.15: SEM images showing peeling of the Ni etch mask following the mesa etch

of μ-ILEDs.

Page 99: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

90

a)

b)

c)

Figure 3.16: Optical images of transparent GaN μ-ILEDs during undercut of the

underlying, Si(111) (dark regions). The undercut profile is strongly dependent on the

location of the anchors on the device which influence which crystallographic planes are

exposed. These images represent the same devices under the same undercut etching

conditions with anchors located a) in the center, b) on the same side, and c) opposite

corners.

Page 100: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

91

247W/m/K

Stojanovic et al., 2007 Spina et al., 2006Schmidt et al., 2010Bourgoin et al., 2010Current model

250

Th

erm

al C

on

du

ctiv

ity

(W/m

/K)

200

150

100

50

0

Thickness (nm)

4000 1200800 1600

Figure 3.17: Reported thermal conductivities of thin-film Al from several references.

Page 101: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

92

Te

mp

era

ture

(°C

)

LED Size (m)

0 800600400200

200

400

600

800

1000

1200

1000

0

Te

mp

era

ture

(°C

)

LED size (m)

0 80604020

60

80

100

100

120 Analytical

FEM

Figure 3.18: (a) A plot from analytical results on the surface temperature as function of

LED size up to 1 x 1 mm2. Inset provides comparison between analytical solution and

FEM simulations on the surface temperature as function of LED size ranging from 10 x

10 m2 up to 100 x 100 m

2.

Page 102: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

93

3.9 Chapter 3 References

1. Schubert, E.F. and J.K. Kim, Solid-State Light Sources Getting Smart. Science,

2005. 308: p. 1274-1278.

2. Tsao, J.Y., et al., Solid-State Lighting: An Integrated Human Factors,

Technology, and Economic Perspective. P. IEEE, 2010. 98: p. 1162-1179.

3. Chen, G., et al., Performance of High-Power III-Nitride Light Emitting Diodes.

Phys. Stat. Sol. (a), 2008. 205: p. 1086-1092.

4. Narukawa, Y., et al., White Light Emitting Diodes with Super-High Luminous

Efficacy. J. Phys. D: Appl. Phys., 2010. 43: p. 354002.

5. Solid-State Lighting Research and Development: Multi-Year Program Plan,

U.S.D.o. Energy, Editor 2010.

6. Karlicek Jr., R.F. High Power LED Packaging. in CLEO. 2005. Baltimore, MD.

7. Yeh, H.-J.J. and J.S. Smith, Fluidic Self-Assembly for the Integration of GaAs

Light-Emitting Diodes on Si Substrates. IEEE Photonic Tech. L., 1994. 6: p. 706-

708.

8. Kim, D.-H., et al., Materials and Noncoplanar Mesh Designs for Integrated

Circuits with Linear Elastic Responses to Extreme Mechanical Deformations. P.

Natl. Acad. Sci. USA, 2008. 105: p. 18675-18680.

9. Park, S.-I., et al., Printed Assemblies of Inorganic Light-Emitting Diodes for

Deformable and Semitransparent Displays. Science, 2009. 325: p. 977-981.

10. Yoon, J., et al., Ultrathin Silicon Solar Microcells for Semitransparent,

Mechanically Flexible and Microconcentrator Module Designs. Nat. Mater.,

2008. 7: p. 907-915.

11. Yoon, J., et al., GaAs Photovoltaics and Optoelectronics using Releasable

Multilayer Epitaxial Assemblies. Nature, 2010. 465: p. 329-334.

12. Meitl, M.A., et al., Transfer Printing by Kinetic Control of Adhesion to an

Elastomeric Stamp. Nat. Mater., 2006. 5: p. 33-38.

13. Lee, J., et al., Growth of High-Quality InGaN/GaN LED Structures on (111) Si

Substrates with Internal Quantum Efficiency Exceeding 50%. J. Cryst. Growth,

2011. 315: p. 263-266.

14. Kumar, V., et al., Thermally-Stable Low-Resistance Ti/Al/Mo/Au Multilayer

Ohmic Contacts on n-GaN. J. Appl. Phys., 2002. 92: p. 1712-1714.

15. Kim, R.-H., et al., Waterproof AlInGaP Optoelectronics on Stretchable Substrates

with Applications in Biomedicine and Robotics. Nat. Mater., 2010. 9: p. 929-937.

16. Gong, Z., et al., Size-dependent Light Output, Spectral Shift, and Self-Heating of

400 hm InGaN Light-emitting Diodes. J. Appl. Phys., 2010. 107: p. 013103.

Page 103: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

94

17. Sou, Z., et al., Mechanics of Rollable and Foldable Film-on-Foil Electronics.

Appl. Phys. Lett., 1999. 74: p. 1177-1179.

18. Arik, M. and S. Weaver. Chip Scale Thermal Management of High Brightness

LED Packages. 2004. Bellingham, WA, INTERNATIONAL: Society of Photo-

Optical Instrumentation Engineers.

19. Moon, S.-M. and J.S. Kwak, High-Current Electro-Optical Degradation of

InGaN/GaN Light-Emitting Diodes Fabricated with Ag-Based Reflectors. J.

Korean Phys. Soc., 2009. 55: p. 1128-1131.

20. Senawiratne, J., et al., Junction Temperature Measurements and Thermal

Modeling of GaInN/GaN Quantum Well Light-Emitting Diodes. J. Electron.

Mater., 2008. 37: p. 607-610.

21. Incropera, F.P., et al., Fundamentals of Heat and Mass Transfer. 2007, Hoboken:

Wiley.

22. Christiaens, I., et al., Thin-Film Devices Fabricated with Benzocyclobutene

Adhesive Wafer Bonding. J. Lightwave. Technol., 2005. 23(2): p. 517-523.

23. Lee, S.M. and D.G. Cahill, Heat Transport in Thin Dielectric Films. J. Appl.

Phys., 1997. 81(6): p. 2590-2595.

24. Sarua, A., et al., Combined Infrared and Raman Temperature Measurements on

Device Structures. CS Mantech Conf., 2006: p. 179-182.

Page 104: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

95

Chapter 4

Fabrication of High Performance Silicon Solar Microcells

Integrating Passivation and Antireflection Coatings

4.1 Introduction

Green energy demands have prompted increased development of photovoltaic (PV)

technologies that harvest and convert abundant solar energy into electrical power [1]. To

date, crystalline silicon (single- or poly-crystalline) has been the dominant PV material in

industrial modules (85% share) due to its natural abundance, established processing

routes, and excellent conversion efficiencies [2, 3]. Widespread implementation,

however, is limited by high costs, over 50% of which originates from wafer production

[4].

An interesting approach to reduce materials costs is utilizing ultrathin silicon (5-50

μm, ut-Si) as the active PV material for solar cells [3, 5]. Due to large losses from

wafering [6], kerf-free ut-Si, generated by epitaxial growth [7-11], amorphous Si

deposition and recrystallization [12], and exfoliating Si layers from ingots [13], is very

attractive. Narrow thicknesses of these PV devices now permit relaxed purity

requirements (shorter carrier diffusion lengths) within the Si [14], further improving their

economic viability. These classes of solar cells, however, have challenges, both

Page 105: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

96

fundamental and practical, to overcome. Surface recombination from high surface-to-

volume devices requires careful attention as well as the design of suitable light trapping

schemes to compensate for their thin geometry and poor optical absorption. Additionally,

superstrates and/or transfer methods of these mechanically fragile films must be carefully

chosen that are cost effective and tolerant to the extreme environments in solar cell

processing.

A recent report has shown that thin (2-50 μm), micro-scale (μ-scale) PV devices of

single-crystalline Si can be fabricated via traditional semiconductor processing routes

then, following complete device fabrication, transferred from the donor wafer and

assembled into deterministically sparse arrays on transparent or flexible substrates [15].

This approach is particularly interesting since it (1) overcomes challenges associated with

deleterious handling of thin, fragile semiconductor membranes, (2) removes restrictions

on high temperature processes since full device fabrication is completed on-wafer prior to

transfer, (3) allows for easy integration with secondary optical components (e.g., micro-

lens arrays [15], nanostructured surfaces [16], and luminescent waveguides [17]) to

compensate for inherently low absorption, (4) allows for application-specific distribution

and orientation of devices to create, for example, semitransparent displays or high voltage

modules [18], and (5) is amenable to fabricating multiple generations of devices from a

single Si wafer [15] to decrease material costs. The design of these μ-cells, however,

fails to address fundamental criteria for optimal carrier collection efficiency. While bulk

recombination is reduced due to their thin geometry, surface recombination becomes a

substantial loss mechanism from unpassivated dangling bonds since a passivation layer

Page 106: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

97

(e.g., thermal oxide, silicon nitride) is not incorporated [15]. Additionally, optimization

of the doping profile was not reported. These μ-cells likely suffered from low carrier

lifetime in the highly doped regions (i.e., heavy doping effects) [14]. Finally, their

fabrication protocol uses two costly depositions of Cr/Au which can act as deep level

traps for carriers [19, 20] as well as furnace impurities if the wafers are not cleaned

properly.

Here, by adopting a different doping profile and utilizing robust masking layers, we

report a simplified fabrication process whereby thin, ribbon-like solar cells are fabricated

from a Si wafer. After fabrication, these μ-cells can be assembled onto secondary

substrates with low-cost materials that work synergistically with the Si to harvest solar

radiation. This protocol uses a thermally grown oxide layer as a robust etching and

diffusion mask to bolster fabrication reliability while also doubling as an effective

passivation and anti-reflection layer to further improve device performance. With an

optimized emitter doping concentration, we demonstrate a best cell conversion efficiency

of 11.7% (fill factor (FF) = 0.762) under a simulated AM1.5D solar spectrum. We

present design criteria for assembling printed μ-cells in optimized spatial distributions

and strategies for integrating them with backside reflectors as well as polymer

waveguides that efficiently manage (e.g., reflect and redirect) incident solar radiation. By

increasing the optical path length within the μ-cells and orthogonalizing light to

illuminate once-invisible sidewalls, we leverage the increased surface area of the μ-scale

devices, while mitigating inherent loss mechanisms from them, to dramatically improve

power generation.

Page 107: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

98

4.2 Fabrication of Silicon Solar Microcells

The μ-cell design is shown at the bottom of Figure 4.1a, highlighting the layout of

the phosphorus- (n+, green), and boron- (p

+, blue), and intrinsically (gray) doped regions.

The entire top surface is phosphorus doped to form the emitter (n+) layer, while the back

surface field (BSF, p+) is formed by boron doping the whole bottom surface following

undercut etching. The entire device, except the bottom surface, is covered (i.e. passivated)

by thermal oxide (purple layer) to eliminate dangling bonds and minimize surface

recombination, while at the bottom, minority carrier recombination is reduced by the

built-in electrical field from the BSF [14]. A vertical contact scheme is adopted here as

opposed to Yoon, et al. where both emitter and base contacts are coplanar on the top

surface. This doping geometry enables us to eliminate two photolithography steps from

the previous process that defined the doping regions on top of the device, simplifying the

fabrication process and reducing shading losses from contacts and metal

interconnections.

Figure 4.1a and Figure 4.2 schematically illustrates the key processing steps for

creating the devices. (See Chapter 4.9 for a detailed fabrication protocol.) A Si (111) p-

type, boron-doped, single crystalline Czochralski wafer (resistivity 7-15 Ωcm), which

we refer to as the donor wafer, is first doped by unmasked phosphorous (n+) diffusion

from solid doping sources to form the emitter layer. (See Chapter 4.9.2 and Fig. 4.15-4.17

for optimization of the phosphorus doping condition.) After growth of the top oxide mask

layer (~500nm) by wet oxidation, patterning of this layer by photolithography and deep

etching with inductively coupled plasma reactive ion etching (ICP-RIE) define the lateral

Page 108: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

99

dimensions and thickness of the devices, respectively. For purposes of anisotropic

undercut etching in tetramethylammonium hydroxide (TMAH), the short axes of the

devices are aligned parallel to the <110> direction, exposing the fast etching 110

planes [21] along their sidewalls (Figure 4.1b). The exposed sidewalls of the Si structures

are then briefly etched in a TMAH solution to create a canopy-type structure where the Si

sidewall is recessed ~2.5 μm below the edge of the top oxide layer (Figure 4.3). A

second wet oxidation creates ~250 nm of SiO2 on the sidewalls and trench floor while

increasing the thickness of the top oxide to ~600 nm. An unmasked flood RIE step fully

removes oxide on the trench floor while preserving the sidewall oxide, due in part to the

canopy structure. With the protection of the top and sidewall oxide, the devices are then

released in an anisotropic undercut etch in TMAH yet tethered to the donor wafer in their

lithographically defined locations, connected by break-away anchors (Figure 4.1b) to

structures which are not undercut—long bars whose short axis is orthogonal to the <110>

direction. Utilizing the existing thermal oxide as a diffusion mask, the exposed bottom

surfaces of the μ-cells are doped with a boron solid-source to form the BSF yielding fully

functional devices. SEM images show that a thermal oxide of ~170 nm and ~25 nm

remains on the top and sidewalls of the devices, respectively (Figure 4.4c), serving as

both passivation and AR layers, as discussed later. The fabricated μ-cells are 1.5 mm

long, 100 μm wide and 30 μm thick (Figure 4.1b). Contact pads (50 μm×50 μm) on top of

the device are formed through photolithography, oxide etching, and metal (Ti/Au)

deposition. Devices are now amenable to transferring and assembling into arbitrarily

large and sparse arrays, as shown in Figure 4.5.

Page 109: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

100

4.3 Silicon Solar μ-Cell Characterization

Electrical measurements of individual μ-cells are carried out by transfer-printing

[22] them to a glass substrate and testing in the dark as well as under simulated AM1.5D

illumination of 1000 W∙m-2

at room temperature. Baseline analysis of the solar μ-cell is

performed by mounting the glass substrate on a non-reflective anodized aluminum (AA)

plate to suppress reflection (Chapter 4.9.12) from the testing table, assuming the

illumination area is only the top surface area of the device for current density (J) and

efficiency (η) calculations. Measurements are also performed with additional components

that enhance the PV performance of the μ-cell and as such are differentiated from this

baseline analysis (subscript “base”): A polymeric planarization layer (subscript “PL”), a

diffuse backside reflector (BSR, Labsphere Spectralon® target, subscript “BSR”), or both

(subscript “PL,BSR”), as discussed subsequently.

The inset of Figure 4.6a shows a representative dark curve for an individual μ-

cell, with a diode ideality factor (m) of 1.82. The baseline light curve of our best cell is

presented in Figure 4.6a, with Jsc, open circuit voltage (Voc), fill factor (FF), and η of 28.7

mA∙cm-2

, 0.534 V, 0.762, and 11.7%, respectively. These values all show substantial

improvement from the previous report on Si solar μ-cells [15] and compare favorably

with traditional Si solar cells.

4.4 Thermal Oxide Passivation and Anti-Reflection Coating

A fundamental requirement to efficient transfer printing is releasing the active

device from the donor wafer by undercut etching of a thin sacrificial layer between them.

Page 110: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

101

Previous reports of transfer-printed μ-scale, inorganic semiconductor devices rely on

differential etching rates of disparate materials in the semiconductor stack.

Homogeneous systems, such as that discussed here, have no etch rate distinction between

the active device layer and the donor wafer. Low pressure and plasma enchanced

chemical vapor deposition (LPCVD and PECVD, respectively) SiO2/SiN and/or Au

layers [15, 23-27] have been used as masking structures on the device’s sidewalls to

prevent complete lateral etching across the wafer surface, but thermally grown SiO2 is

known to be denser and less prone to etchant permeability as a masking material [28, 29]

compared to PECVD films and is simpler and potentially more cost effective in this form

of implementation than these previous processes. As shown in Figure 4.7a, the superior

wet-etch masking properties of thermally grown oxide in TMAH allow the creation of

dense and defect-free arrays of devices over large areas. The fabrication yield is greatly

improved for these μ-cells, generally >95%, and primarily limited by imperfections in the

photolithography mask and particulates on the wafer during processing. The conformal

oxide layer coats the top and sidewall surfaces and also serves as a diffusion mask to

protect the p-n junction during the subsequent boron doping step, which is essential to

achieving consistent device performance over large areas. Figure 4.7b and Figure 4.8

shows statistical analyses of PV metrics from μ-cells evenly distributed (i.e., every sixth

device) over the entire fabricated device area. Low standard deviation of all the PV

metrics further confirms the robustness of our process which allows for large collections

of devices to be fabricated over areas governed only by photolithography capabilities and

constrictions in achieving reliable doping profiles over this area.

Page 111: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

102

Two of the key attributes to efficient terrestrial solar cell operation are (1)

efficiently harvesting solar radiation across the entire solar spectrum, and (2)

minimization of parasitic charge carrier recombination. Thermal oxide, in addition to

excellent wet-etch and diffusion masking, also acts as an anti-reflection (AR) coating and

a passivation layer on these silicon solar μ-cells. The AR coating suppresses optical loss

from reflection, while exacerbated surface recombination for these µ-scale devices due to

increased surface area is minimized by effective passivation. This is clearly observed by

simply comparing photocurrent generation with and without the conformal oxide layer.

The J-V plot shown in Figure 4.9a shows a 29.3% decrease (8.2 mA·cm-2

) in Jsc and a

2.8% (0.014 V) decrease in Voc, resulting in a significant drop of η from 10.89% to 7.54%

for a μ-cell after removal of the oxide layer in BOE. This decrease is partially attributed

to increased reflection at the air/Si interface and partially attributed to the lack of

passivation. Results from external quantum efficiency (EQE) measurements support this

conclusion and provide more insights and details, as presented in Figure 4.9b. The μ-cell

with oxide (black curve) shows greater EQE at all wavelengths compared to the μ-cell

without oxide (red curve). This enhancement is especially evident in the blue region

where photons with relatively high energy are absorbed near the surface and generated

charge carriers are lost by recombination when the passivation layer is absent. The AR

effect is highlighted by the peak in the black curve at around 340 nm. (It is worthy to note

that the top oxide thickness was not optimized for optical absorption, but the layer

thickness is tunable for higher absorption at different wavelengths.) Although better AR

coatings can be developed through multi-layer dielectric deposition [30, 31], the single

Page 112: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

103

layer thermal oxide is advantageous in its simplistic integration into the fabrication

protocol and its multifunctionallity—also acting as a robust mask material and

passivation layer—without requirements for additional processing.

4.5 Packaging Silicon Solar μ-Cell on Secondary Substrates

A unique feature of these μ-cells is the ability to deterministically assemble them

into organized and functional layouts via transfer printing techniques. Rapid delivery of

devices onto receiving substrates can be achieved via multiple stamp inking and printing

cycles with automated machines in high throughput. A glass slide, in our case, is used as

the receiving substrate for devices. A UV-curable adhesive is spin-coated on the substrate

then partially cured to remain tacky to mediate the transfer process. For vertically-

contacted devices shown here, electrical contact must be made to the backside of the μ-

cell. For this purpose, bottom electrodes are deposited through a shadow mask onto the

adhesive layer, serving as the electrical interconnect for the p+ layer of the printed μ-cells,

as shown by the schematic and SEM images in Figure 4.10a. A viscoelastic stamp with

pyramid relief features is used to selectively retrieve the devices from the donor wafer

and deliver them to the glass substrate in high transfer yields (100% in the case of Figure

Figure 4.5). (See Chapter 4.9.9 for experimental details.) The configuration shown in

Figure 4.10a yields fully functional Si solar μ-cells and is the configuration for which

device testing in Figure 4.6 and Figure 4.8 was conducted (i.e., without planarization).

Printed μ-cells are planarized in a simplistic approach that eliminates the need for tedious

and precise photolithography. A PDMS block is laminated to the printed array of μ-cells,

Page 113: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

104

creating an air gap between the PDMS and substrate (Figure 4.10 top). NOA is then

flowed by capillary action through the air gap and cured under UV light through the

PDMS block, resulting in planarized μ-cells embedded in a polymer matrix (Figure 4.10

bottom) with low topological variations (Figure 4.11) over very large areas (Figure 4.5).

The intimate contact between the planarization layer and the device (i.e., no air gap) is

highlighted by the inset SEM image in Figure 4.10b bottom inset. The planarization

layer electrically insulates the bottom electrode, thereby rendering the array ready for

top-side (n+) interconnection. Ag paste (Ted Pella, PELCO Conductive Silver 187) is

screen-printed through a manually aligned stencil mask (Figure 4.10c, see Chapter 4.9.11

for experimental details) to achieve a relatively narrow (~150 μm (Figure 4.10c)) and thin

(20-50 μm (Figure 4.12)) electrode.

4.6 Optical Enhancements to Packaged μ-Cells

The system described here—Si solar μ-cells embedded in a transparent polymer

matrix on a glass substrate of comparable refractive indices—inherently creates a simple

geometric concentrator especially when coupled with a BSR, as illustrated in Figure

4.13a. Light incident on the transparent areas between μ-cells can be reflected by the

diffuse BSR and then redirected to the µ-cell through internal reflection within the

planarization layer and glass substrate. Additionally, since µ-cells fabricated here have a

thickness smaller than the optical absorption length of Si for near-infrared light, BSRs

can reflect the transmitted photons and elongate the effective optical path within the

device for improved performance [32]. We use a modified Monte Carlo ray tracing

Page 114: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

105

method that deals with reflection and scattering processes at all interfaces to calculate

photon flux (Φ) incident on the various surfaces of the μ-cell as described in reference

[33] with more details to be published later. As shown in Figure 4.13b, for a planarized

μ-cell on a diffuse BSR, the photon flux, normalized to Φtop, on the top, bottom, and

sidewall surfaces are 1, 0.55, and 0.36, respectively. This data shows that photons are

reflected from the BSR then redirected to the μ-cell’s sidewalls and bottom via the

planarization layer, serving as a simple waveguiding structure. Similar concentrating

structures have been studied for macro-scale solar cells, though the light is waveguided to

the top of the device rather than the sidewall and bottom [34]. Although attaining

maximum concentration is limited by leaky waveguiding and scattering loss, the system

here, compared to lenticular concentrators, has a 180º acceptance angle [34] and utilized

the otherwise un-illuminated sidewalls and bottom of these µ-scale devices for effective

photon capture.

The increased illumination on devices through the planarization layer and diffuse

BSR results in a direct boost of their PV performance. This is shown experimentally in

Figure 4.13c by comparing I-V curves of printed individual μ-cells with different optical

components (i.e., planarization layer and BSR). In all cases Isc increases significantly

with a slight increase in Voc. The trends are consistent with an increasing incident photon

flux onto the cell. As explained subsequently, a μ-cell on an AA plate without

planarization receives illumination on the top surface only, therefore the red curve

(Isc=37.1 μA) is used as the baseline for comparison. Adding a planarization layer or BSR

to the baseline results in a 16.7% increase (blue curve, Isc,PL=43.3 μA) and 63.6%

Page 115: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

106

increase (green curve, Isc,BSR=60.7 μA), respectively. Incorporating both of these optical

components to a μ-cell increases photocurrent generation by 95.7% (Blue curve,

Isc,PL,BSR=72.6 μA). Overall, power generated from the three conditions increases 17.3%,

59.7%, and 101% from the baseline power (13.9 μW). Assuming photocurrent is

proportional to incident photon flux, the concentration ratio achieved here for a single

cell is 1.96. The modest concentration ratio is primarily attributed to light in the

waveguide not satisfying total internal reflection (TIR) criteria.

For a μ-cell in an array layout, the edge-to-edge spacing (s) between the cells

controls available waveguiding area with positive implications on device performance

with increasing s. To optimize s for the highest concentration ratio and maximum power

generation, cells in arrays with varying s are tested. Results of Isc/Isc,0 extracted from

these measurements are shown in Figure 4.13d, where Isc,0 represents photocurrent

generated only by photons incident on the top surface of the μ-cell (i.e., s=0),

approximated by Isc of a non-planarized μ-cell on an AA plate at s=25 μm. On a diffuse

BSR with planarization (black filled squares in Figure 4.13d), Isc/Isc,0 increases with s

because of the expansion of the capture cross-section area for subsequently waveguided

light to the printed μ-cells. Isc/Isc,0 plateaus (maximum value 1.95) at s≈500 μm due to

leaky waveguiding and propagation losses in the NOA matrix. Using optical simulation

as discussed earlier (with modification to accommodate the array layout), the total

incident photon flux (Φtotal) is calculated and normalized to the direct illumination on the

top surface of the μ-cell (Φtop) to determine the relative enhancement (Φtotal/Φtop) at

different s. Φtotal/Φtop is compared directly with experimental Isc/Isc,0 in Figure 4.13d, and

Page 116: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

107

assumes that Isc increases linearly with incident photon flux. The simulation without

parameter fitting shows good agreement with experiment as the enhancement saturates at

a similar value (1.90) at comparable s. The small discrepancy is possibly attributed to

more efficient utilization of photons for these μ-cells when under concentration which is

not accounted for in simulation since light absorption, not photocurrent generation, is

being modeled. Microcells in the control group (measured on an AA plate without

planarization) shows negligible enhancement (Isc/Isc,0 and Φtotal/Φtop ≈1) for all s in both

experimental and simulated results, a result of both the non-reflectivity of the substrate

and absence of the waveguide layer.

4.7 Silicon Solar μ-Cell Module

We use guidance from Figure 4.13d to build an interconnected array of printed μ-

cells with optimized spatial distributions. 10 Si μ-cells were transferred onto a glass

substrate with an edge-to-edge spacing of 500 μm to maximize their power generation,

then planarized and interconnected with strategies described earlier (Figure 4.14a). (It

should be noted that the small-scale array shown here can be, in a straightforward

approach, scaled up to match large area arrays as shown in Figure 4.5.) I-V

characteristics of the 10-cell module measured on a non-reflective AA plate is plotted as

the blue curve in Figure 4.14b, with a Isc,PL of 0.395 mA, Voc,PL of 0.512 V, FFPL of

0.731, and a maximum power (Pmax,PL) of 0.148 mW. The Isc,PL for the module is 8.7 %

lower than the sum of Isc,PL (0.432 mA) of all the individual solar cells prior to

interconnection due to shading losses from the Ag paste, which covers around 10% of the

Page 117: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

108

top surface of the μ-cells. With a diffuse BSR (see red curves in Figure 4.14b), the

module shows an Isc,PL,BSR of 0.711 mA, Voc,PL,BSR of 0.534 V, FFPL,BSR of 0.709, and a

Pmax,PL,BSR of 0.269 mW. Both Isc and Voc are increased due to enhanced illumination on

the μ-cells via waveguiding and reflection from the BSR. Given the cells in this module

only weighs around 0.1 mg (calculated with bulk density of Si and the ideal dimensions

stated previously), the Si consumption is only 0.4 g/W and is much lower than that

(around 10 g/W) of commercial Si solar cells [6], demonstrating the low material

consumption of these miniscule devices. Although costs associated with cell fabrication

and module assembly need to be taken account, the lightweight geometry and sparse

array layout of these µ-scale devices offer a potential route to decrease the cost of PV

energy by reducing the expensive Si component in the module through integration with

and increased photocurrent generation from cheaper optical components.

4.8 Conclusion

In summary, fabrication and assembly (i.e. transfer-printing, planarization, and

interconnection) strategies have been developed for this new generation of Si solar μ-

cells. A simplified fabrication method is demonstrated, which offers improved device

performance as well as higher process reliability compared to previous work. The thermal

oxide on the μ-cells not only acts as an etching and diffusion mask during fabrication, but

serves as a passivation and simple AR coating that enhances the overall energy

conversion efficiency without additional processing for its integration. We report here a

route to integrate active solar μ-cells in a simple concentrator system with low-cost

Page 118: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

109

polymer materials that efficiently manage incident solar radiation by waveguiding light to

otherwise shadowed surfaces. Transfer-printing techniques allow these μ-cells to be

assembled into arbitrary geometries to maximize incident illumination on devices and

target low semiconductor material consumption per watt at the module level. In this

report, a 10-cell module, interconnected by screen-printing Ag paste, demonstrates a Si

consumption of only 0.4 g/W. Important areas of future work include coupling devices

with light trapping structures (e.g. luminescent waveguide) to achieve higher

concentration ratios and lower material consumption, as well as conforming printed μ-

cell arrays on flexible substrates to three-dimensional structures to further improve

angular light capture properties of the module.

4.9 Detailed Processing Protocol for Silicon Solar Microcells

4.9.1 Cleaning

The process starts with a boron-doped Czochralski silicon wafer (“prime” grade,

3-inch diameter, (111) ±0.1° orientation, 7.5-10 Ω∙cm, 380 µm thickness, Silicon

Materials).Wafers are rinsed with acetone and isopropyl alcohol (IPA). Prior to all high

temperature processes, RCA cleaning procedures were used, detailed below:

RCA1 Clean

Etch in RCA1 solution (H2O:NH3∙H2O:H2O2= 5:1:1) at 80°C for 10 min

RCA2 Clean

Etch in RCA1 solution (H2O:HCl:H2O2= 5:1:1) at 80°C for 10 min

Page 119: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

110

Native oxide on the wafer is removed by buffered oxide etchant (BOE 6:1, Transene)

before and after each cleaning step.

4.9.2 Phosphorous Doping

The cleaned wafers are first doped by solid-state sources of phosphorus (PH-

1000N, Saint Gobain). Etching the phosphosilicate glass (PSG) layer in BOE 10:1 and

RCA cleaning completes the doping process.

Phosphorous doping over a variety of temperatures (900°C, 950°C, and 1000°C)

and times (5 min, 15 min, and 30 min) is conducted to ultimately optimize conversion

efficiency of the μ-cells. To directly compare the effect of doping time and temperature

parameters, three batches of three wafers were processed simultaneously, each wafer

within the batch doped for a different time (Batch 1: 1000°C, Batch 2: 950°C, Batch 3:

900°C). To eliminate possible variations in light absorption from anti-reflective, oxide

layers, the µ-cells were immersed in BOE 6:1 to remove all oxide prior to testing. Figure

4.15 shows normalized efficiency vs. doping time for the three temperatures of interest.

The times and temperatures giving the highest conversion efficiency were 1000°C for 5

min, 950°C for 5 min, and 900°C for 15 min, giving sheet resistance values of 22 Ω/, 80

Ω/, and 78 Ω/, respectively, measured by four point probe (Pro4, Signatone). These

conditions were then directly compared in a fourth batch to eliminate batch-to-batch

variability. Figure 4.16a shows the doping profiles measured by quantitative secondary-

ion mass spectrometry (SIMS) and Figure 4.16b shows averaged I-V curves of cells

doped at 900°C, 950°C, and 1000°C for 15 min, 5 min, and 5 min, respectively. Table 4.1

and Figure 4.17 shows the parameters of Jsc, Voc, η, and FF. The data shows that doping

Page 120: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

111

at 900°C for 15 min gives the highest conversion efficiency, leading us to adopt this as

the standard protocol for μ-cell fabrication.

Temp (°C) Jsc (mA/cm2) Voc (V) η (%) FF

900 27.15 0.538 10.6 0.727

950 26.55 0.541 10.0 0.699

1000 23.83 0.543 9.3 0.719

Table 4.1

4.9.3 Thermal Oxidation and Deep Etching

A top oxide mask layer is then grown by wet oxidation at 1000 °C for 100 min

under steam atmosphere (<0.2 LPM O2 flow through boiling DI water) in a quartz tube

furnace following RCA1 and RCA2 cleaning outlined above. The lateral dimensions of

the µ-cells are defined by standard photolithographic processing (AZ5214-E, AZ

Electronic Materials):

Spin-coat hexamethyldisilazane (HMDS) at 5000 rpm for 30 sec

Spin-coat AZ5214-E at 5000 rpm for 30 sec

Softbake at 110°C for 60 sec

Expose 310 mJ/cm2 at 320 nm (Karl Suss MJB3 mask aligner)

Develop in metal ion free (MIF) 327 for ~100 sec

O2 descum (20 sccm, 250 mTorr, 50 W) for 60 sec

Hardbake at 110°C for 180 sec

Page 121: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

112

The oxide not covered by the photoresist is then removed by BOE 6:1 (~12 min).

Inductively coupled plasma reactive ion etching (ICP-RIE, Surface Technology Systems)

defines the vertical depth of the trench structures. Standard Bosch Processing conditions

are used: chamber pressure of 81.5 mTorr, etching for 7 sec/cycle with 130 sccm SF6 and

13 sccm O2 under 600 W coil power and 12 W platen power, then passivating for 5

sec/cycle with 110 sccm C4F8 under 600 W coil power and 0 W platen power. Etch times

were controlled such that an etch depth of 35-40 μm was achieved, typically 75 cycles.

The remaining photoresist is then removed by sonicating in acetone and RCA1 cleaning.

4.9.4 Brief Undercut and Sidewall Oxidation

A short anisotropic etching step of the wafer is performed in 25 wt.%

tetramethylammonium hydroxide solution (TMAH, Sigma-Aldrich) at 70°C for 2 min to

achieve a canopy-type structure. The exposed sidewalls (110) of cells etch at ~1

μm/min according to cross-sectional SEM images shown in Figure 4.3. After RCA

cleaning, a second wet oxidation step (1000 °C, 20-25 min) forms an oxide layer on the

side and bottom of the trenches and increases the oxide thickness on top of the cells as

well, shown in Figure 4.4a.

4.9.5 Flood RIE and Undercut

The oxide on the bottom of the trenches is selectively removed by reactive-ion

etching (Plasma Therm 790 series) using CHF3 (20 sccm) and O2 (8 sccm) under 50

mTorr and 175 W for ~9 min. This dry etch is pseudo-anisotropic such that little oxide

on the sidewall is consumed (Figure 4.4b). After immersion in BOE 6:1 for ~20 sec, the

Page 122: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

113

wafer is immersed into 25% TMAH solution at 80°C for ~90 min to fully undercut the µ-

cells and release them from the substrate except at anchor locations.

4.9.6 Boron Doping

After RCA cleaning, the base layer is formed by solid-source boron doping (BN-

975, Saint Gobain) at 1000 °C for 30 min. The borosilicate glass (BSG) layer is etched in

20:1 HF for ~2 min. Oxide remaining on the top and sidewall surfaces (Figure 4.4c) now

serves as passivation and anti-reflection layers.

4.9.7 Patterning Top Contacts

Square contacts (50 μm ×50 μm) are defined through photolithography with a

photoresist (AZ P4620, Clariant):

Spin-coat HMDS at 2000 rpm for 30 sec

Spin-coat AZ5214-E at 1000 rpm 10 sec, then at 2000 rpm for 30 sec

Softbake at 65°C for 5 min and then at 95°C for 20 min,

Expose 410 mJ/cm2 at 365 nm

Develop in 3:1 AZ400K(AZ Electronic Materials) for ~1 min

Expose under UV-generated ozone for 2 min

Hardbake at 65°C for 30 min (ramp from 35°C with a ramp rate of 0.5°C/min)

The oxide not protected by the photoresist is then removed by BOE 6:1 (~5.5 min) and Ti

(10 nm) /Au (300 nm) contact pads are formed by e-beam evaporation and photoresist

lift-off in acetone.

Page 123: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

114

4.9.8 Statistical Analysis of Fabricated Si Solar μ-Cells

To test the large-area reliability of the fabrication process, we transferred μ-cells

from the entire patterned wafer area. Specifically, every sixth device in all six rows was

picked-up with a PDMS stamp and transferred to a glass substrate (discussed

subsequently) for I-V testing. The μ-cells were tested under simulated AM1.5D

illumination of 1000 W∙m-2

at room temperature prior to planarization of the μ-cells on an

anodized Al plate to minimize reflection from the back plane. Using the top surface area

of the μ-cells as the device area, the PV metrics of Jsc, Voc, η, and FF were measured and

plotted with histograms shown as insets (Figure 4.7 and Figure 4.8).

4.9.9 Transfer Printing μ-Cells

The receiving substrate is made by spin-coating a UV-curable polymer (NOA61,

Norland Products, 3000 rpm for 45 sec) on a pre-cleaned glass slide (5 cm × 5 cm × 2

mm ) and partially curing under UV light (1260 mJ/cm2 at 365 nm) with a Karl Suss

MJB3 mask aligner. Bottom electrodes are then deposited by e-beam evaporation of Ti

(10 nm) /Au (300 nm) through a shadow mask.

Microcells are selectively picked up by a PDMS (10:1 ratio of pre-polymer to

initiator, Sylgard 184, Dow Corning) stamp with pyramidal structures and printed on the

receiving substrate with an automated transfer printing machine. By applying modest

pressure on the inked stamp to the receiving substrate, excellent electrical contact is

achieved between the bottom electrode and device. Step and repeat printing allows

Page 124: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

115

formation of arrays with arbitrary configurations. The fabrication of the pyramid PDMS

stamp and the manipulation of the automated machine are described elsewhere.

After printing, the µ-cells are tested and baked at 150 °C on hotplate for 10 min.

This postbake step possibly causes the partially cured NOA layer to reflow and intimately

coat the bottom side of the μ-cell, preventing liquid pre-polymer to flow underneath the

device and block the back contact in the following planarization step.

4.9.10 Planarization

A flat PDMS block (10:1 ratio of pre-polymer to initiator) is laminated on top of

the printed devices to achieve conformal contact with the µ-cells. Silica particles (30 µm

diameter, SPI) are dispersed on the substrate around the device area to support the PDMS

stamp and avoid sagging. Another UV-curable polymer (NOA73, Norland Products) is

dispensed in a reservoir created in the PDMS block with a 3 mm biopsy punch. The

liquid pre-polymer flows by capillary action to fill the air gap between the PDMS and

substrate. The entire system is then exposed under UV light to cure the NOA, before

PDMS is peeled off from the substrate, leaving the µ-cells embedded in the NOA matrix.

A profilometry line scan (Figure 4.11) of the resulting planarization layer shows minimal

topological variation between cells (~5 μm).

4.9.11 Interconnection

Interconnecting the planarized μ-cells is accomplished by manually aligning a

stencil mask (two pieces of Scotch® tape) to form the shape of the interconnect. A small

volume (~0.5 mL) of Ag paste (Ted Pella, PELCO® Conductive Silver 187) is dispensed

Page 125: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

116

on one end, using a razor blade to “squeegee” the Ag paste over and into the stencil

mask. The Scotch® tape stencil mask is peeled away immediately, leaving behind a

well-formed interconnection line ~150 μm wide and ~50 μm tall (Figure 4.12). The Ag

paste line is cured in air at STP for 30 min prior to any electrical measurements.

4.9.12 Solar μ-Cell Characterization

I-V characteristics of the μ-cells are measured with a source meter (Keithley,

Model 2400) under a full-spectrum solar simulator (Oriel, Model 91192) with AM 1.5D

filter calibrated to 1000 W∙cm-2

at room temperature using a Si reference cell (Newport-

Oriel, Model # 91150V).

Due to the inherent transparency of printed arrays of these μ-cells, reflection from

the subjacent layers will have a significant impact on PV measurements. We suppressed

reflection from the solar simulator’s sample stage by mounting the glass substrate on an

anodized Al (AA) plate to obtain PV metrics with minimal influence of reflected light.

Figure 4.18 shows reflectivity data, relative to a Spectralon® target, of the anodized Al

(AA) plate used during PV testing of some devices. To elucidate enhancements to the

PV metrics of printed μ-cells from diffuse reflection, the glass substrate was also

mounted on a Spectralon® target, in some measurements. We distinguish reported PV

metrics when using this diffuse backside reflector by using subscript “BSR”.

Page 126: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

117

4.10 Future Directions for Silicon Solar μ-Cell Research

At the time of depositing this dissertation, I would consider this Si solar μ-

cell project an “open book”. I think there are a few obvious directions in which

research within this field can proceed.

4.10.1 Three-Dimensional Architectures for Solar Micro-Cell Arrays

The most relevant and ready-to-start direction for this project is building 3D

solar cell architectures for enhanced optical absorption over a large range of incident

angles of irradiation. One of the drawbacks of conventional, flat-plate type solar

cell modules is the inherent angular dependence of absorption. Most first generation

modules—solar cells of single- or multi-crystalline Si—do not use systems which

track the motion of the sun throughout the day. These types of modules are

relatively low efficiency and inexpensive compared to the large cost associated with

a tracking system. Therefore, power generation from these modules is limited,

primarily, to the mid-day hours. Conversely, high-efficiency solar cells with

concentrating optics rely critically on precise focusing and alignment of the sun to

the module. These modules use highly sophisticated tracking systems which add a

significant cost to the overall system. All of these systems rely on wafer-based

materials which are inherently rigid in design. This limits their implementation

primarily to planar geometries.

Using the protocol explained in this dissertation, possibilities are now

available to create architectures which can be mechanically flexed and distorted to

Page 127: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

118

define a geometry matching some predetermined optimal configuration. By

extending the work discussed in this chapter—printing Si solar μ-cells on glass

substrates—to incorporate flexible substrates (e.g., printing onto flexible PET

substrates instead of rigid glass substrates), modules will have to ability to assume

geometries that are unobtainable to traditional, wafer-based materials. For example,

μ-cells could be assembled onto a flexible PET film and interconnected to form a

working array (I don’t foresee any difference in this protocol versus that which has

already been developed for glass substrates). This array can then be mounted

around the periphery of a lens-type structure having a very high numerical aperture,

similar to a drum lens structure. The lens will accept light from a large range of

incident angles, dramatically decreasing the need for active tracking systems.

Future work in this area will be dependent upon optical simulations to guide the

researcher in choosing appropriate module geometries.

4.10.2 Spectral Tuning with Luminescent Solar Concentrators

As has been shown in reference [17] and [35], materials can be incorporated

into the printed μ-cell array which absorbs light of low quantum efficiency then

isotropically re-emits light in a spectral range more suitable for high efficiency

utilization by the solar cell. Work has already begun in this realm and has produced

very promising results [33, 35]. Future work might focus on an implementation

strategy whereby 3D luminescent structures, such as those discussed in Chapter

4.10.1, are created then used as the substrate for subsequent flexible solar cell array

integration. This strategy is promising because current work has already shown

Page 128: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

119

near-zero optical quenching within the luminescent solar concentrator up to very

large inter-cell spacings. This suggests that large optical path lengths, such as those

created with mirrored surfaces, would be amenable to such a materials system.

4.10.3 Silicon Solar Microcell Design Optimization

Although much work has been done to optimize the processing conditions

needed to create the aforementioned Si solar μ-cells, little work has been done to

optimize geometrical parameters of the μ-cell for improved gravimetric power

density. For example, Yoon, et al. [15] showed that efficiency of their μ-cells

started to plateau at thicknesses around 30 μm, stemming from more efficient optical

absorption with thicker geometries. Li [33, 35] has shown via Monte Carlo

simulation that a trade-off exists between increasing absorption of incident radiation

on the top surface and decreasing optical flux upon the sidewall and bottom surfaces

as μ-cell becomes thicker due to waveguiding effects within the planarizing polymer

layer. Lateral dimensions of the cell should also play an important role, one that has

not yet been explored. Detailed study of these parameters, first with Monte Carlo

simulations, might lead to geometries giving additional enhancements to the

module’s performance. There are other features that can be incorporated into the μ-

cell which will enhance performance. Some of these features might include

optimized oxide layer thicknesses and/or surface texturing for light-trapping

purposes or annealing strategies to reduce surface recombination.

Page 129: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

120

4.10.4 Integration of Silicon Solar Microcells into Multi-Junction Devices

Shockley and Queisser first defined the theoretical maximum efficiency of

solar cells and found it to depend strongly on the bandgap of the material [36]. One

route to seemingly bypass this limit is to stack, using wafer-growth methods of

metal organic chemical vapor deposition, etc., multiple, functional solar cells on top

of one another such that each solar cell in the stack utilizes only the spectral range

which results in the highest power generation per incident solar power. But this

process is limited to very expensive wafer-growth techniques in limited form factors

(i.e., wafer-based). The development of robust fabrication protocols and assembling

strategies has given rise to heterogeneous integration of dissimilar materials. To this

end, a route can be easily conceptualized whereby materials of dissimilar bandgaps

can be co-assembled onto substrates. Silicon best utilizes portions of the solar

spectrum centered around ~500-600 nm. A materials set may consist of two III-V

based solar cells, one that efficiently harvests solar radiation above ~600 nm and one

that efficiently operates below 600 nm. This approach gives the distinct advantage

of individually optimizing the materials properties to operate in these spectral ranges

then independently assembling them, as opposed to the traditional method whereby

optimal alloy compositions may require altering to appropriately adjust the lattice

parameters of the layers. Such a constraint can be avoided if each layer is created

independently of one another. This type of work, in fact, has already begun within

Professor Ralph Nuzzo’s group, focused currently on procuring acceptable III-V

solar cell material and developing the appropriate fabrication protocol to create

Page 130: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

121

functional solar cells. A challenging aspect of this work will focus on electrically

interconnecting these geometrically- and electrically-different devices into modules

which are compatible with one another. This work has yet to begin.

Page 131: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

122

4.11 Chapter 4 Figures

Figure 4.1: a) Schematic illustration of important processing steps in fabricating Si solar

μ-cells on a Si(111) wafer with colorized SEM insets of different features throughout the

fabrication process. B) Colorized SEM plan-view image of fully fabricated Si solar μ-

cells (blue) with deposited Au contacts (yellow). The inset is a zoomed in image of the

break-away anchors tethering the devices to the host wafer.

Page 132: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

123

Wet Oxidation

Photolithography

Oxide Etch

ICP-RIE

Shallow Undercut

Flood RIE

Undercut

Phosphorous Doping Boron Doping

Si(111)

Wet Oxidation

Si(111)

n+

PhotoresistSiO2

Si(111)

SiO2

Si(111)

PhotoresistSiO2

Si(111)

SiO2

Si(111)

SiO2

Si(111)

SiO2

Si(111)

SiO2

Si(111)

SiO2

p+

n+Si(111)

Figure 4.2: Schematic illustration of the fabrication protocol for creating Si solar μ-cells.

Page 133: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

124

200 µm

2.5 µm

200 µm

2.5 µm

a) b)

Figure 4.3: Cross-sectional SEM images of Si μ-cells (blue) with a thin layer of

thermally-grown oxide (pink) a) immediately following deep etching via ICP-RIE and b)

following a shallow anisotropic etch in TMAH. The latter highlights the canopy-type

structure that is created following this undercut etch.

Page 134: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

125

200 nm

604 nm

200 nm

355 nm

200 nm

168 nm

200 nm

250 nm

200 nm

208 nm

200 nm

25 nm

a) b) c)

Figure 4.4: Cross-sectional SEM images of Si μ-cells (blue) with thermally grown oxide

layers (pink) on the top (top frame) and sidewall (bottom frame) surface after different

steps in the device fabrication process.: a) sidewall oxidation, b) unmasked, flood RIE

etch, and c) full device fabrication.

Page 135: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

126

Figure 4.5: Optical image of planarized Si solar μ-cells (100 μm x `1500 μm) printed on a

glass substrate over ~600 mm2.

Page 136: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

127

Figure 4.6: Light and dark J-V plot of our best Si solar μ-cell measured on an AA plate.

Inset shows the same dark J-V curve plotted in semi-log form with ideality fit of m=1.82.

Page 137: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

128

Figure 4.7: a) Tilted SEM image of fully fabricated Si solar μ-cells on the host wafer.

These images highlight the high yield obtained from using thermally grown oxide

masking layers, high device density, and low defects over large areas. b) Statistical

analysis and histogram plot (inset) showing efficiency. The measurement was performed

by transferring every sixth device from the entire donor wafer area onto a glass slide

substrate and testing them on an AA plate without planarization under AM 1.5D solar

simulation.

Page 138: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

129

0 10 20 30 40 50 60 70

5

10

15

20

25

30

Js

c (m

A/c

m2

)

Device #

Mean = 26.27

StdDev = 0.375

24 25 26 27 28 290

5

10

15

Co

un

tJsc (mA/cm2

)

0 10 20 30 40 50 60 70

0.40

0.42

0.44

0.46

0.48

0.50

0.52

0.54

Vo

c (

V)

Device #

Mean = 0.5114

StdDev = 0.00432

0.48 0.50 0.52 0.54 0.560

5

10

15

Co

un

t

Voc (V)

0 10 20 30 40 50 60 70

0.0

0.1

0.2

0.3

0.4

0.5

0.6

0.7

0.8

0.9

Fill F

acto

r

Device #

Mean = 0.745

StdDev = 0.0114

68 72 76 800

5

10

15

Co

un

t

Fill Factor (%)

9.5 10.0 10.5 11.00

5

10

15

Co

un

t

Efficiency (%)

0 10 20 30 40 50 60 70

0

2

4

6

8

10

12

Eff

icie

nc

y (

%)

Device #

Mean = 10.01

StdDev = 0.241

a)

c)

b)

d)

Figure 4.8: Statistical analysis and histogram plots (insets) showing key performance

metrics of a) Jsc, b) efficiency, c) Voc, and d) fill factor. The measurements were

performed by transferring every sixth device from the entire donor wafer area onto a

glass slide substrate and testing them on an AA plate without planarization under AM

1.5D solar simulation.

Page 139: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

130

Figure 4.9: a) J-V plot and b) EQE of a Si solar μ-cell (same device used for both

measurements) taken with a passivating oxide layer (black curve) then measured

following removal of the oxide layer.

Page 140: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

131

Figure 4.10: Schematic illustration (top frame) and colorized SEM with inset (bottom

frame) of the important processes used to build interconnected modules of Si solar μ-

cells: a) printing devices onto pre-patterned Au electrodes, b) planarizing with

photocurable Norland Optical Adhesive, and c) patterning Ag paste electrodes.

Page 141: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

132

0 400 800 1200 1600 2000

-20

-15

-10

-5

0

5

10

15

20

H

eig

ht

(m

)

Distance (m)

Figure 4.11: Profilometry line scan across three printed Si solar μ-cells following

planarization. The data shows that planarizing by capillary filling under a PDMS block

results in ~5 μm height variation across 500 μm edge-to-edge spacing between μ-cells.

Page 142: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

133

Figure 4.12: Profilometry line scan across a Ag paste interconnect formed by screen

printing through a stencil mask.

Page 143: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

134

Figure 4.13: a) Schematic cross-section illustration of printed and planarized Si solar μ-

cells on a glass substrate above a diffuse backside reflector. b) Photon flux Monte Carlo

simulation of a printed and planarized Si solar μ-cell on a diffuse BSR. c) I-V plot of Si

solar μ-cells with different optical geometries. d) Experimental (black symbols) and

simulated (red lines) photocurrent at various μ-cell spacing, normalized to photocurrent at

s=0 on AA plate without planarization (Isc/Isc,0 for experimental and Φtotal/ Φtop for

simulation).

Page 144: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

135

Figure 4.14: a) Optical image of a 10 cell module of interconnected Si solar μ-cells using

a screen-printed Ag paste top electrode. b) I-V and P-V plots of the same module

measured on an AA plate (blue lines) and on a diffuse backside reflector (red lines).

Page 145: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

136

Figure 4.15: Efficiency normalized to the maximum efficiency for each data series (i.e.,

doping temperature) vs. phosphorous doping time. The data shows that the maximum

efficiency is achieved at doping times of 15 min, 5 min, and 5 min at doping temperatures

of 900°C, 950°C, and 1000°C, respectively. Processing-induced variations are

minimized between data points in a series by processing all samples within that series

identically, changing only the phosphorous doping time. Three data points for each

series were chosen because only three wafers could be processed simultaneously with the

available equipment.

Page 146: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

137

Figure 4.16: a) SIMS depth profiling of the phosphorous concentration of samples doped

at 900°C for 15 min, 950°C for 5 min, and 1000°C for 5 min. b) Light J-V curves of Si

solar μ-cells doped at 900°C for 15 min, 950°C for 5 min, and 1000°C for 5 min on an

AA plate without planarization under AM 1.5D solar simulation.

Page 147: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

138

Figure 4.17: Using the doping times found to give the highest efficiency for each doping

temperature of interest (900°C for 15 min, 950°C for 5 min, and 1000°C for 5 min) from

Figure 4.15, three wafers were processed identically changing only the doping conditions.

Full device fabrication was carried out, μ-cells were printed on a glass slide substrate then

tested on an AA plate without planarization under AM 1.5D solar simulation. The key

performance metrics of a) Jsc, b) Voc, c) efficiency, and d) fill factor were measured.

Page 148: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

139

Figure 4.18: Reflectivity of an anodized aluminum, relative to a Spectralon® target, used

as the backside optical absorbing medium for solar cell measurements.

Page 149: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

140

4.12 Chapter 4 References

1. Ginley, D., M.A. Green, and R. Collins, Solar Energy Conversion Toward 1

Terawatt. MRS Bull., 2008. 33(04): p. 355-364.

2. Bagnall, D.M. and M. Boreland, Photovoltaic Technologies. Energ. Policy, 2008.

36(12): p. 4390-4396.

3. Wolden, C.A., et al., Photovoltaic Manufacturing: Present Status, Future

Prospects, and Research Needs. J. Vac. Sci. Technol. A, 2011. 29(3): p. 030801-

16.

4. Rodriguez, H., et al., Bulk Crystal Growth and Wafering for PV, in Handbook of

Photovoltaic Science and Engineering. 2011, John Wiley & Sons, Ltd. p. 218-

264.

5. Dross, F., et al., Crystalline Thin-Foil Silicon Solar Cells: Where Crystalline

Quality Meets Thin-Film Processing. Prog. Photovoltaics Res. Appl., 2012. 20(6):

p. 770-784.

6. Aberle, A.G. and P.I. Widenborg, Crystalline Silicon Thin-Film Solar Cells via

High-Temperature and Intermediate-Temperature Approaches, in Handbook of

Photovoltaic Science and Engineering. 2011, John Wiley & Sons, Ltd. p. 452-

486.

7. Bett, A.W., et al. Highest Efficiency Multi-Junction Solar Cell for Terrestrial and

Space Applications. in 24th European Photovoltaic Solar Energy Conference and

Exhibition. 2009. Hamburg, Germany.

8. Bullis, K., Alta Devices Breaks Solar-Cell Record, in Technol. Rev.2011, MIT.

9. Keavney, C.J., V.E. Haven, and S.M. Vernon. Emitter Structures in MOCVD InP

Solar Cells. in Photovoltaic Specialists Conference, 1990., Conference Record of

the Twenty First IEEE. 1990.

10. King, R.R., et al., 40% Efficient Metamorphic GaInP/GaInAs/Ge Multijunction

Solar Cells. Appl. Phys. Lett., 2007. 90: p. 183516.

11. Venkatasubramanian, R., et al. 18.2% (AM1.5) Efficient GaAs Solar Cell on

Optical-Grade Polycrystalline Ge Substrate. in Photovoltaic Specialists

Conference, 1996., Conference Record of the Twenty Fifth IEEE. 1996.

12. Green, M.A., et al., Crystalline Silicon on Glass (CSG) Thin-Film Solar Cell

Modules. Sol. Energy, 2004. 77(6): p. 857-863.

13. Dross, F., et al., Stress-Induced Large-Area Lift-Off of Crystalline Si Films. Appl.

Phys. A Mater., 2007. 89(1): p. 149-152.

14. Gray, J.L., The Physics of the Solar Cell, in Handbook of Photovoltaic Science

and Engineering. 2011, John Wiley & Sons, Ltd. p. 82-129.

Page 150: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

141

15. Yoon, J., et al., Ultrathin Silicon Solar Microcells for Semitransparent,

Mechanically Flexible and Microconcentrator Module Designs. Nat. Mater.,

2008. 7: p. 907-915.

16. Shir, D., et al., Performance of Ultrathin Silicon Solar Microcells with

Nanostructures of Relief Formed by Soft Imprint Lithography for Broad Band

Absorption Enhancement. Nano Lett., 2010. 10(8): p. 3041-3046.

17. Yoon, J., et al., Flexible Concentrator Photovoltaics Based on Microscale Silicon

Solar Cells Embedded in Luminescent Waveguides. Nat. Commun., 2011. 2: p.

343.

18. Baca, A.J., et al., Compact Monocrystalline Silicon Solar Modules With High

Voltage Outputs and Mechanically Flexible Designs. Energy Environ. Sci., 2010.

3: p. 208-211.

19. Sah, C.-T., Fundamentals of Solid State Electronics. 1991: World Scientific

Publishing Co.

20. Tavendale, A.J. and S.J. Pearton, Deep Level, Quenched-In Defects in Silicon

Doped with Gold, Silver, Iron, Copper or Nickel. J. Phys. C Solid State, 1983.

16(9): p. 1665.

21. Tabata, O., et al., Anisotropic Etching of Silicon in TMAH Solutions. Sensor

Actuat. A-Phys., 1992. 34: p. 51-57.

22. Meitl, M.A., et al., Transfer Printing by Kinetic Control of Adhesion to an

Elastomeric Stamp. Nat. Mater., 2006. 5: p. 33-38.

23. Lee, S., S. Park, and D.-i. Cho, The Surface/Bulk Micromachining (SBM)

Process: A New Method for Fabricating Released MEMS in Single Crystal

Silicon. J. Microelectromech. S., 1999. 8: p. 409-416.

24. Lee, S., S. Park, and D.-i. Cho, A New Micromachining Technique with (111)

Silicon. Jpn. J. Appl. Phys., 1999. 38: p. 2699.

25. Park, S., et al., Mesa-Supported, Single-Crystal Microstructures Fabricated by

the Surface/Bulk Micromachining Process. Jpn. J. Appl. Phys., 1999. 38: p. 4244.

26. Ko, H.C., A.J. Baca, and J.A. Rogers, Bulk Quantities of Single-Crystal Silicon

Micro-/Nanoribbons Generated from Bulk Wafers. Nano Lett., 2006. 6(10): p.

2318-2324.

27. Baca, A.J., et al., Printable Single-Crystal Silicon Micro/Nanoscale Ribbons,

Platelets and Bars Generated from Bulk Wafers. Adv. Funct. Mater., 2007.

17(16): p. 3051-3062.

28. Madou, M.J., Fundamentals of Microfabrication: The Science of Miniturization.

2002: CRC Press LLC.

29. Yan, G., et al., An Improved TMAH Si-Etching Solution Without Attacking

Exposed Aluminum. Sensor Actuat. A-Phys., 2001. 89(1–2): p. 135-141.

30. Schultz, O., S.W. Glunz, and G.P. Willeke, Multicrystalline Silicon Solar Cells

Exceeding 20% Efficiency. Prog. Photovoltaics Res. Appl., 2004. 12(7): p. 553-

558.

31. Zhao, J., et al., 24% Efficient PERL Silicon Solar Cell: Recent Improvements in

High Efficiency Silicon Cell Research. Sol. Energ. Mat. Sol. C., 1996. 41–42(0):

p. 87-99.

Page 151: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

142

32. Yablonovitch, E. and G.D. Cody, Intensity Enhancement in Textured Optical

Sheets for Solar Cells. IEEE T. Electron Dev., 1982. 29(2): p. 300-305.

33. Li, L., et al. Modeling and Optimization of Luminescent Solar Concentrator for

Micro Cell Array Module. 2012. Optical Society of America.

34. Smestad, G. and P. Hamill, Concentration of Solar Radiation by White Backed

Photovoltaic Panels. Appl. Opt., 1984. 23(23): p. 4394-4402.

35. Li, L., Ph.D Dissertation, in Materials Science and Engineering2012, University

of Illinois at Urbana-Champaign.

36. Shockley, W. and H.J. Queisser, Detailed Balance Limit of Efficiency of p-n

Junction Solar Cells. J. Appl. Phys., 1961. 32(3): p. 510-519.

Page 152: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

143

Chapter 5

Nanotextured and Anti-Reflective PDMS

5.1 Introduction

Reflective losses from the surface of high refractive index semiconductors has

long been known [1, 2] to serve as an impediment to achieving high photovoltaic

conversion efficiency for solar cells. Reflection from the surface of a material of known

refractive index n is governed by Fresnel’s equations:

2

21

21

coscos

coscos

ti

ti

snn

nnR

5.1

for s-polarized light, and

2

21

21

coscos

coscos

it

itp

nn

nnR

5.2

for p-polarized light where R and θ are the reflection coefficient and angle of incidence

normal to the sample surface for incident (i) and refracted (t) light. Unpolarized light

through air (n=1) incident on silicon (n=3.42) exhibits R=30% for θi = 0~60⁰.

High reflective losses from Si and other optical media demands methodologies to

circumvent this problem, the so-called “anti-reflective coating” (ARC). Commercial

examples of ARC have predominately relied upon texturing of the semiconductor surface

Page 153: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

144

[2-4] or single [5] or multilayer stacks [6-8] of dielectrics, where n1 < nARC < n2. Reflection

coefficient for a single film is

hknnnhknnn

hknnnhknnnR

ARCARC

ARCARCARC

0

222

210

22

21

2

0

222

210

22

21

2

sincos

sincos

5.3

and reduces to

22

21

22

21

ARC

ARCARC

nnn

nnnR

5.4

for ARC optical thickness h=1/4λ. RARC = 0 when 21

2 nnnARC . Therefore, 1/4 λ

dielectric layers or multilayers have found wide acceptance as ARC for semiconductor

materials. A host of deposition and growth methods have been employed to fabricate

ARC, some of which are PECVD deposition of SiN [6, 7], sputtering MgF2 [8], and

others have been used and well-documented in literature.

As external quantum efficiencies of the semiconductor materials in PV modules

continues to improve and approach the Shockley-Quissar Limit [9], efficiency

enhancements from further optimization of the semiconductor device becomes restricted.

Therefore extremely good optical response of all components within the module become

paramount in importance, namely transmission and transmission over tens of years in

harsh environmental conditions of polymeric encapsulants and focusing optics. To

address the latter point, silicone encapsulants have taken an increasingly dominant role as

Page 154: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

145

their chemical and thermal stability far surpasses previous encapsulants of EVA and other

epoxy-based materials [10]. Under prolonged UV light exposure, these materials assume

a yellow hue, dramatically degrading transmission characteristics [11, 12]. The

aforementioned techniques for ARC fabrication cannot be used as deposition

temperatures are too high and/or the condition of n1 < nARC < n2 does not hold due to

generally low (1.3~1.5) refractive indices of polymers. Therefore more exotic

methodologies must be adopted to suppress reflection at the polymer/air interface.

Bernhard [13] first observed that the surface of the corneal lenses of moths

displayed a regular array of conical protrusions measuring roughly 200 nm in height and

spacing. His hypothesis, confirmed by scaled up features active in the microwave region,

stated that these protrusions aided in reducing reflection by acting as a graded change in

refractive index from air to medium. Subsequent studies were able to reproduce visible

wavelength-active features by recording the straight line interference fringes of a krypton

laser in a photoresist-coated glass slide to produce features roughly 210 nm in spacing

[14]. Further studies conducted in this field have created similar arrays by holographic

projection on photoresist-coated lenses [15], spin-coated colloidal silica sphere etching

masks [16], nanoimprinting [17], and electron beam lithography [18]. Most of these

experimental procedures rely on extremely high resolution photolithographic patterning

of sub-wavelength features, often requiring electron beam lithography, and controlled

etching of underlying materials. Moreover, these methods do not address the critical

issue of fabricating moth-eye structures on bulk silicone substrates, especially in single-

step processes that do not require molding to a pre-patterned master.

Page 155: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

146

The Gogolides group has developed an approach to creating nanoscale columnar

structures on PDMS surfaces with SF6 plasma in an inductively coupled plasma reactive

ion etcher (ICP-RIE) [19]. Their reported process used PDMS coated glass slides etched

in an ICP-RIE without the use of photolithographic pattering or any prior surface

treatments. This single-step protocol resulted in columnar structures 1.45 µm tall with a

~230 nm periodicity, and achieving columns as tall as 6µm, but without mention of

feature periodicity, in an effort to create super hydrophobic, self-cleaning surfaces.

Later studies from this research group focused on the mechanism responsible for

the creation of these nanocolumnar features [20]. Poly(methyl methacrylate) (PMMA)

films were etched in an ICP-RIE with different chamber materials and coatings to

examine the possible effects of chamber wall sputtering and re-deposition onto the

samples serving as a hard mask to plasma etching. The ICP-RIE used in those

experiments consisted of an alumina dome and an anodized Al ring. Examining etched

PMMA samples via x-ray photoelectron spectroscopy (XPS), the authors observed an

increasing concentration of Al for longer etch times. They hypothesized that Al particles

were being sputtered from the chamber walls and re-deposited onto the PMMA samples

serving as a hard mask to subsequent plasma etching. Subsequent experiments were

conducted to examine the surface topology after coating the chamber walls and/or

replacing components to effectively eliminate Al sources from the chamber. Results

show a dramatic decrease in surface roughness following coating of the alumina dome

and Al ring with a photoresist polymer (RMS roughness of 100 nm with native ICP-RIE

chamber, 3.4 nm for polymer coated dome and ring). These findings strongly suggest

Page 156: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

147

influence from the ICP-RIE chamber walls as a source of nanoscale, hard masking

materials.

With strong evidence from the aforementioned research group, along with other

sources citing nanoscale texturing of PDMS [21-23] and other silicones [24, 25] after

fluorine-based plasma etching for the purposes of achieving superhydrophic surfaces, I

believe there is a great opportunity to exploit this nanoscale phenomenon for creating an

anti-reflection layer on silicone surfaces.

5.2 Experimental

5.2.1 Materials

Commercial PDMS (Sylgard 184, Dow Corning) was used as the substrates,

mixed at 10:1 ratio by weight (60 g base, 6 g curing agent). Uncured PDMS was poured

into a 150 mm petri dish, de-gassed under vacuum for 30 min, then cured at 70⁰C for at

least 8 hrs. After fully cured, 1” x 1.25” rectangular samples were cut with a razor blade.

No cleaning steps were performed to the PDMS samples.

5.2.2 Plasma Etching

PDMS samples were etched in a reactive ion etcher (RIE, March CS-1701) under

varying conditions to examine the process-dependent topology of the etched surface.

Preliminary experiments held process pressure and base pressure constant at 200 mTorr

and 100 mTorr, respectively. Etching power, time, and gas composition were varied

from 100 W – 400 W, 1 min – 7 min, and 10 sccm CF4 or 5 sccm CF4 / 5 sccm O2,

Page 157: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

148

respectively, for six different batches of samples. Every batch contained five samples,

one sample of native PDMS and four samples to be etched. In all cases, the native

PDMS sample was extracted from the same petri dish as the other four samples to reduce

thickness and PDMS composition variations. A small (~5mm x ~5mm) PDMS sample

was inserted in the RIE with the 1”x1.25” sample to be used to examine surface topology

via scanning electron microscopy (SEM). Table 5.1 shows process conditions for each

batch.

Process Pressure Power Time Gas

Batch 1 200 mTorr

100W, 200W,

300W, 400W

3 min 10 sccm CF4

Batch 2 200 mTorr 200 W

1min, 2min,

4min, 7min

10 sccm CF4

Batch 3 200 mTorr 300 W

1min, 2min,

4min, 7min

10 sccm CF4

Batch 4 200 mTorr

100W, 200W,

300W, 400W

3 min

5 sccm CF4 +

5 sccm O2

Batch 5 200 mTorr 200 W

1min, 2min,

4min, 7min

5 sccm CF4 +

5 sccm O2

Batch 6 200 mTorr 300 W

1min, 2min,

4min, 7min

5 sccm CF4 +

5 sccm O2

Table 5.1

Page 158: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

149

5.2.3 Optical Transmission Characterization with UV-Vis Spectrophotometry

Optical transmission of RIE etched PDMS samples was characterized in a UV-

Vis spectrophotometer (Agilent, Cary 5000) fitted with diffuse reflectance attachement.

The light source was powered on and allowed to warm up for 45 min to ensure consistent

bulb intensity throughout all measurements. An anodized Al plate with a 3.5 mm

aperture was taped to the sample plate in front of the integrating sphere. PDMS samples

were mounted to the anodized Al plate with the spring clips on the sample plate with the

impinging light incident on the textured PDMS surface. Transmission measurements

were taken relative to an air reference.

5.2.4 Scanning Electron Microscopy of Nanotextured PDMS

As mentioned previously, a small (~5mm x 5mm) PDMS sample was inserted in

the RIE with the larger PDMS sample. This method was assumed to produce two

samples of identical surface topology. SEM samples were sputtered with Pd/Au for 20

sec to reduce charging during SEM imaging. I used a JEOL 7000F SEM under 15 kV

accelerating voltage. All images shown here were taken under 55000x magnification at a

sample stage tilt of 50⁰.

5.3 Nano-Scale Columnar Texture on Plasma-Etched PDMS

Preliminary experiments examining the surface topology and optical transmission

of plasma-etched PDMS has shown promise in achieving anti-reflective properties.

Initial etching experiments, not shown here, were conducted merely to examine the

Page 159: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

150

surface topology following RIE treatments of varying power, time, and gas content.

Examination via SEM showed definitive nano-structures, columnar in shape, protruding

from the surface of the PDMS sample, especially for high power and/or long etching

times. These initial experiments prompted further and closer examination of etched

PDMS and more in-depth analysis of its topological and optical properties.

Follow-up experiments were focused on replicating the previously obtained

results on substrates that were appropriately sized for mounting in a UV-Vis

spectrophotometer for the purposes of optical transmission characterization. As

mentioned before, smaller samples were etched alongside large samples (those used in

UV-Vis measurements) for SEM examination. The SEM images show very clearly the

emergence and evolution of columnar/fibular-type structures as etching time increases

from one to seven minutes. Also evident is the increasing size and spacing of these

features. Table 5.2 shows these critical topology metrics for selected samples.

Avg. Column

Width (nm)

Avg. Column

Height (nm)

Avg. Edge-to-Edge

Spacing (nm)

Batch 2, 7min 200mT, 200W, 10sccm CF4, 7min

73 (±23) 109 (±35) 56 (±17)

Batch 3, 7min 200mT, 300W, 10sccm CF4, 7min

116 (±41) 319 (±56) 153 (±50)

Batch 5, 7min 200mT, 200W, 5/5sccm CF4/O2, 7min

51 (±16) 112 (±13) 61 (±15)

Batch 6, 7min 200mT, 300W, 5/5sccm CF4/O2, 7min

101 (±53) 190 (±20) 155 (±48)

Table 5.2

From this data and qualitative examination of the SEM images, broad

generalization can be made:

1) Width, height, and spacing all increase with increasing etch time and power.

Page 160: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

151

2) Nano-columns appear to be remaining, un-etched material from the bulk PDMS

substrate, not redeposition of etched by-products. This is most evident at the base

of the taller nano-columns, especially Figure 5.2f.

3) Longer etch times produce structures which appear to be aggregated fibular

networks whereas very short etch time’s topologies produce structures which are

more monodisperse.

Of more interest is the optical performance of these etched PDMS samples. As is

evident from the Transmission and integrated intensity plots in Figures 5.1 – 5.4, optical

performance is, at best, only ~0.4% more transparent and, at worst, almost 0.2% less

transparent, than native PDMS from 350 nm to 1000 nm. While the enhancements are

quite small, data shown here represents only the first experiments conducted in this

project. Further optimization of the RIE conditions to tune the surface topography might

yield higher enhancements in transmission. The data shown here simply serves as a

proof-of-concept demonstrating the viability of this approach.

5.4 Future Work on Nanotextured PDMS

At the time of deposition of this dissertation there remain many unanswered

questions about this system: What is the chemical composition of the nano-columns?

Are the nano-columns mechanically robust? If not, can they be made mechanically

robust? What RIE conditions are most important for optimizing optical transmission?

Does formation of these nano-columns follow similar mechanisms to previous, similar

Page 161: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

152

reports in literature? Can plasma nanotexturing of PDMS be translated to other

silicones?

5.4.1 Optimization of Plasma Etching Conditions for High Optical Transparency

Reactive ion etching contains no fewer than six process variables, all of which

having a profound effect on the morphology and etch rate of any material. Preliminary

experiments have been conducted to begin to elucidate process-dependent variations in

surface topology and optical transmission. But this work is still infantile, requiring

numerous repeated experiments to validate this preliminary data. Also in question is the

reproducibility of the nanotextured structure for a given etching condition. If the process

is not reproducible, is there an inherent variability in the process or etching mechanism or

is the non-reproducibility a function of RIE tools available at the University of Illinois?

5.4.2 Optical Simulations for Optimizing Nanotextured Surface Topology

Due to the large number of process variables inherent to RIE and the larger

number of resulting surface topologies of nanotextured PDMS, optical simulation will

play an important role in determining important topological parameters. Wilson and

Hutley [26] outlined generic guidelines for moth-eye structures. For periodic arrays of

nano-scale protuberances of height h and spacing d,

h/λ > 0.4

5.5

d < λ/n

5.6

Page 162: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

153

for normal incidence transmission, and

d < λ/2n

5.7

for oblique incidence transmission. According to the SEM images in Figures 5.1 – 5.4,

the nano-columns are not periodic and, therefore, will not follow these straightforward

guidelines. Previous research on moth-eye structures has relied heavily on rigorous

coupled-wave analysis (RCWA) [27] which deals with diffraction phenomena of

electromagnetic radiation by periodic structures. Such analyses will be paramount to

efficiently discovering optimal surface topologies such as nano-column height, width,

and spacing.

5.4.3 Compositional Analysis of Nanotextured PDMS

There is a high probability that the nano-pillars formed via RIE have surface

chemistries significantly different from that of native PDMS. For instance, Manca et al.

[21] has studied the surface chemistry of CF4-plasma etched PDMS and found via x-ray

photoelectron spectroscopy (XPS) that the surface “became enriched with fluorine”

following RIE. He goes on to explain that fluorine is present in multiple forms on the

sample’s surface: fluoro-carbon films, F-substituted methyl moieties, and Si—F bonding.

This indicates that significant surface modification is taking place as well as deposition of

fluoro-carbon films.

Page 163: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

154

5.4.4 Surface Characterization with Atomic Force Microscopy

Atomic force microscopy (AFM) of the nanotextured surfaces will allow the

ability to calculate important physical characteristics of the nanotextured surface such as

nano-column height and spacing. These physical parameters will be vital in relating

experimental transmission results to optical simulations. Surface maps of the etched

PDMS can be used in optical simulations, such as RCWA, to theoretically calculate

reflection losses from the surface.

Page 164: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

155

5.5 Chapter 5 Figures

Figure 5.1: a) Transmission, b) change in integrated intensity relative to native PDMS,

and tilted SEM images of nano-textured PDMS (Batch 2 from table 5.1) following RIE

etching under conditions of 200 mTorr, 200W, 10 sccm CF4 for c) 1min, d) 2 min, e) 4

min, and f) 7min.

Page 165: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

156

Figure 5.2: a) Transmission, b) change in integrated intensity relative to native PDMS,

and tilted SEM images of nano-textured PDMS (Batch 3 from table 5.1) following RIE

etching under conditions of 200 mTorr, 300W, 10 sccm CF4 for c) 1min, d) 2 min, e) 4

min, and f) 7min.

Page 166: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

157

Figure 5.3: a) Transmission, b) change in integrated intensity relative to native PDMS,

and tilted SEM images of nano-textured PDMS (Batch 5 from table 5.1) following RIE

etching under conditions of 200 mTorr, 200W, 5 sccm CF4, 5 sccm O2 for c) 1min, d) 2

min, e) 4 min, and f) 7min.

Page 167: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

158

Figure 5.4: a) Transmission, b) change in integrated intensity relative to native PDMS,

and tilted SEM images of nano-textured PDMS (Batch 5 from table 5.1) following RIE

etching under conditions of 200 mTorr, 300W, 5 sccm CF4, 5 sccm O2 for c) 1min, d) 2

min, e) 4 min, and f) 7min.

Page 168: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

159

5.6 Chapter 5 References

1. Redfield, D., Multiple-Pass Thin-Film Silicon Solar Cell. Appl. Phys. Lett., 1974.

25(11): p. 647-648.

2. Yablonovitch, E. and G.D. Cody, Intensity Enhancement in Textured Optical

Sheets for Solar Cells. IEEE T. Electron Dev., 1982. 29(2): p. 300-305.

3. Green, M.A., Lambertian Light Trapping in Textured Solar Cells and Light-

Emitting Diodes: Analytical Solutions. Prog. Photovoltaics Res. Appl., 2002.

10(4): p. 235-241.

4. Campbell, P. and M.A. Green, Light Trapping Properties of Pyramidally

Textured Surfaces. J. Appl. Phys., 1987. 62(1): p. 243-249.

5. Green, M.A., et al. 20% Efficient Laser Grooved, Buried Contact Silicon Solar

Cells. in Photovoltaic Specialists Conference, 1988., Conference Record of the

Twentieth IEEE. 1988.

6. Chen, Z., et al., A Novel and Effective PECVD SiO2/SiN Antireflection Coating

for Si Solar Cells. IEEE T. Electron Dev., 1993. 40(6): p. 1161-1165.

7. Johnson, C.C., T. Wydeven, and K. Donohoe, Plasma-Enhanced CVD Silicon

Nitride Antireflection Coatings for Solar Cells. Sol. Energy, 1983. 31(4): p. 355-

358.

8. Zhao, J. and M.A. Green, Optimized Antireflection Coatings for High-Efficiency

Silicon Solar Cells. IEEE T. Electron Dev., 1991. 38(8): p. 1925-1934.

9. Shockley, W. and H.J. Queisser, Detailed Balance Limit of Efficiency of p-n

Junction Solar Cells. J. Appl. Phys., 1961. 32(3): p. 510-519.

10. Norris, A.W., M. Bahadur, and M. Yoshitake, Novel Silicone Materials for LED

Packaging. Proc. SPIE, 2005: p. 594115-594115.

11. Pern, F.J. and S.H. Glick, Photothermal Stability of Encapsulated Si Solar Cells

and Encapsulation Materials upon Accelerated Exposures. Sol. Energ. Mat. Sol.

C., 2000. 61(2): p. 153-188.

12. Kojima, T. and T. Yanagisawa, Ultraviolet-Ray Irradiation and Degradation

Evaluation of the Sealing Agent EVA Film for Solar Cells Under High

Temperature and Humidity. Sol. Energ. Mat. Sol. C., 2005. 85(1): p. 63-72.

13. Bernhard, C.G., Structural and Functional Adaptation in a Visual System.

Endeavour, 1967. 26: p. 79-84.

14. Clapham, P.B. and M.C. Hutley, Reduction of Lens Reflexion by the Moth Eye

Principle. Nature, 1973. 244: p. 281-282.

15. Baker, K.M., Highly Corrected Close-Packed Microlens Arrays and Moth-Eye

Structuring on Curved Surfaces. Appl. Opt., 1999. 38(2): p. 352-356.

16. Sun, C.-H., P. Jiang, and B. Jiang, Broadband Moth-Eye Antireflection Coatings

on Silicon. Appl. Phys. Lett., 2008. 92(6): p. 061112-3.

17. Chen, Q., et al., Broadband Moth-Eye Antireflection Coatings Fabricated by

Low-Cost Nanoimprinting. Appl. Phys. Lett., 2009. 94(26): p. 263118-3.

Page 169: fabrication and transfer assembly of microscale, solid-state light emitting diodes and solar cells for

160

18. Kanamori, Y., E. Roy, and Y. Chen, Antireflection Sub-Wavelength Gratings

Fabricated by Spin-Coating Replication. Microelectron. Eng., 2005. 78–79(0): p.

287-293.

19. Tserepi, A.D., M.E. Vlachopoulou, and E. Gogolides, Nanotexturing of

Poly(dimethylsiloxane) in Plasmas for Creating Robust Super-Hydrophobic

Surfaces. Nanotechnology, 2006. 17(15): p. 3977.

20. Gogolides, E., et al., Controlling Roughness: From Etching to Nanotexturing and

Plasma-Directed Organization on Organic and Inorganic Materials. J. Phys. D:

Appl. Phys., 2011. 44(17): p. 174021.

21. Manca, M., et al., Influence of Chemistry and Topology Effects on

Superhydrophobic CF4-Plasma-Treated Poly(dimethylsiloxane) (PDMS).

Langmuir, 2008. 24(5): p. 1833-1843.

22. Ming-Hung, C., C. Yun-Ju, and T. Fan-Gang, Self-Masked High-Aspect-Ratio

Polymer Nanopillars. Nanotechnology, 2008. 19(50): p. 505301.

23. Cortese, B., et al., Superhydrophobicity Due to the Hierarchical Scale Roughness

of PDMS Surfaces. Langmuir, 2008. 24(6): p. 2712-2718.

24. Szmigiel, D., et al., Deep Etching of Biocompatible Silicone Rubber.

Microelectron. Eng., 2006. 83(4–9): p. 1178-1181.

25. Szmigiel, D., et al., Fluorine-Based Plasma Treatment of Biocompatible Silicone

Elastomer: The Effect of Temperature on Etch Rate and Surface Properties.

Plasma Process. Polym., 2008. 5(3): p. 246-255.

26. Wilson, S.J. and M.C. Hutley, The Optical Properties of 'Moth Eye' Antireflection

Surfaces. Opt. Acta, 1982. 29(7): p. 993-1009.

27. Moharam, M.G., et al., Formulation for Stable and Efficient Implementation of

the Rigorous Coupled-Wave Analysis of Binary Gratings. J. Opt. Soc. Am. A,

1995. 12(5): p. 1068-1076.