high efficiency ibc solar cell with oxide-nitride-oxide passivation · 2019. 12. 20. · teng choon...

189
High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the Australian National University

Upload: others

Post on 25-Aug-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

High efficiency IBC solar cell

with Oxide-Nitride-Oxide passivation

Teng Choon Kho

August 2019

A thesis submitted for the degree of Doctor of Philosophy

of the Australian National University

Page 2: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

I certify that this thesis does not incorporate without acknowledgement any material submitted

for a degree or diploma in any university and that, to the best of my knowledge, it does not

contain any material previously published or written by another person except where due

reference is made in the text. The work in this thesis is my own, except for the contributions

made by others as described in the Acknowledgements.

Declaration

Page 3: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

I would like to express my deepest gratitude to all my supervisors who they possess one the

brightest minds in solar, Prof. Andrew Blakers, Dr Kean Fong, Dr Matthew Stocks and Dr

Evan Franklin. They mentored and provided me with the opportunity to explore, learn and

fabricate high-efficiency silicon solar cells.

I would also like to express my undying appreciation and respect to rock star mentor Dr Keith

McIntosh, who initially inspired me in research way back in 2008. I’m humbled by his

knowledge and willingness to guide me in many different aspects of life.

I’m privileged to have been able to work with others within the department. I’m thankful to

have constant impromptu discussions Prof. Dan MacDonald and Prof. Andres Cuevas. The

laboratory wouldn’t be operational without the support from Christian Samundsett, Nina De

Caritat, Mark Saunders, James Cotsell, Bruce Condon, Maureen Brauers.

I would like to further thank Dr Fouad Karouta and Dr Kaushal Vora for allowing and assisting

me in using the equipment at the ANFF ACT Nodes.

The journey of completing this thesis would not be completed without thanking the ridiculous

jolly group of friends, Dr Pheng Phang, Dr Er-Chien Wang, Dr Kelvin Sio, Dr Di Yan, Dr

James Bullock, Dr Teck Kong, Dr Chog Barugkin, Dr Tom Ratcliff, Ingrid Haedrich, Dr Soe

Zin and Dr Nicholas Grant.

I would like to thank my immediate and extended family member and friends for supporting

me. To my papa and mama, both my sisters Sue Ann and Jey Vonn, three of my childhood

friends in Malaysia (Chung Lern Lee, Kim Zen Fang and Hui Ni Tan) and especially my

dependable wife Swee Wah Yap, who she had to endure and supported me throughout this.

This thesis has been externally reviewed by The Expert Editors following the Australian

standards for editing practice.

Acknowledgements

Page 4: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

Symbol Description

Al Aluminium

Al2O3 Aluminium oxide

ALD Atomic layer deposition

Alneal FGA with presence of aluminium

ANU The Australian National University

ARC Anti-reflection coating

BBr3 Boron tribromide

BSF Back surface field

BSG Boronsilicate glass

CC Corona charge

CSIRO Commonwealth Scientific and Research Organisation

C-V Capacitance-voltage

Cz Czochralski

DLTS Deep-level transient spectroscopy

Eq Equation

EQE External quantum efficiency

ESR Electron spin resonance

FCA Free carrier absorption

FGA Forming Gas annealing

FSF Front surface field

FZ Float Zone

g value Lande Splitting factor

HF Hydrofluoric acid

HMDS Hexamethyldisilazane

IBC Interdigitated back contact

IQE Internal quantum efficiency

I-V Current-voltage

LPCVD Low pressure chemical vapour deposition

m Local ideality factor

N/Si Nitrogen/silicon ratio

ONO SiO2-SiNx-SiOx dielectric stack

PECVD Plasma enhanced chemical vapour deposition

PL Photoluminescence

POA Post-oxidation annealing

List of abbreviations

Page 5: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

POCl3 Phosphoryl chloride

PSG Phosphosilicate glass

PV Photovoltaics

Rantex Random texture

RCA Radio Chemical America

Si3N4 Silicon nitride

SiNx Amorphous silicon nitride

SiO2 Thermal silicon oxide

SiOx Amorphous silicon oxide

TMA Trimethylaluminum

TMAH Tetramethylammonium hydroxide

TOPcon Tunnel oxide passivated contact

Page 6: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

Symbol Description Unit

capture cross section of electrons cm2

capture cross section of holes cm2

µn Mobility of electron cm2/Vs

µp Mobility of holes cm2/Vs

Dit Interface states density eV-1.cm-2

EC Energy level of conduction band eV

EV Energy level of valance band eV

FF Fill factor %

fO Recombination ratio between 111 to 100 orientation

fV Recombination ratio between area corrected rantex to 111 orientation

G Photogeneration rate m-3

J0 Saturation current fA/cm2

J0e Emitter saturation current fA/cm2

J0s Surface saturation current fA/cm2

Jsc Short circuit current mA/cm2

k Extinction coefficient

kT Boltzmann constant (k) with temperature in kelvin eV

LD Debye length m

n Refractive index

n Electron concentration cm-3

n0 Thermal equilibrium of electron density cm-3

NA Acceptor dopants in substrate cm-3

NC Effective densities of states in the conduction band cm-3

ND Donor dopants in substrate cm-3

ni Intrinsic carrier concentration cm-3

NV Effective densities of states in the valance band cm-3

nλ Refractive index at λ

nλ Refractive index at define wavelength

p Holes concentration cm-3

p0 Thermal equilibrium of hole density cm-3

Qeff Effective charge cm-2

Qox Charges in oxide cm-2

Rseries Series resistance Ω/cm2

List of symbols

Page 7: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

Rshunt Shunt resistance Ω/cm2

Seff Effective surface recombination velocity cm/s

Seff.UL Upper limit of effective surface recombination velocity, assuming infinite bulk

lifetime

cm/s

VFB Flatband voltage V

Voc Open circuit voltage mV

w Thickness of substrate µm

α Absorption coefficient cm-1

∆n Excess carrier density cm-3

σ Conductivity S/m

τbulk Effective lifetime of bulk s

τeff Effective lifetime s

τeff.max Maximum effective lifetime across measured injection level s

τeff_1015 Effective lifetime at injection level of 1015 cm-3 s

τSRH Shockley-Read-Hall lifetime S

τsurface Effective lifetime of surface S

Page 8: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

Abstract

This thesis considers the optimisation of an SiO2-SiNx-SiOx (ONO) dielectric coating for

silicon solar cells comprising thermal SiO2, plasma-enhanced chemical vapour deposition

PECVD SiNx and SiOx, together with the development and fabrication of high efficiency

interdigitated back contact (IBC) silicon solar cells.

The three-layer ONO dielectric stack provides excellent surface passivation and anti-reflection

properties. The initial thermal SiO2 layer provides chemical surface passivation. The second

layer of the PECVD SiNx hydrogenates the underlying thermal SiO2 and provides a field-effect

passivation with positive charges. The third layer of the PECVD SiOx improves the overall

anti-reflection coating (ARC) of an ONO stack in air. Positive corona charging of the ONO

stack further improves the surface passivation, and annealing at 400 ˚C traps the charges and

renders it stable for at least two years. An optimised ONO stack for surface passivation for

crystalline silicon has achieved effective lifetimes above previous parameterised Auger limits.

The n-type wafers with resistivities of 0.5, 1.07, 1.77 and 100 Ω.cm achieved lifetimes of 3.7,

15.1, 25.5 and 170 ms, respectively.

The surface passivation of an ONO stack was fine-tuned for both phosphorus and boron

diffusions for integration into IBC silicon solar cells. A large contribution of positive charges

within the stack was mainly from the thermal SiO2 and PECVD SiNx. Post-oxidation annealing

(POA) of the thermal SiO2 and high SiNx refractive index reduced the overall positive charge

of the ONO stack. A lower positive charge improved the passivation on the boron diffused

surface while retaining excellent passivation on the phosphorus diffused surface.

The cell design and fabrication processes for IBC silicon solar cells were improved from

previous fabrication processes at the Australian National University (ANU) by utilising ONO

surface passivation. The refinement for IBC silicon solar cell design was based on detailed

simulations. The improvements to the cell processing included permanently eliminating low

bulk lifetime defects on high resistivity n-type float zone (FZ) wafers, gettering of

contaminants, and consistent texturing results. Implementing optimised ONO stacks for surface

passivation and an ARC onto the IBC silicon solar cell achieved a certified efficiency of 25.0

± 0.6%. Non-ideal recombination contributed to a relatively low fill factor for the champion

Page 9: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

cell, and disregarding the non-ideal recombination in the simulation results revealed that an

efficiency of 25.2% is achievable.

Page 10: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

Contents

1. Introduction .................................................................................................................. 1

1.1 Motivation ........................................................................................................... 1

1.2 Outline and highlights ......................................................................................... 2

2. Review of surface passivation ..................................................................................... 5

2.1 Introduction ....................................................................................................... 5

2.2 Thermal silicon oxide ......................................................................................... 6

2.2.1 Thermal oxide charges ........................................................................... 8

2.2.2 Growth and optical properties of thermal oxide .................................... 9

2.2.3 Surface passivation of thermal oxide ..................................................... 9

2.3 Plasma-enhanced chemical vapour deposition of silicon nitride ..................... 10

2.3.1 PECVD silicon nitride charges ............................................................ 12

2.3.2 Optics of PECVD silicon nitride.......................................................... 13

2.3.3 Surface passivation of PECVD silicon nitride ..................................... 14

2.4 Atomic layer deposition of aluminium oxide .................................................. 15

2.4.1 ALD aluminium oxide charges ............................................................ 16

2.4.2 Optics of ALD aluminium oxide ......................................................... 16

2.4.3 Surface passivation of ALD aluminium oxide..................................... 17

2.5 Doped polysilicon passivated contacts ............................................................ 17

2.5.1 Optics of polysilicon ............................................................................ 18

2.5.2 Passivation and contact resistivity of doped polysilicon ..................... 18

2.6 Electret ............................................................................................................. 19

2.6.1 Corona charging of silicon oxide and silicon nitride electret .............. 20

2.6.2 Surface passivation of corona charged silicon oxide ........................... 21

2.7 Chapter summary ............................................................................................. 22

3. Characterisation methods ......................................................................................... 24

3.1 Photoconductance measurement and recombination mechanism in solar cell . 24

3.1.1 Bulk recombination .............................................................................. 25

Page 11: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

3.1.2 Surface recombination ......................................................................... 27

3.1.3 Emitter recombination model .............................................................. 28

3.2 Capacitance-voltage measurement.................................................................... 29

3.3 Current-voltage measurement ........................................................................... 31

4. ONO: Growth and deposition conditions ................................................................ 35

4.1 Introduction ..................................................................................................... 35

4.2 ONO baseline procedure ................................................................................... 35

4.3 Thermal oxide growth conditions for an ONO stack ........................................ 36

4.3.1 The effect of varying the thermal oxide thickness ............................... 37

4.3.2 The effect of post-oxidation annealing of thermal oxide ..................... 39

4.4 Deposition conditions of PECVD for ONO stack ............................................ 41

4.4.1 PECVD dielectric stack on thermal oxide ........................................... 45

4.4.2 Deposition conditions of PECVD SiNx in ONO stack ........................ 47

4.4.3 Deposition conditions of PECVD SiOx in ONO stack ....................... 56

4.5 Impact of corona charging and annealing on ONO stack ................................. 59

4.6 ONO surface passivation on various wafer resistivities and surfaces .............. 62

4.7 Chapter summary .............................................................................................. 64

5. Passivation of ONO on diffused surfaces ................................................................. 66

5.1 Introduction ..................................................................................................... 66

5.2 ONO baseline procedure on diffused surfaces .................................................. 66

5.3 Surface passivation of ONO stack on phosphorus and boron diffusion ........... 67

5.4 Thermal oxide growth conditions in an ONO stack on diffused surfaces ........ 69

5.4.1 The effect of varying thermal oxide thickness on diffused surfaces ... 70

5.4.2 The effect of post-oxidation annealing of thermal oxide on diffused

surface .................................................................................................. 73

5.5 Deposition conditions for PECVD SiNx of the ONO stack on diffused surfaces

........................................................................................................................... 76

5.6 Impact of corona charging and annealing on diffused surfaces with ONO stack

........................................................................................................................... 79

5.7 Chapter summary .............................................................................................. 81

Page 12: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

6. Simulation of ONO IBC solar cells .......................................................................... 83

6.1 Introduction ....................................................................................................... 83

6.2 Optical losses .................................................................................................... 83

6.2.1 OPAL2 simulation inputs .................................................................... 84

6.2.2 Simulation of ONO with OPAL2 ........................................................ 87

6.2.3 Free carrier absorption ......................................................................... 90

6.2.4 Simulation of FCA with SunSolve ...................................................... 91

6.3 Recombination losses of ONO IBC solar cell .................................................. 92

6.3.1 Simulation for front and rear surface recombinations with phosphorus

surface field .......................................................................................... 94

6.3.2 Simulation for rear boron-doped emitter ............................................. 99

6.3.3 Simulation for rear phosphorus-doped contact region ....................... 101

6.3.4 Simulation of different pitch sizes for the ONO IBC solar cell ......... 102

6.3.5 Simulation of bulk lifetime, resistivity and edge recombination ....... 104

6.3.6 Simulation of series and shunt resistance .......................................... 106

6.4 Chapter summary ............................................................................................ 109

7. Fabrication and analysis of ONO IBC solar cells ................................................. 110

7.1 Introduction ..................................................................................................... 110

7.2 Fabrication sequence ....................................................................................... 110

7.2.1 Chemical cleaning .............................................................................. 111

7.2.2 Pre-oxidation of wafers ...................................................................... 113

7.2.3 Gettering ............................................................................................ 114

7.2.4 Masking layer..................................................................................... 115

7.2.5 Photolithography pattering................................................................. 116

7.2.6 Diffusion process ............................................................................... 117

7.2.7 Texturing ............................................................................................ 118

7.2.8 Metal contact and fingers patterning .................................................. 120

7.3 Improvements to the ONO IBC solar cells fabrication process ...................... 121

7.3.1 Boron-rich layer gettering .................................................................. 121

7.3.2 Surface passivation on IBC solar cell ................................................ 124

7.3.3 Positive corona charging on the front surface of ONO IBC solar cells

............................................................................................................ 132

Page 13: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

7.3.4 Lift-off metallisation on ONO IBC solar cell .................................... 138

7.4 ONO IBC cell with certified conversion efficiency of 25% ........................... 144

7.5 Chapter summary ............................................................................................ 147

8. Conclusion ................................................................................................................ 148

Appendix ......................................................................................................................... 152

List of publication .......................................................................................................... 158

Bibliography ......................................................................................................................... 159

Page 14: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

1

Chapter 1

1.1 Motivation

Photovoltaics have been growing exponentially for more than the last quarter century [1]. There

has been significant growth over the last decade worldwide, with China leading the way. This

growth was accelerated by government policies and incentives to increase reliability and reduce

the costs of renewable energy.

Advancements in manufacturing processes have drastically reduced the costs of photovoltaics

(PV) [2]. As of 2017, silicon wafer-based technologies accounted for 95% of the total PV

industry [3]. The improved efficiency of silicon solar cells leverages cost reductions across the

entire PV value chain.

1980 1990 2000 20100

10

20

30

40

50

60

70

80

US

D p

er

wa

tt (

$/W

)

Year

Fig 1.1: (a) Global growth of PV capacity with regional shares [4-8]. (b) Price in USD per watt of crystalline-silicon modules [9, 10].

Interdigitated back contact (IBC) solar cells are capable of achieving high efficiencies.

Traditional front contact cells reflect a portion of light from the front contact metallised fingers

and busbar, whereas IBC solar cells avoid shading losses by having all the contacts at the rear.

The highest efficiency IBC silicon solar cell reported to date of a single crystal and

0

100

200

300

400

500

600

0

100

200

300

400

500

600

2012 2013 2014 2015 2016 2017 2018

Glo

bal g

row

th o

f cum

ula

tive

PV

(G

W)

Year

Others

China

APAC

America

Europe

2018 Global Estimate

Introduction

$76.67

(1977)

$0.30 (2015)

(a) (b)

Page 15: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

2

heterojunction structure under one sun are 26.1% (cell area size of 4 cm2) [11, 12] and 26.6%

(cell area of 79 cm2) respectively [12, 13].

The highest independently certified efficiency for an IBC solar cell fabricated at the Australian

National University (ANU) prior to this work was 24.4% (cell area size of 4 cm2), with an open

circuit voltage (Voc) of 703 ± 3.5 mV, a short circuit current (Jsc) of 42 ± 1 mA/cm2 and a fill

factor (FF) of 82.7 ± 0.8% [14]. The cell was fabricated from a moderate-resistivity (1 Ω.cm)

n-type wafer, with a sheet boron emitter, localized phosphorus contact dots, front surface

passivation with PECVD SiNx-SiOx stack and rear surface passivation from a thermal SiO2 and

LPCVD Si3N4 stack.

Developing a high-efficiency IBC solar cell requires a high bulk lifetime, excellent surface

passivation, low contact recombination, low contact resistivity, optimised emitter collection

geometry, an excellent anti-reflection coating (ARC) and, ideally, minimal edge recombination

losses [13-15]. This thesis presents the development of excellent surface passivation and ARC

using a thermal SiO2 and PECVD SiNx-SiOx (ONO) stack integrated into IBC solar cells. The

surface passivation of the ONO was studied on both undiffused and diffused surfaces. The data

obtained for the ONO surface passivation was modelled and simulated with the simulation

package Quokka3 [16]. The cell design was based on the simulation results and the fabrication

sequence of the IBC solar cell was established from a previous ANU IBC process with

additional refinements to maintain a high bulk lifetime and achieve repeatable uniform random

texturing (rantex) processes [14]. The champion cell in this work achieved an independently

certified efficiency of 25.0 ± 0.6%, with a Voc of 717 ± 1 mV, a Jsc of 42.9 ± 0.8 mA/cm2 and

a FF of 81.1 ± 1.9%, setting a new ANU record for a single crystal silicon solar cell.

1.2 Outline and highlights

Chapter 2 reviews the various surface passivation methods commonly used in silicon solar cells.

Each of the reviewed surface passivation techniques is discussed in detail with respect to the

chemical bonding, dielectric charges and optical properties. The thermal SiO2, PECVD SiNx,

and atomic layer deposition (ALD) Al2O3 surface passivation dielectrics were evaluated and

compared with optimised surface passivations achieved at ANU.

Page 16: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

3

Chapter 3 introduces the measurement techniques used to calculate and measure the surface

recombination and effective lifetime, charge density in dielectrics, cell efficiency and shunt

and series resistances.

Chapter 4 presents an extensive study of the ONO structure, which consists of a thermal SiO2,

PECVD SiNx and PECVD SiOx stack. The dependence of the dielectric stack on the growth

conditions were systematically investigated on an undiffused planar and rantex surface of n-

type high resistivity (100 Ω.cm) wafers. The effects of the growth conditions of the thermal

SiO2 on the ONO surface passivation were explored with different thermal SiO2 thicknesses

and post-oxidation annealing (POA) with N2. The deposition conditions for the SiNx were

investigated, including the deposition temperature, SiH4 and NH3 gas flow ratio and chamber

pressure. The deposition conditions for the PECVD SiOx layer were investigated by varying

the deposition temperature and gas ratio. The surface passivation of the ONO stack was further

enhanced using corona charging and annealing to improve the charge stability. Based on these

studies, the fine-tuned ONO dielectric stack achieved lifetimes above the parameterised Auger

limits on n-type 0.50, 1.07 and 1.50 Ω.cm wafers [17].

Chapter 5 reviews the ONO structure on diffused surfaces. The dependence of the thermal SiO2

and PECVD SiNx deposition conditions were investigated on boron and phosphorus diffused

planar wafers as well as on phosphorus-diffused rantex wafers. The growth conditions were

varied for the thermal SiO2 layer, including the thermal SiO2 thickness and the duration of the

POA in N2. The deposition conditions for the PECVD SiNx were varied with different SiH4

and NH3 gas flow ratios. The ONO provides excellent chemical surface passivation and

achieved a J0e of 0.5 fA/cm2 and 1.9 fA/cm2 on a light phosphorus diffused planar (520 Ω/)

and rantex surface (610 Ω/) respectively, and 4.8 fA/cm2 on the light boron diffused planar

surface (154 Ω/) after corona charging.

Chapter 6 investigates the different losses for an IBC solar cell using the simulation package

Quokka3, the optical simulator OPAL2, the emitter modelling EDNA2 and the ray tracer

simulator SunSolve [16, 18, 19]. The optical simulation included varying the thicknesses of the

thermal SiO2, PECVD SiNx and SiOx. The device simulation conditions included the surface,

bulk and emitter recombination, various optical losses, contact resistivity, series and shunt

resistances and edge recombination. The obtained simulated results were applied to the cell

fabrication process.

Page 17: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

4

Chapter 7 employs the simulation results from Chapter 6 in the IBC solar cell design while

considering fabrication limitations. The fabrication process was improved from the previous

ANU IBC process by introducing boron rich layer (BRL) gettering and improving the texturing

consistency [14]. Three different aspects of IBC solar cells were investigated: surface

passivation, corona charging and annealing and rear metallisation of the cell.

Chapter 8 summarises the key finding from each chapter.

Page 18: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

5

Chapter 2

Review of surface passivation

2.1 Introduction

Surface recombination remains one of the limiting factors in developing high-efficiency solar

cells. Termination of bonds at the surface of the atomic lattice of a semiconductor causes highly

active recombination sites for electron and hole carriers. The termination of bonds which are

left unbound to any other bonds are known as ‘dangling bonds,’ and the recombination sites at

the surface are commonly referred to as either ‘surface’ or ‘interface’ states. These dangling

bonds allow the recombination of generated carriers within the semiconductor, which reduces

the solar cell efficiency.

In PV, surface passivation is defined as a reduction of recombination events at the surface.

There are two common ways to achieve surface passivation: i) reducing the dangling bonds on

the surface (chemical passivation), and ii) reducing the carrier concentration of one type

(electrons or holes) of free charge carriers at the surface (field-effect passivation).

The ideal chemical surface passivation would be to have chemical bonding at all the surface

dangling bonds of the semiconductor to deactivate the surface recombination sites. Chemical

passivation can be achieved by establishing covalent bonding onto the surface. Dielectric

materials with a bandgap larger than the semiconductor are generally used for surface

passivation, such as thermal SiO2 on silicon. However, due to lattice mismatch of dielectrics

and semiconductors, not all of the dangling bonds are passivated. These additional

unpassivated dangling bonds can be further reduced by subjecting them to additional hydrogen

at the interface using a forming gas annealing (FGA) or hydrogen plasma process.

Field-effect passivation can be achieved by either introducing an external field to repel specific

carriers or doping to limit the recombination of specific carriers at the surface. The

conventional method of creating field-effect passivation is to dope the near-surface regions

with dopants such as phosphorus, boron or Al. The diffused region creates an electric field and

chemical potential within the semiconductor to suppress specific carrier recombination at the

surface [20]. Alternatively, an electric field can be attained by depositing a dielectric layer with

inbuilt charges or charges deposited onto the dielectric surfaces through corona charging. With

Page 19: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

6

sufficient charges, the surface recombination can be reduced by either accumulation or

inversion of the carriers at the surface of an undiffused wafer.

Incorporating both these passivation mechanisms (chemical and field-effect passivation)

simultaneously can significantly reduce surface recombination. Dielectrics such as Al2O3 and

SiNx have been shown to provide both excellent chemical and field-effect passivation with

effective surface recombination velocity (Seff) of <1 cm/s [21].

In the following sections, the growth conditions, characterised defects, optical properties and

chemical and field-effect passivation of the dielectric layers are reviewed.

2.2 Thermal silicon oxide

Thermal oxidation has long been regarded as a reliable passivation method in both the

electronics and PV industries. Atalla et al. in the late 1950s were the first to report surface

passivation with thermal SiO2 on p-n junction diodes [22]. Progressive improvements of silicon

semiconductor materials similarly improved the understanding of silicon PV. By late 1970, the

first silicon solar cells passivated with thermal SiO2 achieved efficiencies of nearly 17%, and

by mid-1980s, Blakers et al. managed to develop a solar cell with an efficiency of 20% by

passivating the cell via thermal oxidation [23, 24].

Thermal SiO2 is normally grown in high-temperature quartz furnaces, from 600 to 1100 C [25,

26], in the presence of O2 or steam. Oxygen atoms diffuse through the silicon lattice and

chemically react at the silicon surface to form thermal SiO2. With steam, the oxide growth is

five to ten times faster due to the formation of hydroxide ions (OH-) from the water, which

diffuses faster through the oxide layer as compared with O2 [27, 28]. POA is normally

performed after oxidation with an inert gas (N2 or Ar) at temperatures between 950 and 1200 °C

to reduce charges within the oxide [29, 30].

Thermally grown SiO2 structures are generally assumed to be amorphous [31, 32]. However,

Takahashi et al. and Munkholm et al. have shown that the SiO2 could maintain an epitaxial

relation with the silicon substrate on 001 wafers [33, 34]. During thermal oxidation,

differences in the bond lengths of the crystalline silicon and thermal SiO2 induce stresses at the

Si-SiO2 interface. The local relaxation of the stress at the interface may cause dangling bond

defects that are defined as Pb centres [35-40], which are commonly characterised via electron

spin resonance (ESR). Early work by Poindexter et al. described single defect as Pb on a 111

Page 20: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

7

orientated wafer with dangling bonds as an sp3[111] hybrid (ᵒSi≡Si3) [41]. Poindexter et al. later

reported a 100 oriented wafer that exhibited two defect types, defined as Pb0 and Pb1 [41].

Although the Pb and Pb0 defects were reported to be nearly identical, there is still ambiguity

around the Pb1 defect [36]. Stesmans et al. suggested that the Pb1 defect is electrically inactive

[42]; however, Campbell et al. reported that the defect was active within the silicon bandgap

[43]. It has been observed from various studies that the Pb centres of 100 orientation are

smaller compared with the 111 orientated wafer [41, 42, 44]. Studies regarding these Pb

centres with ESR can be further correlated to interface state densities.

Interface traps due to dangling bonds that lie between the semiconductor and the passivating

dielectric can be quantified in terms of the interface states densities (Dit). The Dit for Si-SiO2

was first reported by Terman in the early 1960s using a metal oxide silicon (MOS) structure

with capacitance-voltage (C-V) measurements [45]. The measured Pb centres and Dit for

different levels of wafer doping, orientation and resistivity from various publications are

plotted in Figure 2.1 [46-48]. It is observed that the overall measured Pb centre densities are

correlated relatively linearly with Dit. Similarly, 100 wafers were reported to have a lower

Dit compared with 111 orientated wafers at the Si-SiO2 interface [49]. The introduction of

FGA, which is normally a mixture of 5% H2 and 95% N2 or Ar at temperatures from 400–

450 °C, has been shown to reduce active Pb centres and similarly lower the midgap Dit [30, 50-

52].

0 5 10 15 20 25 300

5

10

15

20

25

30

Mikawa et al. P 111 N

2

Yong et al. N 100 N

2

Poindexter et al. N 111 N

2

P 100 N2

P 111 N2

P 100 O2

P 111 O2

N 100 O2

N 111 O2

N 111 FGA

Pb s

pin

con

cen

tratio

n (

10

11/c

m2)

Midgap Dit (1011

/eV.cm2)

Fig 2.1: Correlation between the Pb defects and the interface state densities with different wafer orientations, doping and annealing processes [46-48].

Page 21: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

8

2.2.1 Thermal oxide charges

It has been known since 1970 that thermal SiO2 introduces fixed positive charges at the Si-

SiO2 interface [53]. After nearly half a century, the origin and nature of these charges remain

debated [54-56]. Thermal SiO2 has been observed to contain mostly positive charges, which

can be classified as mobile ion charges, fixed oxide charges, trapped oxide charges or interface

charges. The formation of these charges is dependent on the oxidation temperature, wafer

orientation, POA and ion contamination species [30, 53, 55, 56]. The work performed by Cheng

et al. (plotted in Figure 2.2a) shows that the charges within the oxide reduce as the oxidation

temperature increases [53]. Moreover, Deal et al. and Razouk et al. reported that thermally

grown SiO2 on 111 orientated wafers has more positive charges within the thermal SiO2 as

compared with 100 orientated wafers [30, 56]. Further studies performed by Razouk et al.

concluded that thermal SiO2 not subjected to any POA with N2 and Ar has slightly more

positive charges [30]. Furthermore, Lamb et al. revealed that annealing in inert gases at

temperatures up to 1200 °C reduces the positive oxide charges, as plotted in Figure 2.2b [53,

57].

400 600 800 1000 1200 1400

2

4

6

8

10

12

600 800 1000 1200 1400

5

10

15

20

SiO

2 Q

eff (1

011 c

m-2)

Oxidation Temperature (°C)

Dry P 1.1 Ω.cm

Dry N 0.4 Ω.cm

Wet P 1.1 Ω.cm

Wet N 0.4 Ω.cm

(a) (b)

Annealing Gas

He

Ar

N2

SiO

2 Q

eff (

10

11 c

m-2)

Annealing Temperature (°C) Fig 2.2: (a) SiO2 charges with different oxidation temperatures for dry and wet oxidation [53]. (b) Correlation of thermal SiO2 charges with post-annealing temperatures of the inert gases [53, 57].

2.2.2 Growth and optical properties of thermal oxide

The growth of thermal SiO2 on a wafer depends on various factors including the furnace

temperature, pressure, gasses, humidity, the orientation of the wafer and the doping

concentration.

Page 22: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

9

The Deal-Grove model, also known as the linear-parabolic oxide growth model, is commonly

used to estimate the thermal SiO2 growth under different furnace conditions [58]. The initial

oxide has a linear growth rate, and after reaching a certain thickness, the growth transitions to

a diffusion-limited parabolic growth. The Deal-Grove model was introduced in 1965 and is

well established for thermal SiO2 thicknesses above 30 nm. However, it has been

experimentally observed that the oxidation rate is much faster at the initial stages of oxidation

and for thin oxides under 30 nm. Various groups have proposed solutions for the limitations of

the Deal-Grove model by suggesting a second parallel oxidation reaction model [59], by fitting

thermal SiO2 experimental growth data to a power function [60] or by including a second term

in the Deal-Grove model [61].

The refractive index describes a change in the speed of light from one medium to another with

attenuation given by the complex value. The real component (i.e., refractive index), n, describes

the light velocity and the imaginary component (i.e., the extinction coefficient), k, describes

how the light is absorbed and scattered. The refractive index of thermal SiO2 has been well

characterized. At 632 nm, thermal SiO2 typically has a refractive index (n632) of ~1.45, which

is lower in comparison to α-quartz at ~1.53, while the k for thermal SiO2 has been measured to

be ~0 from 400–2000 nm [62, 63].

2.2.3 Surface passivation of thermal oxide

The first known silicon solar cell reported with thermal SiO2 surface passivation from the late

1970s achieved a Voc of 620 mV [24]. The improved development and understanding of the

thermal SiO2 surface passivation through the years has led to an increased Voc to above 700

mV [64]. These improvements have been achieved via hydrogenation and corona charging to

reduce the Dit and enhance the field-effect passivation, respectively [65, 66].

A variety of methods have been used to hydrogenate the interface of thermal SiO2-Si for solar

cells, such as FGA, alneal (annealing in the presence of Al) and plasma hydrogen. As

mentioned in Chapter 2.1, FGA is typically performed at 400–450 ˚C to reduce the defects of

Si-SiO2 through molecular hydrogenation. Thanh et al. showed that thermal SiO2 with a

thickness of 73 nm could reach a Dit as low as 1.2 × 1010 eV-1cm-2 [67]. Atomic hydrogenation

through alneal process or plasma hydrogenation typically reduces the Dit to be approximately

1010 eV-1cm-2 for thin thermal SiO2 layers [21, 68, 69].

Page 23: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

10

10-1 100 101 102

10-1

100

101

102

Seff U

L (

cm

/s)

Si wafer resistivity (Ω.cm)

Thermal SiO2

P-type 100

N-type 100

N-type 100 - this work

N-type rantex - this work

Fig 2.3: Compilation of silicon oxide passivation wafers with various resistivities from the literature (transparent symbols) [64, 69-72] compared with the thermal SiO2 performed for this work (filled symbols).

The Seff concept introduced in the early 1990s was to describe surface recombination at specific

injection levels [73]. In this section, a comparison between the upper limit surface

recombination velocity (Seff.UL) for hydrogenated thermal SiO2 obtained from literature

together with FGA thermal SiO2 performed at the ANU on undiffused wafers, are plotted in

Figure 2.3 [64, 69-72]. The Seff.UL assumes negligible bulk recombination with effective

lifetime (τeff) reported at carrier injections between 1013 and 1016 cm-3.

2.3 Plasma-enhanced chemical vapour deposition of silicon nitride

Plasma-enhanced chemical vapour deposition (PECVD) was invented by Stirling and Swann

in 1965 [74]. More recently, PECVD technology in the silicon PV industry has achieved

remarkable surface passivation with advancements in plasma reactor configurations.

There are three common types of PECVD reactor configurations used in the PV industry: direct,

remote and dual mode. In the direct PECVD reactor, all the reactant gases are excited while

the substrate is exposed to the plasma. In the remote PECVD reactor, only one of the reactant

gasses is excited, and the plasma is spatially separated from the substrate. In the dual mode

PECVD, the gas is excited within the chamber, but the substrate is outside of the plasma. For

the direct and remote reactors, the plasma frequency is commonly set to 13.56 MHz and 2.45

Page 24: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

11

GHz, respectively. However, in a dual mode PECVD, these two frequencies operate

simultaneously within the chamber.

100 200 300 400 500 600

101

102

103

1.6 2.0 2.4 2.8 3.2

101

102

103

SiN

x S

eff

UL (

cm

/s)

Deposition Temperature (°C)

Aberle et al.

Aberle et al.

Aberle et al.

Duttagupta Wan Kerr

(a) (b)

SiN

x S

eff U

L(c

m/s

)

Refractive Index (n)

Wan

Lauinger et al.

Mackel et al.

Duttagupta

Fig 2.4: Influence of the surface passivation for the PECVD SiNx with respect to (a) the deposition temperature [75-78] and (b) the different refractive indexes n632 by varying SiH4:NH3 gas ratio [76, 77, 79, 80].

PECVD SiNx is most commonly deposited using SiH4 and NH3 source gases together with

either a He or N2 as the diluent gas. The parameters that affect the deposition of PECVD SiNx

include the deposition temperature, plasma power, active gas ratio, chamber pressure and total

gas flow. A compilation of the different deposition temperatures is plotted in Figure 2.4a and

shows that optimal surface passivation can be obtained with films deposited at temperatures

around 300–450 ˚C. The surface recombination generally reduces with increasing SiNx

refractive index, as shown in Figure 2.4b. However, Wan stated that such observed trends are

due to optimised PECVD deposition conditions [77].

The defects in SiNx are not as well established as thermal SiO2. Silicon dangling bonds in SiNx

have many different configurations and charge states that can bond with silicon or nitrogen

atoms. Jousse et al. associated three distinct defect configurations at three different Lande

Splitting Factor (g values) and energy levels from the ESR and C-V measurements,

respectively [81]. Schmidt et al. defined four different defects (A, B, C and D) using deep-level

transient spectroscopy (DLTS) without knowledge of the atomic configurations for the defects

except for Defect D, which is thought to be the K centre [82, 83]. Stesmans et al. reported

similar ESR signature on dangling bond defect of a thermally grown silicon nitride (Si3N4)

(Pbn), and dangling bond defect of thermal SiO2 (Pb), on a 111 orientated silicon wafer which

Page 25: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

12

suggests both dielectrics exhibit similar defect types [84]. This was similarly observed by

Garcia et al. using DLTS on PECVD SiNx wafers, showing that the Dit increases with the Pbn

[84, 85]. The relationship between the Dit at the midgap and the SiNx charge are normally

compared with the defect configuration and the N/Si ratio or with the refractive index of SiNx.

The PECVD SiNx is known to have large quantities of hydrogen within the film. Furthermore,

hydrogen molecules and radicals are generated during the deposition of SiNx [86, 87].

The deposition of PECVD SiNx has been shown to hydrogenate grain boundaries and the bulk

regions of multi-crystalline solar cells [88, 89], passivate the dangling bonds of the Si-SiO2

interface [90-94] and even hydrogenate FZ bulk wafers [95]. While the mechanism in which

SiNx improves the surface passivation is not unequivocal, the SiNx film is capable of releasing

hydrogen through an annealing process. Sheoran et al. demonstrated the release of deuterium

into the silicon bulk via the rapid thermal annealing of SiNx:D at 750 ˚C for one second [96].

2.3.1 PECVD silicon nitride charges

The specific silicon-nitrogen dangling bonds in which three N atoms are back-bonded to a Si

atom (Si≡N3), known as K centre defects, are the primary charge-trapping defects present in

the SiNx film [97-100]. The K centres are reported to be responsible for the large charges

measured within SiNx. The K centre exists in three different charge states: neutral K0, positively

charged K+ and negatively charged K-, with one, zero or two electrons on the dangling bond,

respectively. Sharma et al. managed to alter the K trapped centres using both positive and

negative corona charging on the SiNx wafers [101]. The altered charges injected into the SiNx

were found to be stable after a year in ambient conditions.

Mackel and Ludemann proposed that N-H bonds act as precursor sites with the •Si≡N3 dangling

bonds as the K-centre [80]. Therefore, it is predicted that Qeff increases as the SiNx becomes

more N-rich. However, Wan et al. compiled information on various deposition conditions from

different PECVD reactors and found that the Qeff in SiNx is strongly dependent on the

deposition conditions [77]. Moreover, Wolf et al. showed that MOS structures with a dielectric

SiNx refractive index ranging from 2 to 3 were capable of trapping positive and negative

charges by altering the direction of voltage sweep when performing C-V measurements [102].

A compilation of the Qeff of PECVD SiNx with various refractive indexes is shown in Figure

2.5 with no observed trend over the considered range.

Page 26: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

13

1.6 2.0 2.4 2.8 3.2

1010

1011

1012

1013

Wan et al. (µW/RF dual-mode)

Lamers et al. (µW Remote)

De Wolf et al. (LF Direct)

Lelievre et al. (LF Direct)

Bagnoli et al. (RF Direct)

Dauwe (µW Remote)

Garcia et al. (ECR)

Hazel et al. (LF Direct)

Landheer et al. (ERC)

Helland (RF Direct)

SiN

x Q

eff (

cm

-2)

SiNx refractive index (n)

Fig 2.5: Effective charge Qeff measured within SiNx for various refractive indexes n632 [85, 102-109].

2.3.2 Optics of PECVD silicon nitride

The PECVD SiNx has a large range of refractive indexes that generally range between 1.9 and

2.7. An ARC of single layer PECVD SiNx has potential benefit over thermal SiO2 due to its

higher refractive index, which reduces the optical reflection of solar cells at an optimal

thickness in air or in modules encapsulated with glass.

The PECVD SiNx refractive index, extinction coefficient and thicknesses vary depending on

the conditions of the deposition parameters (power, pressure, temperature, gas flow, etc.). The

amount of silicon in the SiNx film dictates both the refractive index and the absorption

coefficient, as will be discussed in the following paragraph. Each PECVD system would,

therefore, require calibration to achieve the desired SiNx deposition conditions.

The refractive index of SiNx could be represented as the bonding-density-weighted linear

combination of reference refractive indexes at N/Si = 0 (i.e., amorphous silicon) and at N/Si =

4/3 (i.e., Si3N4) [110]. By neglecting the hydrogen bonds in the SiNx, Bustarret et al. proposed

that this ratio could be calculated using Eq 2.1 [111]:

= 1.33(3.3 − 632)632 − 0.5 (2.1)

n632: refractive index measured at 632 nm

Page 27: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

14

As discussed previously is section 2.2.2, the extinction coefficient, k, is the sum of scattering

coefficient and absorption coefficient (α). When the extinction is only due to absorption (i.e.,

no scattering), the relationship between α and k is presented in Eq 2.2:

∝= 4 (2.2)

λ: measured wavelength k: extinction coefficient

A compilation of different SiNx with N/Si ratios using Eq 2.1, and extinction coefficients using

Eq 2.2, are plotted in Figures 2.6a and 2.6b. It is observed that an increasing N/Si ratio causes

both the n and k values to decrease as the SiNx become more N-rich.

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.61.6

2.0

2.4

2.8

3.2 Bustarret et al.

Verlaan et al.

Verlaan et al.

Kessels et al.

This work

Bustarret Model

Re

fra

ctive

In

de

x (

n632)

N/Si Ratio

1.8 2.0 2.2 2.4 2.6 2.8 3.0 3.210-4

10-3

10-2

10-1

100(b)

Wan

Duttagupta et al.

Doshi et al.

van Erven et al.

This work

Extin

ctio

n c

oe

ffic

ien

t k a

t 3

60

nm

(k 3

60)

Refractive Index (n) at 632 nm

(a)

Fig 2.6: (a) Correlation between the refractive index n632 and the N/Si ratio for various SiNx films [111-113], and (b) the correlation between the extinction coefficient k measured at 360 nm (k360) and the refractive index n632 [77, 114-116].

2.3.3 Surface passivation of PECVD silicon nitride

The successful adoption of PECVD SiNx for commercial use in silicon solar cells resulted from

a variety of advantages, including a low-temperature deposition, atomic hydrogenation of the

surface and bulk [106], field-effect passivation from high positive charges and excellent ARC

properties. The PECVD SiNx has been reported to achieve a low Dit of ~1.8 × 1010 eV-1cm-2 on

a 1 Ω.cm n-type silicon substrate [117], which is comparable to the FGA of thermal SiO2. The

passivation of PECVD SiNx applied on the front surface of an IBC solar cell at ANU achieved

excellent surface passivation with a Voc of 703 mV in addition to the advantages of the

Page 28: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

15

transparent SiNx ARC properties with a low absorption coefficient [14]. The excellent surface

passivation of SiNx deposited with Roth Rau AK400 dual mode plasma reactors were thought

to arise from an experimental artefact where the radio frequency (RF) substrate bias ignites an

unintentional deposition plasma before the ignition of the main microwave plasma [118]. The

RF plasma deposited a Si-rich layer, intervening SiNx layer in a film otherwise stoichiometric

SiNx [118].

As the solar industry utilises SiNx for surface passivation, continuous effort is being devoted

to reducing the surface recombination. The Seff of PECVD SiNx with various wafer resistivities

and refractive indexes from literature [14, 17, 115, 119-122] were compared to the optimised

PECVD SiNx deposited using Roth Rau AK400 at the ANU, is plotted in Figure 2.7 [117].

10-1 100 101 102

10-1

100

101

102

Seff U

L (

cm

/s)

Si wafer resistivity (Ω.cm)

P-type 100

N-type 100

N-type 111

N-type rantex

P-type 100 - This work

N-type 100 - This work

N-type rantex - This work

Fig 2.7: A compilation of SiNx passivations for various silicon wafer resistivities from the literature [14, 17, 115, 119-122] (transparent symbols) compared to the PECVD SiNx deposited using PECVD Roth Rau AK400 (filled symbols).

2.4 Atomic layer deposition of aluminium oxide

ALD can produce thin films through a layer-by-layer process [123]. The reaction cycle of a

two-precursor process is commonly executed as the first precursor reacts with the surface to

form a monolayer of material. After the excess precursor and by-products are purged from the

first reaction, the second precursor reacts with the first previously deposited layer on the surface

in a sequential, self-limiting manner. Under controlled deposition conditions, the film thickness

can be varied by only adjusting the number of reaction cycles.

Page 29: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

16

There are two typical reactor configurations: Thermal and plasma-enhanced ALD (Pe-ALD).

Thermal ALD requires a certain temperature (temperature window) that is typically between

100-400 °C for reactants to react within the tight tolerances of processing conditions to achieve

the desired thickness. Temperatures outside the temperature window could decrease or increase

the growth of the film, either due to the lower thermal energy required for a reaction,

condensation of the precursor or from desorption or decomposition of the reactants. On the

other, PE-ALD uses plasma to enhance the reactivity of the reactants or influence the deposited

material characteristics. By increasing the reactivity of the reactants, quality films with high

growth rates can be achieved at relatively lower temperatures.

The deposition of Al2O3 with ALD is commonly performed using trimethylaluminum (TMA)

together with either O2 or H2O as the oxidizer. It has been reported that a thin layer of SiO2

forms between the silicon and the deposited Al2O3 [124-126]. The interface defects are reported

to be of the Pb0 type, which is similar to the Si-SiO2 interface on a 100 orientated wafer [127].

Furthermore, the Al2O3 film has been shown to release hydrogen when post-deposition

annealing (PDA) is performed above the deposition temperature, but the amount of released

hydrogen may be less than of PECVD SiNx [95, 128, 129].

2.4.1 ALD aluminium oxide charges

The high fixed charge density reported for Al2O3 is still a matter of discussion. The tetrahedral

coordination of AlO4 predominant at the silicon interface is considered a possible origin of the

fixed charge density [130-132]. The charges were reported to be located within the first few

nanometres above the silicon surface, and are independent of the Al2O3 thickness [130, 133].

Another explanation for the origin of the negative charges is due to interstitial oxygen atoms

caused by silanol rehydroxylation within the film [134].

2.4.2 Optics of ALD aluminium oxide

The refractive index n of ALD Al2O3 with a deposition temperature of ~200 ˚C on silicon is

approximately ~1.64 [135]. Kumar et al. similarly found ALD Al2O3 to have a n between 1.64–

1.67 on silicon with thicknesses of 200–400 nm [136]. However, Barbos et al. found the

refractive index of Al2O3 increases for thicknesses less than 10 nm [137]. The changes in the

Page 30: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

17

refractive index of the film are thought to be due to the strong influence of the interfacial SiOx

between the silicon and Al2O3 [137].

2.4.3 Surface passivation of aluminium oxide

As ALD Al2O3 is less often employed commercially due to its slow deposition rates and low

throughput, the development of deposition alternatives for Al2O3 to increase throughput

includes PECVD, Spatial ALD and atmospheric pressure chemical vapour deposition

(APCVD). However, ALD Al2O3 is utilised extensively in laboratories to achieve excellent

surface passivation.

10-1 100 101 102

10-1

100

101

102

N-type 100 - this work

N-type 111 - this work

Seff U

L (

cm

/s)

Si wafer resistivity (Ω.cm)

P-type 100

N-type 100

Fig 2.8: A compilation of the data for Al2O3 passivation on various resistivities from the literature (transparent symbols) [17, 138-140] compared to Al2O3 undertaken using plasma ALD on a Beneq TS 200 (filled symbols).

The ALD Al2O3 provides both excellent chemical passivation and field-effect passivation from

negative charges. The Al2O3 deposited on the front of a boron diffused surface with ALD and

a backend phosphorus-doped polysilicon achieved an efficiency of 24.7% with a Voc of 703

mV [141]. Black et al. reported a measured Dit for ALD Al2O3 of ~4 × 1010 eV-1cm-2 on a 1

Ω.cm p-type silicon substrate [142], which is slightly higher than the average Dit for

hydrogenated thermal SiO2. A compilation of Seff.UL of Al2O3 passivation from literatures are

compared with the optimised Al2O3 deposition employed in this dissertation using the ALD

Beneq TS 200, as shown in Figure 2.8 [17, 138-140].

Page 31: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

18

2.5 Doped polysilicon passivated contacts

Carrier selective passivation of contacts has attracted significant attention within the PV

research community. Doped polysilicon or tunnel oxide passivated contact (TOPcon) are

among the many technologies in this category, which are of high interest due to their excellent

electrical performance, commercial scalability and compatibility to existing silicon PV

production technologies. They typically consist of an ultra-thin SiO2 layer at 1-2 nm and a

heavily doped (phosphorus or boron) silicon layer with a polycrystalline phase. The IBC solar

cells utilising doped polysilicon technology recently achieved efficiency of 26.1% [11] . The

key advantage of doped polysilicon contacts is that they form a well passivated surface that

also enables carrier extraction through the thin SiO2 layer [143-145].

The thin SiO2 layer can be grown either chemically [146], using ozone oxidation [147] or in a

high-temperature furnace [148] for a short duration. The amorphous silicon is deposited either

through PECVD or with low-pressure chemical vapour deposition (LPCVD) and is doped and

annealed at high temperatures to form the doped polysilicon with either thermal diffusion [148,

149], spin-on doping, or ion implantation [150]. The thin SiO2 layer between the doped

polysilicon and the silicon substrate provides chemical passivation and acts as a tunnel oxide

for holes or electrons to tunnel through [151, 152], or as a low-density pinhole layer that allows

carriers to channel through [151]. The doped polysilicon provides a field-effect passivation that

can induce a band bending at the polysilicon-silicon interface which prevents recombination of

minority carriers onto the surface [153]. The surface passivation can be further improved with

hydrogenation process of the polysilicon structure [154-156].

2.5.1 Optics of polysilicon

A thick polysilicon layer annealed at high temperature has a relatively similar refractive index

as crystalline silicon [157]. However, the optical properties of thin polysilicon layers have not

been well characterized. Laghla et al. showed the refractive index of polysilicon varies with its

thickness [158]. Using LPCVD, they observed variations in the polysilicon refractive index

due to surface roughness with increasing deposition time. Due to the absorptive nature of doped

polysilicon at visible and infrared wavelengths, high-efficiency cells have polysilicon

deposited mostly at the back side to avoid absorption losses [159].

Page 32: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

19

2.5.2 Passivation and contact resistivity of doped polysilicon

The growth conditions of silicon oxide between the doped polysilicon and silicon are important

for the quality of the surface passivation and contact resistivity for a solar cell. The silicon

oxide should be sufficiently thick to provide excellent passivation while thin enough for

carriers to flow. Therefore, a trade-off is commonly observed between the quality of surface

passivation and contact resistivity [149, 160].

Phosphorus-doped polysilicon with a chemical oxide or thermal SiO2 has been reported as

having excellent properties. Yan et al. reported phosphorus-doped polysilicon with a SiOx at

1.4 nm thick that achieved a saturation current density (J0) ~2 fA/cm2 and contact resistivity of

~3 × 10-3 Ω.cm2 [141]. Fong et al. reported a thermal SiO2 that was under 1.5 nm achieved a

J0 ~3 fA/cm2 and contact resistivity ~7 × 10-4 Ω.cm2 [161]. A compilation of the J0 and contact

resistivity for passivated contacts with phosphorus-doped polysilicon is plotted in Figure 2.9a.

10-3

10-2

10-1

100

101

102

103

100

101

102

103

10-3

10-2

10-1

100

101

102

103

100

101

102

103

Satu

rate

d c

urr

ent

J0 (

fA/c

m2)

Contact resistivity (mΩ.cm2)

Gan et al.

Feldmann et al.

Peibst et al.

Yan et al.

Fong et al.

(a) (b)

Satu

rate

d c

urr

ent

J0 (

fA/c

m2)

Contact resistivity (mΩ.cm2)

Gan et al.

Feldmanns et al.

Yan et al.

Peibst et al.

Fig 2.9: Doped polysilicon comparing the measured contact resistivity and the saturated current J0 for (a) doped phosphorus [149, 151, 162, 163] and (b) doped boron [148, 151, 162, 163].

A compilation of boron-doped polysilicon passivated contact results from the literature is

plotted in Figure 2.9b and shows an overall slightly higher J0 and contact resistivity as

compared with phosphorus-doped polysilicon [148, 151, 162, 163]. To the author’s knowledge,

the report of Peibst et al. is one of the best for boron-doped polysilicon with thermal SiO2,

which achieved a J0s of ~4 fA/cm2 with a contact resistivity of ~10-2 Ω.cm2 [151].

Page 33: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

20

2.6 Electret

An electret is a stable dielectric material that exhibits a quasi-permanent electric charge or

dipole polarisation. There are two classes of electrets: dipolar electrets and space charge

electrets. Dipolar electrets are electrically neutral but have a macroscopic electric dipole

moment (like a magnet), whereas space charge electrets have a net macroscopic electrostatic

charge (like a capacitor).

The two types of electrets are fabricated using different techniques. Dipole electrets are

constructed by heating a material above its glass-transition temperature and cooling it slowly

within a strong electric field. As the material solidifies, the individual molecular dipoles

become ‘frozen’ with a net orientation in the direction induced from the electric field. Space

charge electrets are fabricated by incorporating changes into the surface (or bulk) through a

bombardment with an electron beam, corona discharge or the direct contact with a charged

electrode.

2.6.1 Corona charging on silicon oxide and silicon nitride electret

Corona discharge is a process where an electrode with a high potential ionizes the air around it

to create a plasma. The ions (ionized air molecules) pass charges to a nearby lower potential

region or recombine to form a neutral gas molecule. A corona charger is therefore used to

deposit charges onto the surface of a wafer.

A conventional corona charger has a needle that is placed several centimetres above the desired

sample [164]. A high potential is then applied to the source needle, which ionizes the

surrounding air molecules. Depending on the polarity of the applied potential, the majority of

the ionized air molecules, of either COor H3O+, are deposited onto the surface of the sample

[165]. A mesh grid with lower potential between the sample and the needle can be used to

control the amount of charges and their distribution on the surface [164].

The formation of a silicon oxide electret can benefit due to both its excellent chemical

passivation and additional field-effects to reduce the surface recombination [65, 166, 167]. An

ideal corona charged silicon oxide electret should retain the deposited charges to render it

useful; however, there are reports of the corona charge electrets decaying at high temperatures,

Page 34: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

21

humidity, light exposure, and bulk and lateral conduction of the dielectric material [168-172].

Nonetheless, there are ways to overcome certain decay mechanisms, such as by reducing the

surface conduction by using hexamethyldisilazane (HMDS), while additional annealing can

drive surface charges into the bulk of the dielectric [167, 170, 172].

Corona charging to passivate dielectrics can be of benefit by inducing a field-effect via either

accumulation or inversion to reduce the surface recombination [65, 166, 167, 173, 174].

However, as mentioned, corona charging treatments of surfaces are known to decay over time.

The silanol group at the surface of the SiO2 was reported by Voorthuyzen et al. to be

responsible for the lateral conduction of charges at the surface, either due to mobile protons

from these groups or the ability of the groups to attract water molecules [172]. Olthuis et al.

showed that the lateral conduction of charges can be reduced by modifying the polar silanol

surface with a terminating apolar methyl group using HMDS [168]. They also acknowledged

that although using HMDS may reduce the surface conduction, the silanol group within the

bulk of the silicon oxide may pose a threat to the long-term charge stability in humid

environments.

Other methods of charge storage with SiO2 include annealing positive charges at temperatures

of 250–400 ˚C [65, 167, 170]. Minami et al. showed that annealing a corona-charged surface

drives the charges into the bulk and is able to withstand high humidity for long periods of time

[170]. Although Minami et al. performed various characterizations to investigate the effect of

annealing on the charge storage, the results were inconclusive, and the underlying mechanism

was not well understood. Other methods, including capping the SiO2 with Si3N4 (SiO2-Si3N4

structure) or Si3N4-SiO2 (SiO2-Si3Nx-SiO2 structure) have been reported to be superior at

retaining positive charges compared with a single SiO2 layer at 300 ˚C [175]. Zhang et al.

showed that annealing causes the trapped charges to drift from the surface of the ON structure

into the dielectric bulk in a similar manner as SiO2 [176].

2.6.2 Surface passivation of corona charged silicon oxide

As previous discussed in Chapter 2.2.1, thermal SiO2 has a low density of fixed oxide charges.

To demonstrate the outcome of the field-effect passivation, corona charging was performed on

a high resistivity 100 Ω.cm silicon wafer with thermal SiO2 surface passivation.

Page 35: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

22

The surface passivation improves (reduced J0s) with an increasing positive corona charge

deposited onto the SiO2 surface, as plotted in Figure 2.10. The accumulation of majority

carriers at the surface prevents minority carriers from recombining. Corona charging over 80

seconds has shown to have a reduced lifetime; however, the J0s remains relatively constant,

which suggests that corona charging may have cause bulk degradation via dehydrogenation.

Such observations are not new, and other research groups have observed similar occurrences

related to dehydrogenation [177, 178].

0 20 40 60 80 100

4

8

12

16

20

τeff

(∆n ~ 1015

)

J0s

Corona charging duration (s)

τ eff

(ms)

4

8

12

16

20

J0s (fA

/cm

2)

N-type 100 Ω.cm

Fig 2.10: Accumulative corona charging for an n-type 100 Ω.cm FZ wafer with at a 406-µm thick thermal SiO2.

These results demonstrate that the field-effect improves the surface passivation; however,

caution has to be taken so that sufficient charge is deposited without degrading the surface or

bulk via dehydrogenation. The use of corona charging will be extended onto surface

passivation with an ONO stack as discussed in subsequent chapters.

2.7 Chapter summary

This chapter discussed and summarised the various methods of surface passivation commonly

used in silicon solar cells. Thermal SiO2 can attain a low Dit after hydrogenation with relatively

low fixed oxide charges. PECVD SiNx can achieve a low Dit and high positive charges, whereas

ALD Al2O3 can similarly achieve a low Dit and high negative charges. The charges create a

field-effect passivation that may accumulate or invert the surface.

Page 36: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

23

The following chapters present the fabrication, optimisation and implementation of ONO stack,

which includes thermal SiO2, PECVD SiNx and SiOx, together with corona charging to

investigate the best methodology to improve chemical and field-effect passivation in IBC solar

cells.

Page 37: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

24

Chapter 3

Characterisation methods

This chapter discusses the characterisation methods used throughout this work, which include:

1. Characterisation of the recombination in the silicon bulk, surface, and diffused region

2. Characterisation of the charge density in dielectrics

3. Current-voltage (I-V) measurements and analysis of I-V data in solar cells.

3.1 Photoconductance measurement and recombination mechanism in

solar cells

Photoconductance measurements performed on wafers determines the decay in the

conductivity after illumination to calculate the minority carrier lifetime [179, 180].

The initial conductivity of a silicon wafer is measured with the wafer placed above a coil to

form a bridge circuit. A flash is then applied to the wafer, causing an increased conductance.

As the flash intensity decreases, the differences in the decrease conductivity can be used to

calculate the minority carrier concentration using Eq 3.1

∆ = ∆ ! + !# (3.1)

µn: mobility of electrons µp: mobility of holes w: thickness of wafer ∆σ: conductivity

With known carrier concentration given by Eq 3.1, the effective lifetime can be calculated

using Eq 3.2

$%&& = ∆' − (∆()

(3.2)

where ∆n is the time-dependent average excess carrier density and G is the photogeneration

rate for the electron-hole pairs.

Page 38: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

25

There are two different types of measurement modes: quasi-steady state (QSS) and transient.

During QSS measurements, the illumination period is long relative to the lifetime. Thus, the

decay of the excess carrier concentration can be ignored, and the lifetime measurement can be

calculated as in Eq 3.3. While the QSS mode is not limited by the lifetime, the photogeneration

rate is required for accurate measurements.

During transient measurements, the illumination period is shorter than the lifetime. Thus, the

photogeneration rate can be ignored, and the lifetime measurement can be calculated as in Eq

3.4. The transient mode is limited by the measured lifetime, as it requires the lifetime to be

much longer than the decay of the illumination period (commonly with measured lifetimes

above 200 µs).

$%&& = ∆' (3.3)

$%&& = ∆(∆() (3.4)

All photoconductance measurements performed in this thesis were in transient mode because

the measured wafers were well-passivated and have relatively high lifetimes. The measured

effective lifetime can be modelled to represent the surface, bulk and emitter recombination as

given by Eq 3.5 because the minority carrier diffusion length is large compared with the wafer

thickness. Each of the recombination will be discussed in the following sections.

1$%&& = 1$*+,- + 1$.+/&01% + 1$%2344%/ (3.5)

3.1.1 Bulk recombination

There are three main fundamental recombination mechanisms within bulk silicon material:

Radiative, Auger and Shockley-Read-Hall.

Radiative recombination occurs when an electron from the conduction band recombines with

a hole in the valance band and emits a photon. This recombination dominates in direct bandgap

semiconductors such as GaAs; however, an indirect bandgap material such as silicon has low

radiative recombination relative to other modes that is unlikely to impact the overall lifetime.

Page 39: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

26

Auger recombination occurs when an electron recombines with a hole by releasing excess

energy to a third charge carrier (electrons in the conduction band or holes in the valance band).

The kinetic energy given to the third charge carrier is lost as phonons when the excited carrier

relaxes to its original energy state via ‘thermalisation’. Three charge carriers are involved in

Auger recombination: either two electrons and a hole, or two holes and an electron. Therefore,

the recombination rate is proportional to the product of the concentration of the three carriers.

An empirical parameterisation based on experimentally measured lifetimes is often used

instead to describe the Auger recombination rate due to difficulties in theoretically determining

the Auger recombination parameters. The parameterised Auger recombination rate

implemented in this work can be written as shown in Eq 3.6 [17].

$34/,%6 = ∆7(2.5 × 109:%%6; + 8.5 × 10=:%667; + 3.0 × 10=>∆;.>= + ?,@A (3.6)

:%%6(;) = 1 + 13 B1 − tanh GH ;3.3 × 109IJKL;.MMNO (3.7)

:%66(7;) = 1 + 7.5 B1 − tanh GH 7;7.0 × 109IJKL;.MNO (3.8)

Blow: Radiative recombination coefficient for lowly doped and lowly injected silicon 4.7 × 10-15 cm3s-1 at 300 K [181]. n0: Thermal equilibrium of electron density p0: Thermal equilibrium of hole density

Crystal defects and impurities introduce energy states or defect levels within the bandgap,

which allows a free electron within the conduction band to transition into the defect level before

recombining with a hole in the valance band. The recombination rate depends on the electron

and hole concentrations, as well as the properties and density of the defect. The recombination

can be modelled using Shockley-Read-Hall (SRH) statistics and for a single defect level, as

shown in Eq 3.9 [182].

QRST = U46V(7 − 3=) + 9 + 7 + 79= 7 − 3=$;( + 9) + $;(7 + 79) (3.9)

$; = 1VU46 (3.10)

$; = 1VU46 (3.11)

Page 40: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

27

9 = 1WX7 YZV − Z1[ \ (3.12)

79 = ]WX7 YZ] − ZV[ \ (3.13)

: capture cross section of holes : capture cross section of electrons V: density of defect U46: thermal velocity charge carriers ET: energy level of the defect n1: electron density when Fermi level coincides with defect energy level ET

p1: hole density when Fermi level coincides with defect energy level ET Nc: effective densities of states in the conduction band Nv: effective densities of states in the valance band

The SRH bulk lifetime for n-type silicon can be derived from Eq 3.9 and expressed as Eq 3.14.

$RST ≈ $;( + 9) + $;(7 + 79)_ + ∆ (3.14)

3.1.2 Surface recombination

Dangling bonds on the surface can give rise to a large density of defect levels within the silicon

bandgap. Unlike bulk SRH centres, which occupy a single energy state, multiple states can

exist on the surface throughout the bandgap. The surface recombination can be modelled using

SRH statistics from the bulk by replacing NT with Dit and integrating over the entire bandgap,

as given by Eq 3.15 [182].

QR = ` U46(.7. − 3=). + 9(Z) + 7. + 79(Z)a1

a] b34(Z)(Z (3.15)

ns: electron density at the surface ps: hole density at the surface

The surface values of ns, ps, σp, σn and Dit are generally not known; however, the surface

recombination in Eq 3.15 can be further simplified using a surface recombination velocity (Seff)

with the respective carrier concentration at the surface, as shown in Eq 3.16. Alternatively, for

a highly charged surface passivation, the surface recombination can be described using a

saturation current density J0s as in Eq 3.17 [183]. The J0s can be measured using Kane and

Swanson method from photoconductance measurements of symmetrical passivation on silicon

wafers [184]. In most of our work, J0s is used to describe the surface passivation of an ONO

Page 41: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

28

stack as it satisfies the condition Q2/Ns >1.5 × 107 cm. The advantage of J0s over Seff is that J0s

is carrier injection independent within the measured range of excess carrier concentration [183].

QR = %&&∆. (3.16)

QR = 3= × c;.(7.. − 3=) (3.17)

When a wafer has a high bulk lifetime and the excess carrier concentration is uniform

throughout the sample, the surface carrier concentration is assumed to be similar to the bulk

carrier concentration (Eq 3.18) or for a highly charged surface passivation (Eq 3.19).

$.+/&01% = d2%&& (3.18)

$.+/&01% ≈ 3=d2c;.(e + ∆) (3.19)

It should be noted that Kane and Swanson method does have limitations. One of the limitations

includes assuming τbulk is constant across the measured carrier injection level. Such limitation

is observed in some of the measured J0s presented in Chapter 4.

3.1.3 Emitter recombination

Recombination within the emitter depends on several factors, including the minority carrier

concentration, Auger recombination, SRH recombination, carrier mobility, effective masses

and the band gap narrowing. Moreover, the surface concentration from the emitter doping

affects the surface recombination. Cuevas et al. [185, 186] solved the continuity equation and

used a 3rd order analytical approximation to calculate the recombination of the emitter, as

shown in Eq 3.20. However, accurate calculation of J0e requires details of the surface

recombination velocity SP, the diffusion profile and the Auger lifetime τp within the emitter,

which are generally not well known.

c;% = 7;(d)f1 + ?=()g + h (X 7;(X)$(X) i=(X)j;1 + h (X 7;(X)$(X) k9(X) + 7;()fk9() + k()gA;

(3.20)

Page 42: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

29

k() = ` (X9 1b(X9)7;(X9)` (X= 7;(X=)$(X=) ` (X 1b(X)7;(X)l=

;l9

;j

;

k9() = ` 1b(X9)7;(X9)j

; (X9 (3.21)

Nevertheless, the measured values of J0e can be extracted using the Kane and Swanson method

from photoconductance measurements of a high resistivity wafer with symmetrical diffusion,

as shown in Eq 3.22 [184].

$%&& ≈ 3=d2c;%(e + ∆) (3.22)

The reported J0s and J0e in this thesis were an average of at least of 3 photoconductance

measurements.

3.2 Capacitance-voltage measurement

Capacitance-voltage (C-V) measurements of metal-oxide-semiconductor (MOS) capacitors are

widely used to measure the charges within dielectrics and the Dit of MOS structures. The MOS

structure is a two-terminal device composed of a metal deposited on top of a dielectric oxide

with the rear contact on grounded silicon. The following discusses the MOS structure for

thermal SiO2.

Thermal SiO2 may contain various charges trapped within the dielectric. As discussed in

Chapter 2, thermal SiO2 charges can be categorized as either fixed charges (Qf), oxide-silicon

interface charges (Qit), oxide trapped charges or mobile oxide charges. The fixed charges

within thermal SiO2 are generally positive and can be reduced using POA in either N2 or Ar

gas. The interface trap charges are either positive or negative and are caused by either dangling

bonds at the Si-SiO2 interface, oxidation-induced defects or metal impurities. The interface trap

charges within the Si-SiO2 interface can be reduced via hydrogen annealing at low temperatures.

Oxide trapped charges are either positive or negative and are the result of holes and electrons

within the thermal SiO2. The causes of such charges include ionizing radiation, avalanche

injection or they occur during the operational conditions of the device. Mobile oxide charges

can also consist of ionic impurities within the oxide, such as sodium, lithium, potassium or

hydrogen ions. These impurity contaminations are commonly found within the oxide due to

Page 43: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

30

exposure during equipment cleaning processes that use chemicals including vacuum pump oil

and chemical-mechanical polish. The ion contaminations on the wafers can be reduced using

Radio Corporation of America (RCA) cleaning.

In an ideal condition when no charges are presented within the oxide, the flatband voltage (VFB)

is equal to zero. However, the presence of charges commonly found within the oxide (Qox)

shifts the VFB. Under such condition, the flatband condition in a MOS structure illustrated in

Figure 3.1, occurs when no charges are present within the semiconductor. The solution of Eq

3.23 can provide the quantity of charges within the thermal SiO2 at the flatband voltage.

Fig 3.1: MOS capacitor structure with the (a) energy band diagram and (b) charge density under flatband conditions

m@l = i@l nop × (@lk((@l − X) (3.23)

A: Area of metallised region q: Electron charge of 1.602 × 10-19 C Ф2.: Workfunction differences of metal and silicon Cox: Oxide capacitance VFB: Flatband voltage

During a C-V measurement, the VFB could be obtained from the calculated flatband capacitance

(CFB) as shown in Eq 3.24 and 3.25.

Ef,m

Фm

Qox

Ef,Si

n-type Si

SiO2

Metal

Evac

EC

EV

ρ(x)

(a)

(b)

χ

dox

qVFB

E

i

x

Page 44: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

31

1iop = u_k. vR3 + 1i@l (3.24)

u_ = wvR3[=_ (3.25)

The LD in Eq 3.25 is defined as the Debye length, which is known as an ‘effective shielding

distance’ and is described by Evans as a ‘characteristic distance that some external electric field

can penetrate a neutral semiconductor surface without substantially perturbing the

semiconductor away from neutrality.’ [187].

The work function difference between a metal and silicon can be determined using Eq 3.26.

Ф2. = nop = Ф2 − x. − (Z1 − Z3) + Z& − Z3# (3.26)

Z1 − Z3 = [ y Y13 \ (3.27)

Z& − Z3 = [ y Y_3 \ (3.28)

Ф2: Workfunction of metal x.: Electron affinity Nc: Effective densities of states in the conduction band Nd: Concentration of donor doped atoms

3.3 Current-voltage measurement

Current-voltage (IV) measurements are used to establish the performance of a solar cell. The

cell performance can be analysed from each I-V curve to determine the short circuit current

(Jsc), open circuit voltage (Voc), fill factor and the efficiency of the cell. Other information, such

as the series and shunt resistances, non-ideality factor(s) and temperature coefficient, can be

extracted from multiple I-V measurements under different lighting and temperature conditions.

The I-V measurements for an IBC solar cell were performed using a custom designed

measurement jig with an independently measured aperture mask of 3.99 ± 0.01 cm2. The layout

of the jig is presented in Figure 3.2 and has multiple probes that can measure six IBC solar

cells on each wafer. Each cell has eight probes in total; six current and two voltage probes. The

middle voltage and current probes are 3 mm apart, whereas the distance between the current

probes at a similar busbar polarity is 6.7 mm. Each of the current probes is connected to a

Page 45: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

32

resistor for uniform current distribution across each pin. The jig has a thick Al base and four

Peltier coolers and fans to control the overall temperature while taking the measurements. The

jig was anodised black to avoid light reflections that may cause artificial increases in the Jsc.

The cells loaded onto the jig were held in place with a vacuum pump and the aperture mask

that defines the active cell area was made from a silicon wafer. The mask was painted with

black matt paint on the front and a 100 nm evaporated Al layer at the rear to reduce the

reflection of light back into the light source (mirrors are the light source) and to prevent infrared

wavelength into the cell.

Fig 3.2: Layout design of the cells placed onto the jig with metal contact probes.

Listed below are the steps performed to achieve accurate I-V measurements:

1. For a halogen light source, the lamp was pre-warmed for 30 minutes to an hour. The jig

temperature was set and stabilised for at least ten minutes before the measurements.

Figure 3.3 shows the measured surface temperature of the jig using a calibrated

thermocouple with time at different pre-set temperatures. The pre-set temperatures on

the controller were measured to be 1–2 ˚C lower than the desired jig temperature; thus,

a pre-set temperature of 26.2 ˚C was set to achieved ~25 ˚C on the jig.

2. The current density of a calibrated reference cell was measured at a defined area under

the illuminated light source. The light intensity of the halogen lamp was adjusted to

match the specified current density of calibrated reference cell at 1 sun.

TLM structure for boron

TLM structure for phosphorus

Dielectric isolation pads

Metal line resistance

Page 46: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

33

3. The IBC solar cell, together with the aperture mask, was placed in a similar area and at

a similar height to the previous reference cell location.

4. Measurements of the I-V curve were performed by sweeping the voltage from 0.8 to -

0.05 V using a Keithly source meter. To avoid an unnecessary increase in the cell

temperature caused by the illumination of the light source, an inbuilt shutter was

programmed to open just before the cell measurement.

5. After measuring all cells, the light intensity was remeasured with the calibrated

reference cell at cell measurement area to account for any drift in the halogen light

source.

0 500 1000 1500 2000

20

21

22

23

24

25

26

27

Pre-set

to 26 °C

Pre-set

to 23 °C

Calibrated thermocouple top of stage

Internal thermocouple

Te

mpe

ratu

re (

°C)

Time (s)

Pre-set to 27 °C

Fig 3.3: Different pre-set jig temperatures with measured internal (underneath the Al plate) and external (top of the stage) thermocouples.

The series resistance (Rseries) of a solar cell can be determined using either multiple light I-V

method or comparing light and dark I-V measurements [188]. The measured series resistance

from an I-V curve includes resistance from the bulk, metallised fingers and busbar as well as

the contact resistance of the metal and silicon interface. Figure 3.4a plots the light and dark I-

V measurements for an IBC solar cell, from which the Rseries can be extracted at specific

voltages, as shown in Figure 3.4b, using Eq 3.27. It is noted that the extracted Rseries can be

viewed as a function of the voltage. Therefore, the series resistance, which is commonly

represented as a lumped Rseries, is oversimplified. The shunt resistance in a solar cell is often

due to imperfections of fabricate solar cell device which provide an alternative path for the

light-generated current [189, 190]. The shunt resistance (Rshunt) can be determined from the

Page 47: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

34

dark I-V measurements at low voltages of 0–0.1 V, as shown in Figure 3.4a. However, the

accuracy of measuring Rshunt is limited with non-ideal recombination at low voltages.

zW|W ≈ n(| − ny:ℎ)cJ (3.27)

0 200 400 600 80010

-3

10-2

10-1

100

101

102

0 200 400 600 8000

1

2

3

4(b)

Cu

rre

nt

(mA

/cm

2)

Voltage (mV)

Dark IV

Shifted light IV J

RShunt

J0

Mpp

Fitting Rs

Se

rie

s R

esis

tan

ce

.cm

2)

Voltage (mV)

Mpp

(a)

Fig 3.4: Measurement of (a) the shifted light and dark I-V curves with the simulated shunt resistance and J0, and (b) the extracted Rs measurements from the shifted light and dark I-V curves using Eq 3.27.

Page 48: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

35

Chapter 4

ONO: Growth and deposition conditions

4.1 Introduction

Excellent surface passivation is predominately achieved by having both exceptional chemical

and induced field-effect passivations. This chapter discusses the development of an ONO

surface passivation stack, which consists of a thin thermal SiO2, followed by a PECVD SiNx

and SiOx stack. The initial thermal SiO2 layer provides chemical passivation from its low Dit

between Si-SiO2. The subsequent PECVD SiNx layer enhances the passivation with a field

effect from positive charges and hydrogenates the Si-SiO2 interface. The final PECVD SiOx

layer improves the overall optical properties of the stack in air while retaining additional

positive charges from corona charge deposition.

This chapter begins with the baseline fabrication process sequences used in the formation of

ONO stacks, which include thermal oxidation, the deposition of PECVD SiNx and SiOx, FGA

and corona charging. This is followed by sections that describe the optimisation study of the

film deposition by varying the growth and deposition conditions of each individual layer, while

simultaneously evaluating both the passivation quality and the optics of the stack. Finally, this

chapter summarises the optimal passivation conditions performed on undiffused, n-type,

silicon wafers to achieve lifetimes that exceed the commonly accepted intrinsic recombination

model.

4.2 ONO baseline procedure

High resistivity FZ n-type 100 Ω.cm silicon wafers with a 100 orientation were etched in

TMAH for 10 minutes at 85 ˚C to remove any saw damage, cleaned using the standard RCA

process and subsequently dipped in a 1% HF solution until the surfaces became hydrophobic.

Thermal oxidation was conducted in a quartz tube furnace under ambient O2. The wafers were

loaded at 700 ˚C and ramped up to 1000 ˚C at a rate of 15 ˚C/min in O2. After reaching 1000

˚C, the wafers were annealed in N2 at the same temperature for 45 minutes followed by a ramp

Page 49: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

36

down to 700 ˚C. The thermal SiO2 thickness was measured as 15 ± 1 nm using an ellipsometer

(JA Woollam M2000D) [62].

After the thermal oxidation, amorphous SiNx and SiOx films were deposited individually with

PECVD using an Oxford Instruments PlasmaLab 100 deposition tool. For the SiNx depositions,

gas flowrates of 13 sccm for SiH4, 14 sccm for NH3 and 980 sccm for N2 were used at a

deposition temperature of 400˚C, a chamber pressure of 650 mT and forward plasma power of

20 W, which yielded an SiNx thickness of 50 ± 3 nm and refractive index at 632 nm (n632) of

~1.93. The PECVD SiOx depositions were performed at 250˚C with flows of 9 sccm for SiH4,

710 sccm for N2O and 161 sccm for N2, which produced films with a thickness of 90 ± 5 nm

and n632 of 1.48. The thickness of the SiNx and SiOx films was selected specifically for ARC

purposes based on simulations in OPAL2 [19]. All the wafers received a FGA at 400 ˚C in a

quartz tube furnace for 30 minutes.

The ONO wafers were then positively corona charged in a conventional setup without a mesh

grid using a steel needle with an applied voltage of 5 kV at an elevation of 8 cm above the

wafers [164]. The surface charge deposition was performed symmetrically on both surfaces,

and the wafers were then subjected to an additional FGA at 400 ˚C for 30 minutes to embed

the surface corona charges into the ONO stack [65, 167, 176]. The effective lifetime of the

wafers was measured before the positive corona charging and again after the anneal step.

4.3 Thermal oxide growth conditions for an ONO stack

This section investigates the effect of the thermal SiO2 thickness and duration of the post-

oxidation annealing (POA) on the passivation quality of the ONO stack. The performed

investigation is shown in the flowchart presented in Figure 4.1 with the details described in the

following subchapters.

Page 50: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

37

Fig 4.1: Flow chart detailing the variations for each layer (relative to the baseline) to establish the best passivation conditions of thermal SiO2 for ONO stack (a) The baseline procedure is represented by the white boxes, while variations to the oxidation temperature are highlighted in grey. (b) Variations to the POA time and ambient are highlighted in dark blue.

4.3.1 The effect of varying the thermal oxide thickness

The primary function of the thermal SiO2 is to provide a low Dit between the Si and the

passivating dielectric stack. A thermal SiO2 satisfies the basic criteria of having a low Dit

interface while providing excellent thermal stability. Further layers can be deposited on the

surface without negatively impacting its physical properties. The goal of the thermal SiO2

optimisation is to form a film that provides the desired low Dit characteristics with low

thickness to limit the optical losses associated with having a low refractive index material next

Embedding Charges

Forming Gas Anneal 400 °C

Planar 100, quartered

Thermal

Oxidation

850 °C

PECVD Silicon Nitride 400°C

PECVD Silicon Oxide 250 °C

Forming Gas Anneal 400 °C

Positive Corona Charging

Thermal

Oxidation

900 °C

Thermal

Oxidation

950 °C

Thermal

Oxidation

1000 °C

Nitrogen annealing 1000 °C

(a) Thermal SiO2 thickness

Planar 100, quartered

Thermal

Oxidation

1000 °C

Ramp down in

oxygen

PECVD Silicon Nitride 400 °C

PECVD Silicon Oxide 250 °C

Forming Gas Anneal 400 °C

Positive Corona Charging

Embedding Charges

Forming Gas Anneal 400 °C

Thermal Oxidation

1000 °C

Anneal and ramp

down in nitrogen

(b) Post-oxidation annealing in N2

Page 51: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

38

to the silicon surface [191, 192]. Therefore, this section examines the effects of varying the

thermal SiO2 thickness on the overall electrical performance of the ONO stack.

Experimental details:

A 100 Ω.cm n-type FZ silicon wafer with a thickness of approximately 400 µm was cleaved

into quarters. The growth and deposition of the ONO on all quarters were identical to the

baseline, except for the variable temperature for oxidation (and hence oxide thickness) as

described in Figure 4.1a. Each wafer was loaded in pure O2 at 700 ˚C and ramped to different

temperature set points. The oxide growth occurred during the ramps to 850, 900, 950 and 1000

˚C. Once the temperature set points were reached, the gas flow was switched from O2 to N2

and the wafers were subjected to further annealing at 1000 ˚C for 45 minutes in N2.

Results and discussion:

The measured surface saturation current density (J0s) and effective lifetime (τeff) for all the

wafers are presented in Figure 4.2. Two different sets of lifetimes are reported. The τeff reported

at an carrier injection level of 1015 cm-3 (τeff_1015) corresponds well to the operating maximum

power point of a high-efficiency solar cell, whereas the maximum lifetime (τeff.max) is reported

across measured injection level between 1013–1016 cm-3.

The experimental results indicate that a minimal oxide thickness of around 7 nm is necessary

for passivation levels of <5 fA/cm2. Further increases in the thermal SiO2 thickness do not

provide additional benefits, so this is considered the optimum for electrical and optical

performances of the ONO stack. For practical reasons, a target of 10 nm thick thermal SiO2 is

used when ONO is utilised in cell fabrication (Chapter 6) due to uncertainties in the process

controls.

In most cases, variations in the passivation layer thickness are known to affect the passivation

quality [193]. In general, thicker passivation layers (thermal SiO2, PECVD SiNx and ALD

Al2O3) perform better than thinner films [139, 192, 194, 195]. It is uncertain why the

passivation of thinner thermal SiO2 layers capped with either SiNx or SiNx-SiOx has a worse

performance when less than 10 nm. However, Black et al. suggested that thinner thermal SiO2

layers may be related to the deterioration of the chemical passivation [193]. Moreover, a similar

Page 52: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

39

trend was observed by Kerr et al. and Mack et al. for thermal SiO2 capped with PECVD SiNx

(ON stack), in which they reported no correlation of passivation quality with thermal SiO2

thicknesses above 10 nm [192, 194]. Nevertheless, there is a possibility that in this experiment,

the variations in the oxidation growth temperatures from 850 to 1000 ˚C to achieve the desired

thicknesses may affect the surface passivation.

0

5

10

15

20

25

30

35

τ eff

(ms)

τeff.max

τeff_1015

J0s

4 6 8 10 12 14 160

5

10

15

J0s (

fA/c

m2)

Thermal oxide thickness (nm)

Fig 4.2: Minority carrier lifetimes and J0s measured for ONO wafers with different thermal SiO2 thicknesses.

4.3.2 The effect of post-oxidation annealing of thermal oxide

The passivation of dielectrics with a large number of charges may benefit (accumulation of

majority carriers at the surface) or worsen (deplete majority carriers at the surface) the solar

cell performance [196]. As reviewed in Chapter 2, performing an anneal with N2 at

temperatures of 800–1000 ˚C reduces the number of positive charges in the thermal SiO2 [53,

57]. However, the impact of POA on the surface Dit of Si-SiO2 is more complicated. Razouk

et al. reported that the POA can either increases or reduce the surface Dit depending on the

wafer orientation, annealing duration and subsequent hydrogen annealing [30]. In this section,

POA with N2 on thermal SiO2 was examined to achieve a desired low surface recombination

on the ONO stack.

Page 53: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

40

Experimental details:

A 100 Ω.cm n-type FZ silicon wafer with a thickness of approximately 400 µm was split into

quarters. The growth and deposition of the ONO on all quarters were identical to the baseline,

except for the variable POA as detailed in Figure 4.1b. The wafers were loaded at 700 ˚C and

ramped up to 1000 ˚C in O2. After reaching 1000 ˚C, the wafers were annealed in N2 for 0, 15

and 30 minutes and then cooled in N2. A control quarter without N2 annealing was loaded at

700 ˚C, ramped up to 1000 ˚C and ramped down to 700 ˚C in O2.

Results and discussion:

The experimental results shown in Figure 4.3 indicate there were no improvements to J0s or τeff

as the duration of the POA increases. Instead, J0s increased slightly from 0.2 to 0.5 fA/cm2 and

τeff_1015 decreased slightly from 30 to 25 ms. It can be concluded that for these undiffused ONO

wafers, POA is not beneficial, and can even cause an increase in recombination.

0 10 20 30O2

0

10

20

30

40

50

60

τeff.max

τeff_10

15

J0s

N2 Annealing (minutes)

τ eff

(ms)

Ramp down in O2

0.0

0.4

0.8

1.2

1.6

2.0 J

0s (

fA/c

m2)

Fig. 4.3: Minority carrier lifetimes and J0s measured for ONO wafers with different thermal SiO2 annealing times in N2 or O2 ramp down

The source of the reduced lifetime and passivation quality may be caused by metal impurity

contamination (e.g. Fe, Cr, Ni) during the N2 anneal [197]. Alternatively, the reduced lifetime

may be due to the re-introduction of vacancy-like defects back into the silicon bulk during the

N2 anneal, as demonstrated by Abe [198].

Page 54: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

41

Figure 4.3 also shows that when ramp-down is performed in O2, the surface passivation is

slightly worse as compared to ramp-down performed in N2 (0.8 fA/cm2 instead of 0.2 fA/cm2);

however, the passivation exhibits a higher measured τeff.max (55 ms instead of 43 ms). This

observation reinforces the suggestion that using only ambient N2 during annealing contributes

to the formation of bulk defects. These findings are consistent with the work of Grant et al.

who demonstrated that vacancy related intrinsic defects (incorporated during FZ ingot growth)

can be annealed out during high-temperature oxidation, thereby improving the bulk lifetime of

FZ silicon [199, 200].

While an exact mechanism for bulk degradation resulting from the POA has not yet been

determined, with no POA (O2 rampdown) or 0 min POA (N2 rampdown), the performance of

a solar cell will be similar with operating conditions at ∆n ~1015 cm-3. Thus, the use of POA

under these conditions does not lead to any observable benefit from a device perspective.

4.4 Deposition conditions of PECVD for ONO stack

Stacking an additional dielectric layer onto the surface passivation layer of a solar cell is used

to increase the efficiency by improving the surface passivation, electrical isolation of the metal

contact and for ARC properties. The Al2O3 and SiNx stacked onto thermal SiO2 in some solar

cell designs improve the surface passivation by providing additional field-effect passivation

and hydrogenation of the Si-SiO2 interface [201].

This section initially investigates the surface passivation behaviour of the different dielectric

stacks of SiOx (OO), SiNx (ON) and SiNx and SiOx (ONO) onto thermal SiO2. The investigation

continues the development for the ONO stack by examining the deposition conditions for the

PECVD SiNx and SiOx layers. The investigations are shown in flowcharts as presented in

Figures 4.2 to 4.4 with the details described in the following subsections.

Page 55: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

42

Fig 4.4: Flow chart detailing the deposition of the PECVD SiOx and SiNx on thermal SiO2 to form the SiO2-SiOx (OO), SiO2-SiNx (ON) and SiO2-SiNx-SiOx (ONO) stacks.

Forming Gas Anneal 400 ᵒC

Planar 100

Thermal Oxidation and nitrogen annealing

at 1000 ᵒC

PECVD SiNx

400 ᵒC PECVD SiOx

250 ᵒC

PECVD SiOx

250 ᵒC

Positive Corona Charging

Various PECVD dielectric stack on thermal SiO2

ONO ON OO Control

Page 56: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

43

Fig 4.5: Flow chart detailing variations in each layer (relative to the baseline) to establish the best passivation conditions for PECVD SiNx. (a) The baseline procedure is represented by the white boxes, while variations to the gas flow ratio of SiH4 and NH3 are shown in the blue boxes. (b) The variations to the deposition temperature of SiNx (shown in green box). (c) The variation to chamber pressure of SiNx (shown in orange box).

Planar 100

Planar Rantex

Planar 111

TMAH TMAH and IPA

Thermal Oxidation and nitrogen

annealing at 1000 ᵒC

PECVD Silicon Nitride Deposition Temperature:

250 – 550 ᵒC

PECVD Silicon Oxide 250 °C

Forming Gas Anneal 400 °C

Positive Corona Charging

Embedding Charges Forming Gas Anneal 400 °C

(b) Deposition temperature of SiNx

Planar 100

Thermal

Oxidation

at 1000 ᵒC

and ramp

down in

nitrogen

PECVD Silicon Oxide 250 ᵒC

Forming Gas Anneal 400 ᵒC

Positive Corona Charging

Embedding Charges Forming Gas Anneal 400 ᵒC

(a) Gas ratio of SiNx

PECVD Silicon Nitride Gas Flow Ratio: SiH4: 8 - 24 sccm

NH3: 20 - 4 sccm

Planar 100

Thermal Oxidation and nitrogen

annealing at 1000 ᵒC

PECVD Silicon Oxide 250 ᵒC

Forming Gas Anneal 400 ᵒC

Positive Corona Charging

Embedding Charges Forming Gas Anneal 400 ᵒC

(c) Chamber pressure of SiNx

PECVD Silicon Nitride Chamber Pressure: 600 - 1000 mTorr

Thermal

Oxidation

at 1000 ᵒC

and ramp

down in

oxygen

Page 57: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

44

Fig 4.6: Flow chart detailing variations in the process for each layer (relative to the baseline) to establish the best passivation conditions for PECVD SiOx. (a) The baseline procedure is represented by the white boxes, while the variations to the gas flow ratio for SiH4 and N2O are highlighted in blue. (b) The variations to the deposition temperature of SiOx are highlighted in green.

Planar 100, quartered

Thermal Oxidation and nitrogen

annealing at 1000 ᵒC

PECVD Silicon Oxide

Gas Flow Ratio

SiH4: 4 - 40 sccm

N2O: 679 - 715 sccm

Forming Gas Anneal 400 ᵒC

Positive Corona Charging

Embedding Charges

Forming Gas Anneal 400 ᵒC

(a) Gas ratio of SiOx

PECVD Silicon Nitride 400 °C

Planar 100, quartered

Thermal Oxidation and nitrogen

annealing at 1000 ᵒC

PECVD Silicon Oxide

Deposition Temperature:

200 – 400 ᵒC

Forming Gas Anneal 400 ᵒC

Positive Corona Charging

Embedding Charges

Forming Gas Anneal 400 ᵒC

(b) Deposition temperature of SiOx

PECVD Silicon Nitride 400 °C

Page 58: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

45

4.4.1 PECVD dielectric stack on the thermal oxide

As previously discussed in Chapter 2, the PECVD SiNx stack onto thermal SiO2 is known to

provide an abundance of hydrogen, which help to passivate the dangling bonds of the Si-SiO2

interface and positive charges to create field-effect passivation [202].

The PECVD SiOx stack on thermal SiO2 has been reported to reduce the Dit of the Si-SiO2

interface after alneal (annealing in the presence of Al) with little change to the charge density

[192]. However, most of the hydrogenation sources from alneal is thought to be from FGA.

In this section, the OO, ON and ONO stacks, as well as the control thermal SiO2, were

examined.

Experimental details

A 100 Ω.cm n-type FZ silicon wafer with a thickness of approximately 400 µm was cleaved

into quarters. The growth and deposition of the ONO on one of the quarters was identical to

the baseline, except the wafer was not subjected to further annealing after corona charging.

Two other quarters with the baseline thermal SiO2 had either PECVD SiNx (ON stack) or SiOx

(OO stack) deposited. The PECVD SiNx and SiOx were similar to the baseline process. The

control quarter was not subjected to any PECVD deposition. No annealing was performed on

any sample after positive corona charging. The sequences of each process are illustrated in

Figure 4.4.

Results and discussion

The effective lifetimes of deposited dielectric OO, ON and ONO stacks and a control wafer

with only thermal SiO2 are plotted in Figure 4.7. Before corona charging, the measured surface

passivations of the ON and ONO dielectric stacks showed very similar lifetimes across the

measured excess carrier densities (∆n) range. The control thermal SiO2 passivation had the

lowest lifetime, and the OO passivated stack lifetime was between the thermal SiO2 and SiNx

capped (ON and ONO) wafers. This initial observation indicates that the SiNx capped onto

thermal SiO2 is beneficial to improve the surface passivation.

Page 59: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

46

1013

1014

1015

1016

0

10

20

30

40

1013

1014

1015

1016

0

10

20

30

40

Thermal Oxide

OO

ON

Eff

ective

Life

tim

e (

ms)

Excess Carrier Density (cm-3)

ONO

(a) (b)

OO

Thermal Oxide

OO

ON

Eff

ective

Life

tim

e (

ms)

Excess Carrier Density (cm-3)

ONO

Fig 4.7: Lifetime curves for the PECVD SiOx, SiNx and SiNx-SiOx deposited on thermal SiO2 to form OO, ON and ONO dielectric stacks (a) before and (b) after positive corona charging.

The various dielectric stacks and the control wafer all received additional positive corona

charges at the surface to induce the field-effect passivation. All of the passivation stacks

including the control thermal SiO2 passivation layer, were positively corona charged, till no

significant improvement was observed across measured lifetime curves. The corona charges

measured with a Kelvin probe were >3 × 1012 cm-2, assuming dielectric constants for the

thermal SiO2, PECVD SiNx and SiOx were 3.9, 7.0 and 5.0, respectively [203, 204]. The surface

passivations for all dielectric stacks improved with additional positive corona charges, as

plotted in Figure 4.7b. Two sets of lifetimes were observed, the first set with the ON and ONO

stacks and the second set with the OO and only the thermal SiO2 passivated wafer. Each of the

two different passivation schemes on each set had similar lifetimes across the measured range

of ∆n.

Based on these results, the higher lifetime first observed in Figure 4.7a for the OO stack relative

to the thermal SiO2 is most likely due to charges within the SiOx film. This is evidenced by

further corona charging as shown in Figure 4.7b, which improves the thermal SiO2 and OO

stack to a near identical surface passivation due to the addition of charge induced field-effect

passivation, suggesting otherwise equivalent chemical passivation. However, the measured

lifetimes of the corona charged thermal SiO2 and OO stack passivations are both lower

compared to wafers with SiNx. The key hypothesis in the improved lifetime seen on the SiNx

Page 60: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

47

wafers is that they benefit from the hydrogenation of the Si-SiO2 interface, which was

introduced during the PECVD SiNx deposition and subsequent FGA steps.

Therefore, the deposition of PECVD SiNx onto thermal SiO2 is crucial. It further improves the

surface passivation compared to PECVD SiOx on thermal SiO2. The mechanism in which it

further reduces the surface passivation for PECVD SiNx is thought to be atomic hydrogenation

at the Si-SiO2 interface rather than molecule hydrogenation introduced with FGA [106].

4.4.2 Deposition conditions of PECVD SiNx in ONO stack

The important parameters during the deposition of PECVD SiNx are the radio frequency (RF)

incident power, gas flow ratio, chamber pressure and temperature. These parameters affect the

optical properties and the deposition rate of SiNx [117, 194, 205].

In the following sections, the effects of the SiH4:NH3 gas ratio, deposition temperature and

chamber pressure on the electrical characteristics of the deposited PECVD SiNx are discussed.

A similar RF power was used in each experiment set to adequately ignite the plasma while

reducing the possibility of plasma damage and ion bombardment [205]. For each experiment,

only one parameter was varied while the other deposition conditions were held constant.

Experimental details

The experimental details for the deposition of PECVD SiNx are divided into three subsections.

Silane to ammonia gas flow ratio - To investigate the influence of the SiH4:NH3 gas ratio on

SiNx, a 100 Ω.cm n-type FZ silicon wafer with a thickness of approximately 400 µm was split

into quarters. The growth and deposition of an ONO stack on all quarters were identical to the

baseline except for two processing variations: i) the growth of thermal SiO2 and ii) the

SiH4:NH3 gas ratio.

For thermal oxidation, experiments were performed on wafers under two conditions. Under the

first set of conditions, POA with N2 ramp down was performed. The wafers were loaded at 700

˚C and ramped to 1000 ˚C in O2. After reaching 1000 ˚C, the supply of O2 was switched to N2

while cooling to 700 ˚C. The second set of conditions were wafers without POA. The wafers

Page 61: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

48

were loaded at 700 ˚C, ramped up to 1000 ˚C and ramped down to 700 ˚C while maintaining

in O2 ambient

In the second variation, the gas ratio of SiH4:NH3 was varied between 4 to 22 sccm. The

deposition time was varied to maintain the SiNx thickness under different gas flow ratios. For

antireflective purposes, the SiNx and SiOx thicknesses of 50 ± 3 nm and 90 ± 5 nm, respectively,

were maintained, as outlined in the baseline procedure. The performed investigation is shown

in the flowchart presented in Figure 4.5a.

Deposition temperature - To investigate the influence of the SiNx deposition temperature,

several FZ and Czochralski (Cz) n-type 100 Ω.cm silicon wafers featuring different surface

conditions were used: chemically polished FZ 100 planar, chemically textured FZ using a

TMAH and isopropanol (IPA) solution mixture to form rantex and mechanically polished Cz

111 planar surfaces. The wafer thicknesses for the 100 and 111 planar and the rantex

surfaces were measured to be approximately 400, 420 and 380 µm, respectively. The growth

and deposition of the ONO stack on all wafers was identical to the baseline, except for the

PECVD SiNx deposition temperature, which varied between 250–550 ˚C, as shown in Figure

4.5b. The deposition time was varied to maintain the SiNx thickness at different deposition

temperatures (as temperature increased, the deposition rate decreased and the refractive index

n632 increased from 1.9 to 2.0 [205]).

Chamber pressure - To investigate the influence of the chamber pressure on the deposition

of SiNx, a 100 Ω.cm n-type FZ silicon wafer with a thickness of approximately 400 µm was

split into quarters. The growth and deposition of the ONO stack on all quarters were identical

to the baseline, except for the chamber pressure, which varied between 600 to 950 mTorr, as

shown in Figure 4.5c. The deposition time was varied to maintain a similar SiNx thickness

across the different chamber pressure conditions.

Results and discussion

The results of each condition are discussed in the following subsections. The dependence of

the surface recombination, charge density and optical properties of the PECVD SiNx on the

various deposition parameters are shown in Figure 4.8.

Page 62: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

49

109

1010

1011

1012

100

101

ramp down O2

ramp down N2

τ eff (

ms)

(a) (c) (b)

τeff_10

15

J0e

0

5

10

15

20

J0s (f

A/c

m2)

ON

O Q

eff

(cm

-2)

N2

O2

O2

N2

2:5 3:4 1:1 4:3 9:5 5:21.5

2.0

SiH4:NH

3 gas ratio

SiN

x r

efr

active

in

de

x (

n)

250 350 450 550

Temperature (°C)

600 700 800 900 1000

n

k

Pressure (mTorr)

10-4

10-3

10-2

10-1

100

k a

t 3

60

nm

Fig 4.8: The effect of the PECVD SiNx deposition conditions on the ONO dielectric stack with the measured effective lifetime τeff_10

15, extracted effective charge density with C-V and measured SiNx refractive index n632

with different (a) SiH4:NH3 gas ratios, (b) deposition temperatures and (c) chamber pressures. All lifetime measurements were performed on n-type 100 Ω.cm 100 silicon wafers.

Silane to ammonia gas flow ratio - The measured lifetimes and J0s of the SiH4:NH3 ratio

shows two trends in Figure 4.8a. First, the surface passivation deteriorates with an increased

SiH4:NH3 gas ratio in both thermal SiO2 conditions. Second, the N2 ramp down during thermal

oxidation has a slightly better J0s but poorer τeff_1015 as compared to the ramped down O2. To

further investigate this observation, the effective charge (Qeff) of the ONO dielectric was

Page 63: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

50

determined using C-V measurements on low resistivity 1 Ω.cm n-type 100 silicon wafers.

The C-V measurements on the ONO stack with and without the N2-annealed thermal SiO2 as

shown in Figure 4.8a exhibit two trends. As the SiH4:NH3 ratio increased, the effective charge

density of the ONO stack with the N2-annealed thermal SiO2 decreased to as low as 6.9 × 109

cm-2. Secondly, higher Qeff were measured for SiNx n at ~1.94 (SiH4:NH3 gas ratio 1:1) and

above on thermal SiO2 without any N2 annealing.

The Qeff measured on the N2-annealed wafers declined steadily with the increasing SiNx n

above 1.91 as compared to O2 ramped down wafers having no obvious trend with the SiNx n.

Differences in both the annealing conditions may be due to reducing positive charges within

the thermal SiO2 when subjected to N2 annealing [30, 53, 57]. However, the charges measured

within the stack, which affect the field-effect passivation, showed no clear correlation with J0s.

The slightly lower lifetimes observed with the N2 ramp down wafers could be due to the

introduction of bulk defects with the N2 annealing, as similarly observed in Section 4.3.2.

The annealing conditions (O2 and N2 ramped down) for both wafers further proceeded with the

deposition of positive corona charging and FGA. The results shown in Figure 4.9 show

improvements in the surface passivation for a SiNx n of ~1.94 (SiH4 and NH3 gas ratio of 1:1)

and above. However, either slight degradations or no improvement was observed with wafers

below a SiNx refractive index of 1.94.

1.8 1.9 2.0 2.1 2.2 2.31

10

Orange: Before CC

Red After CC and FGA

SiNx refractive index (n)

τ eff

(ms)

0

5

10

15

20

J

0S

(fA

/cm

2)

1.8 1.9 2.0 2.1 2.2 2.3100

101

(b)

τeff_1015

J0s

τeff_1015

J0s

SiNx refractive index (n)

τ eff

(ms)

(a)

0

5

10

15

20

J0S

(fA

/cm

2)

Purple: Before CC

Blue: After CC and FGA

Fig 4.9: The different thermal oxide conditions with (a) O2 ramp down and (b) N2 ramp down on various SiNx refractive indexes with a measured lifetime at ∆n = 1015 cm-3 and surface recombination J0s.

Page 64: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

51

Based on the results shown in Figure 4.9, the lifetime improves more significantly for the O2

ramp down compared with the N2 ramp down. This further supports the hypothesis that

subjecting wafers to N2 annealing introduces bulk defects. The relatively constant J0s observed

with SiNx refractive indexes n632 of 1.94 to 2.21 for the O2 ramp down wafers may suggest

improvements in the bulk with increases in the SiNx refractive index while within this range.

This was not observed with the N2 annealed wafers as the bulk defects may have limited the

bulk lifetime.

Deposition temperature - The measured lifetime and J0s for deposition temperatures from 250

to 550 ˚C shows two trends in Figure 4.8b. First, the surface recombination decreased with the

rising temperature within the range of 250–400 ˚C. Second, an opposing trend is observed

where the surface recombination increased as the deposition temperature exceeded 400 ˚C.

Further investigations show that the measured Qeff on low resistivity 1 Ω.cm n-type 100

silicon wafers are relatively constant when below the deposition temperature of 400 ˚C.

Wafers with similar resistivities and different wafer crystallographic orientations and surface

morphologies were deposited with ONO using SiNx deposition temperatures from 250–550 ˚C.

The plots in Figure 4.10 show similar results were observed with decreased surface

recombination over the range of 250–450 ˚C and increased surface recombination when the

deposition temperature exceeded 450 ˚C. It should be noted that 100 wafer measured J0s <0

fA/cm2 at PECVD SiNx deposition temperatures of 400 ˚C before and after corona charging.

The measured negative J0s might be either due to injection dependent or an underestimation of

τbulk with Kane and Swanson method [184].

A decrease and increase of the surface recombination with temperature were similarly observed

for a single PECVD SiNx passivation layer on silicon [75, 117], suggesting that the deposition

temperature of SiNx strongly influences the surface recombination, either as a stack on top of

the thermal SiO2 or as a single passivation layer. The lowest J0s (highest lifetime) was achieved

when the deposition temperature was within the range of 350-450 ˚C. A decrease in J0s with

the increasing deposition temperatures up to 400 ˚C suggests that either the density of electric

charges in the film increases or that the interface defect concentration decreases with

temperature. The trend is particularly noticeable for 111 and rantex wafers, which have

previously been found to exhibit higher Dit than 100 surfaces [44, 206, 207], even after a

FGA [44].

Page 65: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

52

It is assumed that the trend for the Qeff measured on 1 Ω.cm n-type 100 silicon wafers is the

same for all ONO samples with different resistivities, crystallographic orientations and surface

morphologies. Therefore, the measured lifetime and surface recombination in Figures 4.10a-c

indicate that the decreased J0s with deposition temperatures up to 450 ˚C arises from a decrease

in the interface defect density, rather than an increase in the charge density within the ONO

stack. Although the mechanism for which the SiNx improves the surface passivation is not

unequivocal, the surface passivation from the SiO2, which improves after capping it with SiNx,

is most likely due to the hydrogenation and positive charges [91, 208, 209]. Another possibility

is that the hydrogen molecules and radicals generated during the deposition [86, 87] passivate

the dangling bonds at the Si-SiO2 interface [90-94], causing a reduction in the J0s. The literature

indicates that the common optimum temperature for hydrogenation using a hydrogen plasma

is 400 ˚C [93, 94], which agrees with our observations. Another source of hydrogenation could

be the deposition of SiNx followed by subsequent annealing, which may release hydrogen into

the bulk or Si-SiO2 interface [210-213].

To quantify the amount of hydrogen within the PECVD SiNx film, FTIR measurements were

performed on the PECVD SiNx layers to determine the concentrations of Si-H and N-H bonds

at 2150 and 3350 cm-1. Figure 4.10d plots the absorption rate over the PECVD SiNx thickness

[214] at different deposition temperatures rather than the absolute bond concentration values

[215]. It is thought that as the bond densities of Si-H ([Si-H]) and N-H ([N-H]) reduce with the

deposition temperature, the released hydrogen could passivate the silicon surface and result in

a lower interface defect density (the Dit could not be determined from the low and high-

frequency C-V measurements due to significant measurement uncertainty in the low-frequency

measurement). However, no clear dependence of J0s on the [Si-H] and [N-H] were found. These

findings are similar to those from Wan et al. for the PECVD SiNx passivation layer [117].

After the corona charging and annealing, the J0s decreased to below 20 fA/cm2 for all deposition

temperatures and wafer types. In fact, for planar 100 wafers, the J0s decreased to practically

0 fA/cm2 for temperature depositions across 300 to 400 ˚C. The J0s decreased the most after

corona charging for deposition temperatures below 400 ˚C. Little or no improvements in the

J0s were observed with deposition temperatures above 450 ˚C. The cause for the increased J0s

for deposition temperatures above 450 ˚C remains uncertain. However, it is possible that

dehydrogenation [216] or contamination via the rapid diffusion of contaminants could have

occurred [197].

Page 66: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

53

250 300 350 400 450 500 550100

101

Temperature °C

(a)

0

4

8

12

16

20

Te

ff (

ms)

J0

s (

fA/c

m2)

250 300 350 400 450 500 550100

101

τeff_10

15 before CC FGA

τeff_10

15 after CC FGA

J0s

before CC FGA

J0s

after CC FGA

Temperature °C

0

10

20

30

40

50

60

70

80(b)

Te

ff (

ms)

J0

s (

fA/c

m2)

250 300 350 400 450 500 55010-1

100

101

Temperature °C

(c)

0

40

80

120

160

200

Te

ff (

ms)

J0

s (

fA/c

m2)

250 300 350 400 450 500 5501011

1012

(d) Planar 100O

NO

Qe

ff (

cm

-2)

CV Qeff

FTIR [Si-H] FTIR [N-H]

Temperature (°C)

104

105

FT

IR S

iNx:

[Si-

H]

| [N

-H]

Bo

nd

s (

%)

Fig 4.10: Measured effective lifetime τeff_1015 and surface recombination prefactor J0s for ONO stacks with

different SiNx deposition temperatures on (a) 100, (b) 111 and (c) rantex wafers. The τeff_1015 is measured at

∆n = 1015 cm-3 before and after corona charging and FGA. (d) Qeff of ONO stack and bond density in SiNx is plotted. The charge was measured using C-V measurement on a 1 Ω.cm n-type 100 silicon wafer, as outlined in Section 3.2. The absorption for the [Si-H] and [N-H] on SiNx was measured using FTIR.

Impact of crystal orientation on ONO passivation – The surface passivation for various

films has been shown to depend on the crystal orientation and surface morphology of the silicon

wafers [44, 122, 217]. For example, Baker-Finch et al. found that the effective upper limit of

surface recombination velocity (Seff.UL) at carrier density (∆n) of 1015 cm-3 on thermal SiO2 for

planar 111 to be higher than that for planar 100 with a ratio fO ≈ 4 [217]. The fO here

represents the surface recombination velocity (SRV) ratio of the surface passivation between

111 to 100 where the subscript O represents the orientation. Wan et al. found the surface

Page 67: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

54

orientation to be less influential for PECVD SiNx with an fO ranging from 0.8 to 1.7 at ∆n =

1015 cm-3 [122]. The surface recombination on a rantex surface (which has 111 oriented

facets) is typically greater than for the planar 111 surface due to its higher surface area (√3

times larger) and possible stress-induced defects at concave and convex features [217, 218]. A

comparison of the surface recombination ratio between the corrected surface area of the rantex

to planar 111 is denoted as fV (where the subscript V is for vertices). Accounting for the

differences in the surface area, Wan et al. found that the SiNx passivation (n = 1.93) has a

higher Seff.UL for rantex over planar 111 with fV ≈ 2 [122]. In contrast, amorphous silicon has

fV ≈ 0.9 [122, 219].

In this work, the fO and fV were compared using the Seff.UL at ∆n = 1015 cm-3 instead of the J0s.

Two trends were observed with calculated fO for the lifetimes plotted in Figures 4.10a-b. First,

the calculated fO was observed to gradually decrease with the increasing SiNx deposition

temperatures up to 350 ˚C. Second, before the corona charging, the ONO stack with SiNx

deposited at 350 ˚C has an fO of ~2.6. After the corona charging and FGA of the ONO stack,

the fO reduced to ~1.2, making it behave more like PECVD SiNx than SiO2. Thus, whether the

crystal orientation was 100 or 111, the influence of surface defects on the ONO passivation

is comparable to PECVD SiNx [122, 219].

Similar trends of fO were observed for the calculated fV with the lifetimes plotted in Figures

4.10b-c. The calculated fV gradually decreased with increasing SiNx deposition temperatures

up to 350 ˚C. The area-corrected fV with the deposited SiNx at 350 ˚C before and after corona

charging showed a similar ratio of ~0.9. This is sufficiently close to unity to indicate that the

poorer ONO passivation on the rantex wafers compared to the planar 111 wafers is due to

their higher surface area, with a minimal impact from the convex and concave aspects of the

morphology.

Chamber pressure - For a direct PECVD configuration, a higher pressure reduces the mean

free paths for all gas species within the plasma chamber, decreases the electron temperature

and reduces the dissociation energies [220, 221]. This leads to a increase in the SiH4:NH3

radical ratio and results in a slightly higher refractive index with higher pressures [222], as

shown in Figure 4.8c.

The J0s shown in Figure 4.8c was found to slightly increase with pressures above 640 mTorr,

and the measure Qeff on the 1 Ω.cm n-type 100 silicon wafer were found to be relatively

similar (>1012 cm-2) across deposition pressures from 600 to 940 mTorr.

Page 68: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

55

The observed increase in the surface recombination with increased pressure is postulated to be

due to a reduced dissociation of atomic hydrogen during the SiNx deposition. Interestingly,

Wan et al. observed a decrease in the single layer PECVD SiNx surface recombination with an

increasing chamber pressure on a dual-mode PECVD system from 0.02–0.4 mTorr [77]. An

increased chamber pressure on the dual-model PECVD system causes a decrease in the SiNx

refractive index. The findings from Wan et al. and our results suggest that changes in the

deposition pressure affect the surface passivation and that changes in the refractive index from

only varying the pressure may indicate the quality of the surface passivation.

No improvements in the J0s with the additional corona charging and annealing steps were

observed, as illustrated in Figure 4.11. This suggests that differences in the surface

recombination with chamber pressure are better associated with improvements at the Si-SiO2

interface as compared to the additional field-effect passivation with corona charging. The lower

lifetimes sometimes observed after corona charging and annealing are due to the

dehydrogenation at the Si-SiO2 interface or bulk, and the reduction of charges after annealing

[167, 178].

600 700 800 900100

101

τeff_1015 before CC FGA

τeff.max after CC FGA

J0s before CC FGA

J0s after CC FGA

Pressure (mTorr)

τ eff

(ms)

0

1

2

J

0s (f

A/c

m2)

Fig 4.11: Measured effective lifetimes and J0s on the ONO stacks with different chamber pressures for PECVD SiNx.

Page 69: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

56

4.4.3 Deposition conditions for PECVD SiOx in the ONO stack

Like PECVD SiNx, the important parameters during the deposition of the PECVD SiOx are the

RF incident power, gas flow ratio, chamber pressure and temperature affect both the optical

properties and deposition rate of the PECVD SiOx [117, 194, 205].

In the following sections, the effects of the SiH4:N2O gas ratio and deposition temperature on

the electrical characteristics of the deposited PECVD SiOx will be discussed. A similar RF

power was used for each experiment and was set to adequately ignite the plasma while reducing

the possibility of plasma damage and ion bombardment [205]. Only one parameter was varied

for each experiment while the other deposition conditions were held constant.

Experimental details

The experimental details for the deposition of PECVD SiOx are divided into two subsections.

Silane to nitrous oxide gas flow ratio - To investigate the influence of the SiH4:N2O gas ratio

on the SiOx, a 100 Ω.cm n-type FZ Silicon wafer with a thickness of approximately 400 µm

was split into quarters. The growth and deposition of the ONO stack for all quarters were

identical to the baseline, except the SiH4 flow varied from 4 and 40 sccm and the N2O was

between 679 and 715 sccm, as shown in Figure 4.4a. The deposition time was varied to keep

the SiOx thickness similar at different gas flow ratios.

Deposition temperature - To investigate the influence of the SiOx deposition temperature, a

100 Ω.cm n-type FZ silicon wafer with a thickness of approximately 400 µm was split into

quarters. The growth and deposition of the ONO stack on all quarters were identical to the

baseline, except that the temperature varied between 200 and 400 ˚C, as shown in Figure 4.4b.

The deposition time was varied to maintain the SiOx thickness for different deposition

temperatures.

Results and discussion

The results for each of the conditions are discussed in the following subsections. Figure 4.12

plots the dependence of the surface recombination, charge density and optical properties for

PECVD SiOx on the various deposition parameters.

Page 70: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

57

1010

1011

1012

100

101

τeff_1015

J0s

τ eff

(ms)

0

1

2

J0s (f

A/c

m2)

(b)

ON

O Q

eff (

cm

-2)

(a)

1:17 1:35 1:79 1:1791.0

1.5

2.0

SiH4:N

2O gas ratio

Re

fra

ctive

Ind

ex a

t 63

2 n

m (

n)

200 250 300 350 400

n k

SiOx Deposition Temperature

10-4

10-3

10-2

10-1

100

k a

t 3

60

nm

Fig 4.12: The effects of the PECVD SiOx deposition conditions on the ONO dielectric stack with measured effective lifetime τeff_10

15, the extracted Qeff determined using C-V measurements and the measured refractive index n with different (a) SiH4:N2O gas ratios and (b) deposition temperatures.

Silane to nitrous oxide gas flow ratio – The J0s and refractive index reduced with decreasing

SiH4:N2O gas ratio, as shown in Figure 4.12a. An initial explanation could be the decreased

effective charges that were measured on the ONO stack with increasing SiH4:N2O ratios.

However, this argument does not explain the results for a ratio of 1:179; the lowest J0s was

presented with measured charges that were similar to SiH4:N2O ratio of 1:35.

Further investigations were carried out on these wafers by corona charging and FGA, as shown

in Figure 4.13a. The results in Figure 4.13a show that a similar surface recombination for J0s

Page 71: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

58

was achieved; however, the lifetimes of the wafers were lower compared with the initial

deposition conditions. This suggests that improvements to the interface with respect to the

SiH4:N2O ratio are more associated with reductions in the Dit of the Si-SiO2 interface rather

than the field-effect passivation from corona charges. The observed trends are postulated to be

caused by two factors. First is the stress induced from the SiOx deposited on top of the ON

stack, as this deposition may advance the release of hydrogen from SiNx to the passivation

dangling bonds at the Si-SiO2 interface or in the bulk. Second is the increasing density of SiOx,

as a denser SiOx capping layer may reduce the out-diffusion of hydrogen from the SiNx.

Deposition temperature – A slight improvement in the surface passivation was observed with

an increasing deposition temperature from 200 to 250 ˚C. However, no further improvements

were observed for deposition temperatures over 250 ˚C. Furthermore, there was no observed

trend with measured Qeff at different SiOx deposition temperatures on the ONO stack.

1.44 1.48 1.52 1.5610

0

101

(a)

J0s

τeff_1015

SiOx refractive index

τ eff (

ms)

Light: Initial

Dark: After CC FGA

1.44 1.48 1.52 1.56

0.0

0.2

0.4

0.6

0.8

1.0(b)

J0s

(fA

/cm

2)

200 300 40010

0

101

SiOx Dep Temperature (°C)

τ eff

(ms)

0

1

2

J

0S

(fA

/cm

2)

Fig 4.13: Measured effective lifetimes and J0s on the ONO stacks with different (a) refractive index n632 and (b) deposition temperatures on the ONO SiOx

The results of the corona charging followed by FGA on the wafers in Figure 4.13b shows an

overall marginal improvement in the surface passivation while exhibiting lower lifetimes.

Improvements in lifetimes observed from 200 to 250 C may be caused by stresses or a reduced

out-diffusion of hydrogen from the SiNx, as discussed in the previous section. No changes to

the surface recombination were observed at deposition temperatures greater than 250 ˚C. This

suggests a saturation point in which additional stresses or denser SiOx may not further improve

the ONO surface passivation.

Page 72: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

59

4.5 Impact of corona charge and annealing on the ONO stack

As previously discussed in Chapter 2, corona charging for dielectric-coated silicon provides

field-effect passivation that can improve surface passivation either by accumulating or

inverting the concentration of free carriers at the surface [65, 166, 167, 173, 174].

In this section, we investigate the injection dependence of effective lifetime of corona charged

ONO stack, on a rantex surface, for various durations between 0 to 180 seconds. Furthermore,

we monitor the charge storage of ONO stack by examining the effective lifetime τeff_1015 of

ONO stack after corona charged and annealed over an extended period of two years.

Experimental details

To investigate the influence of corona charging, n-type 100 Ω.cm silicon wafer was textured

using a TMAH and IPA solution mixture which resulted in a thickness of approximately 350

µm. The growth and deposition of the ONO stack was identical to the baseline. The corona

charging was performed accumulatively with a symmetrical deposition of positive charges at

5 kV on both surfaces.

The impact of annealing after corona charging was measured over an extended period of two

years with fabricated process wafers in Section 4.4.2.

Results and discussion

Figure 4.14a shows a family of lifetime curves for an ONO rantex wafer subjected to different

positive charging durations. Each successive corona charge led to a significant improvement

in the effective lifetime at all excess carrier densities. However, the lifetime saturates with

corona charging beyond 120 seconds.

For comparison, Figure 4.14a shows the conversion from the effective lifetime to an upper

limit surface recombination velocity (Seff.UL) using Eq 3.18 and assuming an infinite bulk

lifetime. An extremely low Seff.UL of <1 cm/s was attained from a rantex silicon surface after

corona charging on the ONO stack for just 120 seconds.

It is preferred to quantify the surface recombination with J0s rather than using τeff and Seff.UL

due to its independence from the injection level when the surface is under accumulation [183].

Page 73: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

60

Figure 4.14b plots the measured J0s (between 1014–1016 cm-3) after successive corona charging

of up to 180 seconds, where no further improvements were observed for τeff across carrier

injection.

There are two notable outcomes in Figure 4.14b. The first relates to the ‘linear’ fit for each

measurement, which indicates the surfaces are in accumulation, as would be expected from a

positively charged dielectric on an n-type silicon wafer [183]. This demonstrates that the

deposited surface charges (Qeff) that cause the accumulation of majority carriers at the Si-SiO2

interface (ns) are large enough to accurately extract J0s.

1013

1014

1015

1016

0

5

10

15

20

25

0 2x1015

4x1015

6x1015

0

50

100

150

200

250

300

180s

120s

60s

20s

10s

0s

fit τeff

τe

ff (

ms)

Excess carrier density (cm-3)

ONO Rantex

--

3.5

1.8

1.2

0.9

0.7

Up

pe

r lim

it S

RV

(cm

/s)

5.4 fA/cm2

44 fA/cm2

37 fA/cm2

21 fA/cm2

11 fA/cm2

5.9 fA/cm2

1/τ

eff -

1/τ

bu

lk (

intr

insi

c)(s

-1)

Excess carrier density (cm-3)

0s

10s

20s

60s

120s

180s

best fit J0s

(a) (b)

Fig 4.14: (a) ONO stack on a 100 Ω.cm n-type rantex silicon wafer that was corona charged for an increasingly cumulative period at 5 kV; the calculated upper limit SRV on the secondary y-axis scale is not to scale. (b) Line plots showing the best-fit

Secondly, the results clearly show a substantial reduction in J0s from 44 to ~5 fA/cm2 after

successive corona charging for up to 180 seconds. The additional corona charging caused an

increased ns, which reduced the recombination of minority carriers at the surface.

The deposition of the corona charging is not stable without further treatment. The deposited

charges can be removed by rinsing the wafers in isopropanol (IPA) [65, 167, 223]. The benefit

of a simple corona charge is, therefore, only temporary and would not provide a long-term

efficiency gain in finished devices. However, annealing corona-charged wafers have been

Page 74: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

61

found to embed the surface charges into the dielectric [65, 167, 176]. To examine the stability

of the embedded charges into the ONO stack, the annealed ONO wafers from Section 4.4.2

(100, 111 and rantex wafers) were monitored over a two-year period under ambient

conditions. Figure 4.15 shows the results of the wafers before and after the corona charging,

illustrating the annealed corona-charged wafers after 400 C in FGA. From the measured values

of τeff_1015 and J0s shown in Figure 4.15, the corona-charged wafers were observed to have a

slight increase in the surface recombination after annealing due to their reduced overall charges

or stresses caused by further annealing of the stack [65, 167]. It is evident that the passivation

scheme is stable over a two-year period (stored indoors) and is not dependent on the surface

condition of each wafer.

0 200 400 600 8000

5

10

15

20

25

30 Before CC

After CC

Annealed CC

Before and

after CC

J0s 111

J0s Rantex

τeff_1015 Rantex

τeff_1015 111

τeff_1015 100

Days

τ eff (

ms)

0

5

10

15

20

J0s

(fA

/cm

2)

Fig 4.15: Effective lifetime measured at ∆n = 1015 cm-3 and J0s for the corona-charged wafers plotted after FGA annealing and stored in ambient conditions for up to two years. No data plots are displayed for the J0s of the 100 wafer, as the extracted values were lower than 0 fA/cm2. The wafer thicknesses of the 100 planar, 111 planar and rantex surfaces were measured to be approximately 400, 420 and 380 µm, respectively.

Kelvin probe measurements were performed after the corona charging showed positive surface

charges of 4.3 × 1012 cm-2 for an n-type planar 1 Ω.cm silicon wafer. The C-V measurements

performed before and after the corona charging (with annealing to embed the charges), showed

an increase in the effective charge Qeff from approximately 7 ± 2 × 1011 to 2.4 ± 0.3 × 1012 cm-

2. The measured increase in the effective charge indicates that the deposited surface corona

charges are driven into the ONO dielectric stack after subsequent annealing.

Page 75: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

62

4.6 ONO surface passivation on various wafer resistivities and surfaces

A summary of the principal results obtained from this study on the ONO passivation scheme

is compiled in Table 4.1. The table includes details of the wafer types, thicknesses and surface

morphologies. The conditions for thermal SiO2, PECVD SiNx and SiOx were compiled to

establish an ONO stack with optimised surface passivation. The wafers featuring the ONO

surface passivation were processed as follows:

(i) Ramp up from 700 to 1000 ˚C and the subsequent ramp down to 700 ˚C, all in O2, with

no subsequent POA in N2 (12 ± 2nm).

(ii) 400 C PECVD SiNx (50 ± 3 nm) with a SiH4:NH3 gas ratio of 1:1 and chamber pressure

of 650 mTorr.

(iii) 250 ˚C PECVD SiOx (90 ± 5 nm) with a SiH4:N2O gas ratio of 1:79.

(iv) Anneal the ONO stack in FG at 400 ˚C for 30 minutes.

(v) Deposit positive corona charge at 5 kV onto the ONO passivated wafers for 120 seconds.

(vi) Anneal the charged ONO wafers in forming gas at 400 ˚C for 30 mins to embed and

stabilise the corona charges.

1013

1014

1015

1016

1017

10-5

10-4

10-3

10-2

10-1

100

1013

1014

1015

1016

10-3

10-2

10-1

100 (b)

Richter et al.

Trupke et al.

Dingermans et al.

Hacker et al.

Kerr et al.

Veith-Wolf et al.

Wan et al.

Grant et al.

Niewelt et al.

Steinhauser et al.

Bonilla et al.

Kho et al. (this work)

τ eff (s)

Dopant Concentration (cm-3)

(a)

86.7 Ω.cm

1.77 Ω.cm

1.07 Ω.cm

τ eff (

s)

Excess Carrier Density (cm-3)

0.50 Ω.cm

1.07 Ω.cm

1.50 Ω.cm

86.7 Ω.cm

Intrinsic limit

0.50 Ω.cm

n-type

Fig 4.16: (a) Maximum lifetime as a function of the dopant concentration measured at 300 K on an n-type silicon wafer together with data from the literature ([17, 64, 140, 224-230]). (b) ONO passivation with a range of n-type silicon resistivities compared to the intrinsic limits with a parameterised Auger [17] and radiative recombination [181] (lines).

Page 76: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

63

Figure 4.16a plots the maximum effective lifetime measurements published in the literature on

n-type silicon wafers. This comparison demonstrates that the lifetime of samples with the ONO

passivation exceeds the lifetimes for the various other passivation schemes as it achieved a

lifetime of 170 ms on a 100 Ω n-type silicon wafer. Figure 4.16b shows the effective lifetime

measurements for three different n-type resistivity wafers with ONO passivation, which

exceeds the intrinsic recombination lifetime model [17]. This not only demonstrates the

outstanding passivation quality of ONO stacks, but it also highlights the necessity of revising

the parameterised intrinsic limits for silicon.

Table 4.1. Measured lifetime parameters for a variety of n-type wafers with ONO, NO and SiNx passivations.

Note: J0s is extracted by fitting injection levels from 1 to 5 x 1015. Values are reported at 300 K, where ni is 9.7 × 109 cm-3, τeff_10

15 is at ∆n = 1015 cm-3, and τeff_max is at ∆n from 1013 to 1016 cm-3. Chem. P: Chemical Polish Rantex: Random texture a,b SiNx Deposition technique: PECVD (Roth & Rau AG, system AK400) b SiNx deposited onto silicon wafer followed by PECVD SiOx [14]. *Wafer resistivity measured at the University of Oxford using the Hall Coefficient measurements after thermal processing [231]. *Wafer was not subjected to step (iv) in Section 4.6 All other wafer reported resistivities were based on dark conductance measurements using the QSSPC before any thermal processing.

ρbulk

(Ω·cm)

Type Crystal

Lattice

w

(µm)

Surface Passivation

Layer

τeff_1015

(ms)

τeff_max

(ms)

J0s

(fA/cm2)

0.50 CZ 100 155 Chem. P ONO 3.2 3.7 -

1.07 CZ 100 272 Chem. P ONO 12.4 15.1 -

1.77* FZ 100 398 Chem. P ONO 19.3 25.5 -

2.71 FZ 100 228 Chem. P ONO 18.6 20.6 -

4.97* FZ 100 388 Chem. P ONO 34.9 43.7 0.2

8.93* FZ 100 181 Chem. P ONO 28.0 38.5 0.8

86.7* FZ 100 291 Chem. P ONO 72.5 170 0.3

142 FZ 100 382 Chem. P ONO 44.9 84.6 -

115 FZ - 350 Rantex ONO 18.0 42.6 4.3

105 FZ - 351 Rantex a SiNx 10.1 12.9 9.8

2.96 FZ - 250 Rantex bNO 4.3 4.5 4.7

Page 77: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

64

4.7 Chapter summary

This chapter presents a detailed study for the surface passivation of an ONO dielectric stack on

undiffused n-type silicon wafers. The study is divided into four main areas: thermal SiO2,

PECVD SiNx, PECVD SiOx and corona charging with subsequent annealing.

The initial thermal SiO2 grown on the silicon surface provides good chemical surface

passivation. A thermal SiO2 thickness of 7–15 nm was sufficient to provide excellent surface

passivation, while the introduction of POA with N2 was deemed unnecessary. The POA with

N2 was found to reduce positive charges within the thermal SiO2 and contributed to bulk defects

or contamination.

The second layer of the deposited SiNx improves surface passivation by hydrogenation the Si-

SiO2 interface and induces a field effect passivation with positive charges. The deposition

conditions for the investigated PECVD SiNx include the deposition temperature, SiH4:NH3 gas

flow ratio and the chamber pressure. It was found that the optimum deposition temperature was

from 350–450 ˚C on planar orientated 100 and 111 and rantex wafers. The increase in the

SiH4:NH3 gas ratio was observed to have fewer positive charges in the SiNx. Also, the lowest

surface recombination was observed at a chamber pressure of 650 mTorr.

The third layer capping PECVD SiOx indicates that it traps and embeds the positive corona

charges while improving the overall ARC properties of the ONO stack in air, which will be

discussed further in Chapter 6. The deposition conditions for the investigated PECVD SiOx

includes the SiH4:N2O gas flow ratio and the deposition temperature. It was revealed that the

SiH4:N2O gas ratio of 1:179 and deposition temperatures above 250 ˚C provided the lowest

surface recombination.

The corona charges slowly dissipated in air on the untreated dielectric surfaces. However, by

annealing the corona charged ONO surfaces at 400 ˚C in a forming gas, the charges were

deemed stable for up to two years under indoors conditions. For most occasions, the deposition

of corona charges and further annealing improved the surface passivation and embedded the

charges into the dielectric stack. However, degradation of the surface from corona charging

may occur due to ion bombardment at the Si-SiO2 interface. Therefore, the amount of corona

charging and subsequent annealing should be optimised to avoid damaging the Si-SiO2

interface.

Page 78: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

65

Excellent surface passivation and record-breaking lifetimes for silicon wafers featuring ONO

passivation were demonstrated, proving that this passivation scheme is amongst the best for

dielectric films. Surface passivation on planar surface of 100 and 111 wafer orientation

achieved J0s ≤1 fA/cm2 and random texture surface J0s <5 fA/cm2. As such, these could find

applications in high-efficiency silicon solar cells.

Page 79: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

66

Chapter 5

Passivation of ONO on diffused surfaces

5.1 Introduction

This chapter focuses on using the ONO stack to achieve excellent surface passivation for

silicon substrates with diffused surfaces by optimising their electrical properties. The primary

use of diffusion is to form a p-n junction in semiconductor devices which enables the specified

carriers to be separated and ‘collected’ at the junction. Other uses of diffusion include reducing

the surface recombination via field-effect passivation, lowering the contact resistivity between

metal and silicon by reducing the barrier width through field emissions and improving the

lateral flow of carriers within a solar cell [232]. This chapter investigates ONO passivation on

boron and phosphorus-diffused surfaces, which is an essential study to achieve high-efficiency

silicon solar cells.

Based on the experimental results presented in Chapter 4, an ONO stack with select processing

conditions was implemented on boron and phosphorus-diffused surfaces. Firstly, the basic

wafer preparation procedure, such as saw damage removal, RCA cleaning, diffusion processes,

etc., is briefly described. Secondly, the impacts of the thermal SiO2 layer (thermal SiO2

thickness and post-oxidation annealing), PECVD SiNx film deposition conditions and post-

treatments on the surface passivation of diffused silicon substrates were studied and evaluated

and are presented in the following sections.

5.2 ONO baseline procedures on diffused surfaces

High resistivity n-type 100 Ω.cm silicon wafers with a 100 orientation were either etched in

TMAH for 10 minutes at 85 ˚C to remove saw damage or chemically textured to form rantex

surface using solution mixture of TMAH and isopropanol (IPA). Planar and rantex wafers after

the etching had thicknesses of ~400 µm and ~350 µm respectively. The wafers were

subsequently cleaned using the standard RCA process and dipped in a 1% hydrofluoric (HF)

solution until the surfaces became hydrophobic.

Page 80: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

67

Thermal diffusion was performed in a Tempress 5-inch quartz tube system with wafers loaded

in O2 at 700 ˚C. For phosphorus diffusion, wafers were ramped to 775 ˚C and diffused for 25

minutes with gas flow rates of 130 sccm for the N2 carrier gas with phosphoryl chloride (POCl),

65 sccm for O2 and 2.5 slm for N2 diluent gas. After diffusion, the wafers were ramped down

to 700 °C and unloaded in O2. For the boron diffusion, the wafers were ramped up to 920 ˚C

and diffused for 40 minutes with flow rates of 18 sccm of the N2 carrier gas with boron

tribromide (BBr3), 18 sccm for O2 and 2.5 slm for N2 diluent gas After diffusion, the wafers

were ramped down to 700 °C in O2 and unloaded. The diffused wafers with phosphosilicate

glass (PSG) and borosilicate glass (BSG) layers were removed in 5% HF solution, cleaned

using the standard RCA process and dipped in a 1% HF solution to prepare for the thermal

oxidation.

The growth of thermal SiO2 was similar to those in the baseline procedure in Section 4.2. The

thermal SiO2 thickness was measured using an ellipsometer on the 100 orientated planar

surface with a light phosphorus diffusion of 400–500 Ω/ was 15 ± 1 nm, and was 17 ± 2 nm

for a boron diffusion of ~160 Ω/ (JA Woollam M2000D) [62]. The rantex wafer was assumed

to be similar to the measured phosphorus-diffused 111 wafer with a thickness of 16 ± 1 nm.

The PECVD SiNx and SiOx layer deposited on top of the thermal SiO2 via PECVD (using an

Oxford Instruments PlasmaLab 100) was identical to the procedure given in Section 4.2.

Following the PECVD process, all wafers were subsequently FGA at 400 ˚C in a quartz tube

furnace for 30 minutes.

The diffused wafers with symmetric ONO stacks were corona charged on both the front and

rear surfaces using the conventional setup [164]. Positive corona charging was performed on

the phosphorus diffused surface at 5.5 kV for 120 seconds, whereas negative corona charging

was performed on boron diffused surfaces at 6 kV for 100 seconds.

5.3 Surface passivation of ONO stack on phosphorus and boron diffusion

The net emitter recombination current density pre-factor (J0e) for diffused silicon was

determined based on the quality of the surface passivation, the surface dopant concentration

and the diffusion profile. This section investigates the impact of the ONO stack on the J0e for

different diffusion profiles.

Page 81: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

68

Experimental details

Phosphorus and boron-doped silicon substrates with both planar and rantex surfaces were

fabricated as described in the baseline procedures of Section 5.2, except the diffusion

temperature for phosphorus and boron) were varied between 770 to 800 ˚C and 890 to 940 ˚C

respectively. The formation of the ONO stack and deposition of the corona charges onto the

wafers were similar to the baseline procedures detailed in Section 5.2.

Results and discussion

The measured J0e for phosphorus and boron diffused wafers with the ONO stack are presented

in Figures 5.1a-b, respectively. Two trends were observed in Figure 5.1a. First, the J0e for the

phosphorus diffused wafers increased with lower sheet resistances. Second, positive corona

charging reduced the recombination of phosphorus diffused wafers with ONO passivation

stacks. Prior to corona charging, the J0e for the passivated ONO stack on rantex surface (grey

filled triangle symbols) was comparable to that for the rantex alnealed thermal SiO2 (grey

hollow triangle symbols). After positive corona charging of phosphorus doped silicon substrate,

the surface passivation improved significantly on the rantex surfaces, whereas only slight

improvements were observed on the ONO planar surfaces. The lowest phosphorus diffused

sheet resistivity of 520 Ω/ on planar and 610 Ω/ on rantex wafers with ONO stacks and

positive corona charges achieved J0e of 0.5 fA/cm2 and 1.9 fA/cm2, respectively.

Large variations in the J0e were found on boron diffused silicon wafers with ONO surface

passivation prior to corona charging, as shown in Figure 5.1b. No obvious trends were observed

with the ONO passivation as the J0e values ranged from 20 to 34 fA/cm2 across sheet

resistivities from 125 to 180 Ω/. The ONO surface passivation quality is comparable to the

thermal SiO2 data from Kerr et al. [233]. As PECVD SiNx generally exhibits positive charges,

which are undesirable with boron diffused surfaces, the boron diffused ONO stack was

negatively corona charged with a -6 kV bias [234, 235]. The negative charges substantially

reduced the ONO surface recombination and were found to be comparable to the surface

passivation of the Al2O3 layer and the a-Si/SiNx stack, both of which are typically regarded as

among the best passivations available for boron diffused surfaces [236, 237].

Page 82: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

69

0 100 200 300 400 500 600

100

101

102

0 50 100 150 200 250100

101

102

Rantex Alneal - M.J. Kerr

Rantex ONO

Pos. CC Rantex ONO

Planar ON - M.J. Kerr

Planar ONO

Pos. CC Planar ONO

J0e (

fA/c

m2)

Sheet Resistivity (Ω/ )

(a)

SiO2 - M.J. Kerr

aSi/SiNx - M. Kessler

Al2O

3 - J. Schmidt

ONO

Neg CC. ONO

J0e (

fA/c

m2)

Sheet Resistivity (Ω/ )

(b)

Fig 5.1: Compilation data for different types of passivation schemes comparing the J0e of (a) planar and rantex phosphorus diffused [194, 233] and (b) planar boron diffused samples [236, 237].

The J0e for the ONO passivated phosphorus and boron-doped silicon wafers increased with

lower sheet resistivities due to an increased Auger recombination within the diffusion profile

[18, 238]. The correct polarity of charges with additional corona charging (positive charges on

phosphorus and negative charges on boron) improved both diffused surfaces as it accumulates

majority carriers from the diffusion towards the surface. The low J0e attained on both diffusions

after corona charging demonstrates that the ONO stacks can provide good chemical passivation.

5.4 Thermal oxide growth conditions in an ONO stack on diffused

surfaces

Varying the thermal SiO2 growth conditions changes the surface passivation quality, surface

dopant concentration and the diffusion profile. During the oxidation process, the dopant

redistributes near the Si-SiO2 interface. The growth of thermal SiO2 tends to deplete the boron

dopants at the surface whereas phosphorus dopants tend to pile up. The post-oxidation

annealing (POA) with N2, however, drives the diffusion dopants deeper into the silicon.

This section investigates the effects of the thermal SiO2 thickness and duration of the POA on

the passivation quality of the ONO stack for both boron and phosphorus diffusions. The

conditions investigated in this section are presented in flowchart Figure 5.2 with the details

described in the following subsections.

Page 83: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

70

Fig 5.2: Flow chart detailing the process variations to each layer (relative to baseline) to establish the best passivation conditions for the (a) thermal SiO2 thickness and (b) POA.

5.4.1 The effect of varying the thermal oxide thickness on the diffused surfaces

The difference in the segregation coefficient of silicon and thermal SiO2 for diffused boron or

phosphorus atoms causes the dopant concentration to either deplete or accumulate on the

silicon surface.

As diffusion introduces additional complexities compared with undiffused wafers, this section

examines the effect of varying the thermal SiO2 thickness on the electrical performance of the

ONO stack for similar diffused wafers.

(a) Thermal SiO2 thickness (b) Post-oxidation N2 annealing

Phosphorus Diffusion 780 ˚C

Thermal

Oxidation 850 °C

PECVD SiNx 400 °C

PECVD SiOx 250 °C

FGA 400 °C Boron Diffused

Wafers

Thermal

Oxidation 900 °C

Thermal

Oxidation 950 °C

Thermal

Oxidation 1000 °C

N2 annealing 1000 °C

Planar 100 and rantex

quartered

Boron Diffusion

940 ˚C

Planar 100

quartered

FGA 400 °C Phosphorus Diffused

Wafers

Positive Corona

Charging 5 kV Negative Corona

Charging 6 kV

PECVD SiOx 250 °C

Control

Thermal

Oxidation 1000 °C

Ramp down in O2

PECVD SiNx 400 °C

Thermal Oxidation 1000 °C

Anneal and ramp down N2

Phosphorus Diffusion 780 ˚C

Planar 100 and rantex

quartered

Boron Diffusion

940 ˚C

Planar 100

quartered

FGA 400 °C Boron

Diffused Wafers

FGA 400 °C Phosphorus Diffused

Wafers

Positive Corona

Charging 5 kV Negative Corona

Charging 6 kV

Page 84: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

71

Experimental details

To investigate the impact of the thermal SiO2 thickness, both phosphorus and boron diffusion

were performed using a 100 Ω.cm n-type FZ silicon wafer on planar or rantex surfaces. Each

wafer with specific diffusions and surface morphologies were cleaved into quarters. The

diffusion, growth and deposition of the ONO stack on all quarters followed the baseline, except

the duration of the initial thermal oxidation growth was varied. The wafers were loaded in O2

at 700 ˚C and ramped to different temperature set points. The thermal SiO2 growth on the

wafers occurred during the ramp up to 850, 900, 950 and 1000 ˚C. Immediately after reaching

the defined temperature set points, the O2 was switched to N2 and subjected to further annealing

at 1000 °C for 45 minutes in N2. By altering the duration at which the wafers were exposed to

oxidation, the thermal SiO2 thicknesses were measured to be between 4 to 16 nm. The process

flow is as illustrated in Figure 5.1a.

Results and discussion

The thermal SiO2 thicknesses were varied on the phosphorus diffused planar, phosphorus

diffused rantex, and boron diffused planar wafers. It is possible that variations in the oxidation

growth temperatures may affect the surface passivation; however, the different growth

temperatures of the thermal SiO2 underwent POA in N2 at 1000 °C for 45 minutes. The surface

passivations with different thermal SiO2 thicknesses required to achieve excellent surface

passivation are plotted in Figure 5.3.

Figure 5.3a illustrates that a thermal SiO2 thickness of approximately 7 nm or greater provides

excellent surface passivation on a light phosphorus diffused planar surface. However, a thicker

overall thermal SiO2 was required on the rantex surface as observed in Figure 5.3b, which

demonstrates excellent surface passivation with a thermal SiO2 thicknesses of approximately

10 nm or greater. One possible reason for the thicker thermal SiO2 requirements is due to the

concave and convex corners on the rantex, which have slower thermal SiO2 growth rates [239].

The surface passivation on the diffused planar wafers showed no substantial improvements

after positive corona charging; however, notable improvements were observed after corona

charging for rantex wafers with a reduced J0e. This trend was similarly observed on the

undiffused n-type silicon wafer passivated with ONO. This is most likely due to Dit with

thermal SiO2 on the 111 oriented rantex surfaces is generally higher compared to those on

100 oriented wafers [44, 206, 207].

Page 85: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

72

5 10 15100

101

418 Ω/

532 Ω/

506 Ω/

(a)

τeff_10

15 before CC

J0e

before CC

τeff_1015 after positive CC

J0e

after positive CC

Thermal SiO2 thickness (nm)

τeff (

ms)

423 Ω/

101

100

J0e (

fA/c

m2)

5 10 15100

101

545 Ω/

692 Ω/

602 Ω/

532 Ω/

(b)

τeff_1015 before CC

J0e

before CC

τeff_1015 after positive CC

J0e

after positive CC

Thermal SiO2 thickness (nm)

τeff (

ms)

100

101

J0e (

fA/c

m2)

5 10 1510

-1

100

101

τeff_1015 before CC

J0e

before CC

τeff_1015 after negative CC

J0e

after negative CC

Thermal SiO2 thickness (nm)

τeff (

ms) 89 Ω/

105 Ω/ 135 Ω/

145 Ω/

100

101

102

J0e (

fA/c

m2)

(c)

Fig 5.3: Minority carrier lifetimes τeff_1015 and J0e measured for ONO on (a) phosphorus diffused planar, (b)

phosphorus diffused rantex and (c) boron diffused planar with different thermal SiO2 thicknesses annealed at 1000 ˚C in N2.

Figure 5.4c shows that the J0e for boron diffused wafers reduced as the thermal SiO2 thickness

increased. Reductions in the J0e with thicker thermal SiO2 is caused by the decreasing boron

dopant concentration at the surface (lower Auger recombination) and a reduced field effect

from SiNx. Due to the preferential segregation of boron dopants into the thermal SiO2, the

overall dopant concentration and the surface doping concentration are reduced with the

increased thermal SiO2 growth [240]. This effect consequently reduced the Auger

recombination while also increasing the sheet resistance [18, 238]. As the PECVD SiNx

generally exhibits positive charges within the dielectric film [234, 235], the field effect of SiNx

Page 86: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

73

reduced with the increasing SiO2 thickness due to the charge centroid being driven away from

the silicon surface [192, 241]. As negative corona charges are applied to the surface, the J0e for

all boron diffused wafers decreased. This observation demonstrates the importance of the field-

effect passivation in the passivation of boron diffused surfaces from the ONO film stacks.

5.4.2 The effect of post-oxidation annealing of thermal oxide on diffused surfaces

The POA process on diffused wafers drives dopants deeper into the silicon bulk and slowly

diffuse dopants into the thermal SiO2. Higher POA temperatures with N2 may reduce the

overall Auger recombination of the diffusion profile while maintaining a similar electrical

conductivity, e.g. sheet resistance.

As the effects of the POA on undiffused wafers were discussed in Chapter 4, this section

examines the influence of POA on thermal SiO2 on the electrical properties of ONO stack for

diffused wafers.

Experimental details:

The 100 Ω.cm n-type FZ silicon wafer on planar or rantex surfaces were used to investigate

the effects of the POA in N2 for diffused surfaces. Diffusions performed on the wafers were

done similarly to baseline procedure in Section 5.2. Each wafer with specific diffusion and

surface morphologies were cleaved into quarters. The ONO stacks were similar to baseline

procedures in Section 5.2, except for the POA for thermal SiO2. the subsequent POA in N2

annealing was varied between 0, 15 and 30 minutes at 1000 ˚C and cooled down in ambient

N2. A control quarter without POA in N2 annealing was loaded at 700 ˚C, ramped to 1000 ˚C

and then ramped down in ambient O2. The process flow is illustrated in Figure 5.2b.

Results and discussion:

The results of thermal SiO2 subjected to POA in N2 for both diffusions are shown in Figure 5.4.

POA with N2 from Section 4.4.2 on n-type undiffused planar 100 orientation silicon substrate

has shown to reduce positive charges within the thermal SiO2. The phosphorus diffused planar

wafers in Figure 5.4a show a reduced J0e for a prolonged POA. A longer POA in N2 improves

the J0e by reducing the surface Dit and driving in the diffusion dopants deeper into the bulk,

Page 87: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

74

which may reduce the Auger recombination of the diffused region [53]. Further positive corona

charging of the phosphorus diffused planar wafers shows either a marginal improvement or no

changes to the measured J0e. Phosphorus rantex wafers in Figure 5.5b show no obvious trend

with a prolonged N2 annealing before corona charging. However, the surface passivation

improves with subsequent positive corona charging under all annealing conditions. A

consistent J0e of ~4 fA/cm2 was measured across all annealing conditions. This suggests that

the N2 annealing may not improve the surface Dit of Si-SiO2 interface for diffused rantex wafers,

and that charges have a much larger influence on the surface passivation of 111 orientated

wafers as compared to 100 wafers. Both types of phosphorus diffused wafers (planar and

rantex), which were not subjected to any N2 annealing (i.e. when the ramp down is performed

in O2), have higher lifetimes compared to wafers with N2 annealing. This trend was similarly

observed for undiffused n-type wafers in Section 4.3.2. As also discussed in Section 4.3.2, the

reduction in lifetimes with N2 annealing is likely due to the re-introduction of vacancy-like

defects back into the silicon bulk, as demonstrated by Grant et al. [199, 200].

In Figure 5.5c, boron diffused planar wafers share similar trends with those that have

phosphorus diffusion, except when not subjected to any N2 annealing. The surface passivation

improves with N2 annealing due to a reduced surface Dit at the Si-SiO2 interface, dopants

concentration driven deeper into the bulk and reduced positive charges within the thermal SiO2

[53]. In contrast to phosphorus diffused wafers, a high J0e was observed for the boron diffused

surface with no N2 annealing (boron diffused J0e of 230 fA/cm2 as compared to phosphorus

diffused J0e of 4 fA/cm2). The surface passivation quality improved significantly after inducing

negative charges via corona charging, which approached the J0e values for the wafers that

underwent N2 annealing. This demonstrates that the thermal SiO2 that was ramped down in O2

has good chemical passivation and that the measured high J0e was due to positive charges within

the thermal SiO2.

Page 88: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

75

0.0 0.1 0.2 0.3 0.4 0.51018

1019

0 10 20 30O2

100

101

556 Ω/

558 Ω/

556 Ω/

503 Ω/

(a)

τeff_1015 before CC

J0e

before CC

τeff_10

15 after positive CC

J0e

after positive CC

N2 Annealing (minutes)

τeff (

ms)

10-1

100

101

J0e (

fA/c

m2)

0 5 10 15 20 25 30O2

100

101

654 Ω/

682 Ω/710 Ω/

642 Ω/

(b)

τeff_1015 before CC

J0e

before CC

τeff_1015 after positive CC

J0e

after positive CC

N2 Annealing (minutes)

τeff (

ms)

100

101

J0e (

fA/c

m2)

0 10 20 30O2

10-1

100

101

τeff_1015 before CC

J0e

before CC

τeff_1015 after positive CC

J0e

after positive CC

N2 Annealing (minutes)

τeff (

ms)

100

101

102

177 Ω/175 Ω/176 Ω/

(d)

J0e (

fA/c

m2)

(c)

202 Ω/

180 Ω/

O2

0 N2

15 N2

30 N2

Simulation

Bo

ron D

opan

t C

on

centr

ation (

cm

-3)

Depth (µm)

180 Ω/

Fig 5.4: Minority carrier lifetimes τeff_1015 and J0e measured for the ONO stack samples on (a) phosphorus diffused

planar (b) phosphorus diffused rantex and (c) boron diffused planar surfaces with different thermal SiO2 POA duration. (d) Boron diffused profile with different POA duration as measured using ECV together with the simulated rectangle diffusion profile.

Electrochemical capacitance-voltage (ECV) measurements and EDNA2 simulations were

performed on the boron diffused wafers to better understand the contributions of Auger and

surface recombination (J0s) on the total emitter recombination (J0e). EDNA2 is a software

package that models the recombination of an emitter in one dimension [18]. The measured

boron diffusion profiles with ECV are plotted in Figure 5.5d. The results shown in Table 5.1

indicate that the J0s and Auger recombination within the emitter reduced with prolonged POA

in N2. To demonstrate the reduction in Auger recombination, simulated profiles of boron

diffusion with identical sheet resistivities show that Auger recombination reduces as the boron

Page 89: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

76

dopants are redistributed from the high surface dopant concentration with a shallow junction

to a smaller surface dopant concentration with a deeper junction.

Table 5.1: Simulations for different boron diffusion profiles with EDNA2 after negative corona charging on the ONO stack. The simulation assumes no SRH recombination within the diffusion profile.

5.5 Deposition conditions for PECVD SiNx of the ONO stack on diffused

surfaces

The deposition conditions for the SiH4:NH3 gas flow ratio were examined on undiffused wafers

in Section 4.4.2. The results obtained for undiffused wafers indicate that increasing the ratio

between the SiH4:NH3 increases the SiNx refractive index and reduces the Qeff of the ONO

stack of N2 POA thermal SiO2.

In this section, the effects of the ONO stack with different SiH4:NH3 gas ratios on diffused

wafers were investigated, with the investigation process presented in the flowchart of Figure

5.5.

Experimental details

Several FZ n-type 100 Ω.cm chemical polished 100 planar silicon wafers were used to

investigate the influence of the SiNx refractive index. The diffusion and deposition of the ONO

stack on all wafers was identical to the baseline except for the thermal oxidation and PECVD

SiNx conditions. The thermal oxidation for phosphorus was ramped down in O2 whereas the

thermal oxidation for boron was ramped down in N2. The different gasses performed during

the thermal oxidation ramp down were selected to control the quantity of positive charges

within the thermal SiO2, as discussed in Chapter 4. The phosphorus diffused wafers were

POA Duration Sheet Resistivity Nsurface EDNA 2

(min) (Ω/) (1019 cm-3) Zf (µm) J0e

(fA/cm2) JAuger

(fA/cm2) J0s (fA/cm2)

O2 0 202 0.3 - 21.4 6.5 14.9

N2 0 176 2.0 - 16 8.6 7.4

N2 15 175 2.2 - 12 8.5 3.5

N2 30 177 2.0 - 11 7.6 3.4

Simulation - 180 1 0.540 5.3 5.3 0

Simulation - 180 3 0.203 15 15 0

Page 90: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

77

loaded at 700 ˚C, ramped up to 1000 ˚C and ramped down immediately to 700 ˚C in ambient

O2. The boron diffused wafers were loaded in at 700 ˚C and ramped up to 1000 ˚C in O2, and

ramped down in N2 to 700 C. The sheet resistivity of the phosphorus and boron diffused wafers

was measured to be 570 ± 30 Ω/ and 180 ± 10 Ω/ respectively.

In the case of the PECVD SiNx, the SiH4:NH3 gas ratios were varied between 0.4 to 7, as the

SiH4:NH3 gas flows were varied between 4 to 24 sccm. The deposition time was varied to

maintain the SiNx thickness at different gas ratios. The SiNx and SiOx thicknesses of 55 ± 3 nm

and 90 ± 5 nm, respectively, were achieved.

Fig 5.5: Flow chart detailing the process variations for each layer (relative to the baseline) to establish the best passivation conditions of the gas ratio for PECVD SiNx

Gas ratio of SiNx

Phosphorus

Diffusion 780 ˚C

Thermal Oxidation

1000 °C

Ramp down in O2

PECVD Silicon Nitride 400 °C

(Different gas ratio)

PECVD Silicon Oxide 250 °C

FGA 400 °C Boron

Diffused Wafers

Thermal

Oxidation

1000 °C

Ramp down in N2

Planar 100 and

rantex, quartered

Boron Diffusion

940 ˚C

Planar 100

quartered

FGA 400 °C

Phosphorus

Diffused Wafers

Positive Corona

Charging 5 kV

Negative Corona

Charging 6 kV

FGA 400 ˚C FGA 400 ˚C

Page 91: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

78

Results and discussion

The results for J0e with various SiNx refractive indexes n632 before and after the corona charging

for both diffusions are plotted in Figures 5.6a-b. Prior to corona charging, the lowest SiNx

refractive index n632 of 1.83 was found to have the highest J0e with both diffusions. The surface

passivation for phosphorus diffused wafers improved and stabilised across SiNx refractive

indexes n632 of 1.90 to 2.58. The improvements to J0e with higher could be related to

hydrogenation of the surface and bulk as observed on the undiffused wafer in Section 4.4.2.

Fig 5.6: Minority carrier lifetimes and surface recombination pre-factors as measured for ONO stack samples on (a) phosphorus diffused planar and (b) boron diffused planar surfaces with different SiNx refractive indexes n632.

Similarly, the surface passivation for boron diffused wafers improved with an increased SiNx

refractive index n632 and stabilised across the range of 1.94 to 2.58. In Section 4.4.2, C-V

measurements performed on the various SiNx refractive indexes shows that the effective

positive charges on the ONO stack decrease as the refractive index increased from 1.90 to 2.21.

The reduced positive charges across the refractive indexes of 1.90 to 2.21 are reflected in the

decreasing J0e for the boron diffused wafers plotted in Figure 5.6b. Negative corona charging

on boron diffused wafers significantly improved the surface passivation. Boron diffused wafer

with a J0e of 580 fA/cm2 had reduced surface recombination of merely 8 fA/cm2 after the

negative corona charging. This again demonstrates that the ONO stack has excellent chemical

passivation for boron diffused samples across all SiNx refractive indexes.

1.8 2.0 2.2 2.4 2.6

101

102

Grey: Before CC

Black: After pos. CC

τeff_1015

J0e

Refractive Index (n)

τ eff (

ms)

101

100

(a)

J0e (

fA/c

m2)

1.8 2.0 2.2 2.4 2.610

-1

100

101

Grey: Before CC

Green: After neg. CC

τeff_1015

J0e

Refractive Index (n)

τ eff (

ms)

(b)

100

101

102

103

J0e (

fA/c

m2)

Page 92: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

79

5.6 Impact of corona charging and annealing on diffused surfaces with

ONO stack

Significant improvements were obtained on the ONO passivation as shown with diffused and

undiffused n-type wafers, where immeasurably low J0s values were achieved.

This section presents and discusses the injection dependence of effective lifetime of

phosphorus diffused planar, rantex and boron diffused planar surfaces before and after corona

charging and followed by FGA at 400 ˚C.

Experimental details

To investigate the influence of corona charging, both phosphorus and boron diffusion were

performed using a 100 Ω.cm n-type FZ silicon wafer on planar or rantex surfaces. In the

fabrication of diffused wafers, the ONO stack and corona charging are identical to the baseline

process. However, the wafers were subjected to additional annealing in FG at 400 ˚C for 30

minutes after corona charging.

Results and discussion

The effect of the corona charging on diffused wafers was documented using lifetime

measurements of the deposited ONO stack, corona charging and further annealing in FG at 400

˚C in Figure 5.7.

Degradation of the surface passivation was observed on the phosphorus diffused planar after

attempting to embed corona charging with FGA, as shown in Figure 5.7a. The positive corona

charging on the phosphorus diffused planar wafer had a slight increase in the τeff at injection

levels greater than 1015 cm-3. However, lifetimes at lower injection levels decreased, which

indicates slight degradations that may be due to dehydrogenation of the bulk or the increased

defect interface at the Si-SiO2 interface [95, 177, 178]. The additional FGA to embed the

charges did not improve or recover the lifetime to initial surface passivation conditions.

Page 93: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

80

1013 1014 1015 1016

0

20

40

60

80

100

1013 1014 1015 1016

0

10

20

30

40

50

1013 1014 1015 1016

0

5

10

15

20

J0e

6.9 fA/cm2

J0e

1.9 fA/cm2

J0e

~4.0 fA/cm2

τeff (

ms)

Excess Carrier Density (cm-3)

Initial

CC 120s

CC & FGA

fit τeff

(a) (b)

J0e

5.3 fA/cm2

J0e

3.4 fA/cm2

J0e

9.6 fA/cm2

τeff (

ms)

Excess Carrier Density (cm-3)

Initial

CC 120s

CC & FGA

fit τeff

J0e

~59 fA/cm2

J0e

~25 fA/cm2

J0e

~9 fA/cm2

(c)

τeff (

ms)

Excess Carrier Density (cm-3)

Initial

CC 100s

CC & FGA

fit τeff

Fig 5.7: Minority carrier lifetime measurements of (a) phosphorus diffused planar, (b) phosphorus diffused rantex and (c) boron diffused planar surfaces, prior to corona charging, after corona charging and after the additional annealing in FG at 400 ˚C.

Improvements in the passivation were observed on the phosphorus diffused rantex and boron

diffused planar wafers after corona charging and additional FGA. The positive corona charging

on the phosphorus diffused rantex wafers, as shown in Figure 5.7b, resulted in improved

lifetimes at all injection levels, where a J0e of ~3.4 fA/cm2 was achieved. The negative corona

charging for boron diffused surfaces, as shown in Fig 5.7c, improved the surface passivation

and achieved a J0e of ~9 fA/cm2. The subsequent FGA to embed charges within the ONO stack

for both wafers showed slight degradation, but maintained an overall higher lifetime than when

performed prior to the corona charging. The cause of such degradation maybe due to decrease

Page 94: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

81

in charges or increase in surface recombination. Some of the charges may not have been

embedded into the ONO stack or were lost during the subsequent annealing step [65], moreover

it is possible that the surface passivation may have deteriorated after corona charging due to

ion bombardment [178].

5.7 Chapter summary

Diffusions are important for the operation of a conventional solar cell. This chapter details the

study of ONO passivation on light boron and phosphorus diffused wafers by examining the

effects of the thermal SiO2 thicknesses, POA with N2, SiNx refractive index, corona charging

and subsequent annealing.

Depending on the orientation and surface morphology, the deposition of corona charges onto

ONO stack to accumulate majority carriers of boron and phosphorus diffused surfaces may

improve the surface passivation. The deposition of corona charges and further annealing was

observed to improve the surface passivation for boron diffused planar and phosphorus rantex

surfaces. However, ONO passivation with light phosphorus diffusion on 100 oriented

surface need not require additional corona charging. Only mild improvement was observed

after charging, and excessive corona charging may degrade the surface and bulk lifetime due

to ion bombardment at the Si-SiO2 interface [178].

Thermal SiO2 with thicknesses from 7 to 10 nm in the ONO stack on lightly phosphorus

diffused wafers were found to provide excellent surface passivation on both planar and rantex

surfaces. A thicker thermal SiO2 was required for rantex surfaces due to the concave and

convex corners, which have slower oxide growth rates. Boron diffused planar surfaces

similarly improved with increasing thermal SiO2 thicknesses due to the resulting light doping

profiles and reduced positive charge centroid from the SiNx.

A much more complex situation was observed on the diffused wafers after the N2 POA. The

J0e on the phosphorus diffused wafers with a planar surface improved after a prolonged POA;

however, no improvements were observed on the rantex wafers. Introducing POA in N2

reduced the lifetime on both the phosphorus diffused surfaces compared with the thermal SiO2

ramped down in O2. The reduced lifetimes with POA in N2 were similarly observed for the

undiffused wafers, suggesting the N2 annealing may have contributed to bulk defects or

contaminations. The J0e for boron diffused planar wafers improved with prolonged POA in N2

Page 95: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

82

due to the reduced Auger recombination of the doping profile and the improved surface

passivation of the ONO stack.

Increasing the silane to ammonia gas ratio reduced the J0e for the phosphorus diffused planar

wafers. The decreased J0e with an increased gas ratio may be due to hydrogenation of the

surface and bulk with an increasing SiNx refractive index. The boron diffused surface achieved

improved passivation with higher SiNx refractive indexes due to the hydrogenation and lower

positive charges.

Excellent surface passivation was achieved for phosphorus and boron diffused wafers with

ONO surface passivation after corona charging. Positive corona charged ONO stack passivated

on phosphorus diffused wafer with sheet resistivity of 520 Ω/ planar and 610 Ω/ rantex

surface achieved J0e of 0.5 fA/cm2 and 1.9 fA/cm2, respectively. Negative corona charged ONO

stack passivated on boron diffused wafer with sheet resistivity of 154 Ω/ planar surface

achieved J0e of 4.8 fA/cm2.

Page 96: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

83

Chapter 6

Simulation of ONO IBC solar cells

6.1 Introduction

This chapter presents the loss analysis for an interdigitated back contact (IBC) solar cell with

ONO stack surface passivation. Simulation tools, such as EDNA2, SunSolve, OPAL2 and

Quokka3, were used to simulate the solar cell performance. The three general loss categories

are optical, recombination and resistance, which are analysed and discussed in detail.

6.2 Optical losses

The reflectance of a polished silicon surface at normal incidence can be obtained using the

Fresnel equation as shown in Eq 6.1:

z.3 = (1 − )= + ()=(1 + )= + ()= (6.1)

n: refractive index k: extinction coefficient

The reflectance of a polished silicon sample is ~35% at a 632 nm wavelength with established

n and k values of silicon [242]. To reduce the uncaptured energy from the high reflectance, the

silicon surface can be chemically textured to create isotextures or random pyramids. Chemical

etching to create a random textured pyramidal surface can reduce the overall reflection to ~12%

at wavelengths from 400–1100 nm [243]. Other surface morphologies, such as black silicon

(BSi), have gained significant interest in recent years due to their excellent absorption over a

wide range of wavelengths with reflections of ~1% [244]. However, BSi has yet to overcome

the high surface recombination caused by the increased surface area of random nanoscale

structures [244].

Surface reflection can be further reduced using an ARC, which utilises the destructive

interference of reflected light. Single or multiple layers of dielectric stacks are commonly

Page 97: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

84

deposited on high-efficiency solar cells to reduce the reflectance, such as SiO2, SiNx, TiO2,

MgF2-ZnS and SiNx-SiOx [14, 117, 245, 246].

The following section discusses simulations that were performed using OPAL2, which is an

optical simulator developed by Baker-Finch et al. that is capable of simulating different surface

morphology for a solar cell [19].

6.2.1 OPAL2 simulation inputs

The simulations in OPAL2 requires user inputs for the surface condition, planar fraction,

dielectric film thicknesses, refractive index and extinction coefficient, substrate thickness and

light trapping model to define the generation current density (Jgen) within the wafer. This

section reviews and discusses each of these conditions.

Textured surface – The limiting number of light ray bounces within the pyramids is

determined from the pyramidal angle. The number of light ray bounces for symmetric pyramids

increases to between 1 and 3 as the pyramidal angle increases from 0 to 60˚ (Appendix A).

Ideally, larger pyramidal angles are desired to increase the number of light ray bounces and

reduce the overall surface reflection.

Monocrystalline silicon texturing is typically performed using potassium hydroxide (KOH),

TMAH or sodium hydroxide (NaOH) together with isopropanol (IPA) or other surfactants. Due

to the different etching rates of the 111 and 100 oriented planes, the anisotropic etching

and masking of IPA (discussed in Chapter 7) result in the formation of randomly scattered

pyramids that are typically 3-20 µm across the base. The theoretical pyramidal angle between

the 111 and 100 silicon planes is 54.735˚. However, reported angles of individual pyramid

facets and the substrate surface plane (α) range from 49 to 53˚ [247]. The random pyramid

facets etched with a mixture of TMAH and IPA solution used on fabricated IBC solar cell

devices in this work (details in Chapter 7) were measured to have a pyramid angle of 51˚ using

electron microscopy.

Planar fraction – Incomplete texturing may occur during the texturing process, which results

in a fraction of highly reflective planar surface coverage. The measured reflectance from a

spectrophotometer can be fitted with the simulated reflectance for a planar fraction, where the

Page 98: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

85

simulated reflectance is calculated using the percent of planar region reflectance, including the

remaining percent for the defined surface morphology reflectance.

Refractive index and extinction coefficient – The n and k values were determined for the

applied ARC materials. The ONO stack consists of three different refractive indexes and

extinction coefficients.

The initial bottom layer of the ONO stack is thermal SiO2 with n and k values obtain from Palik

[248]. The second and third layer of the PECVD SiNx and SiOx deposited using an Oxford

Plasmalab 100 have refractive indexes and extinction coefficients that were measured using a

JA Woollam M2000D ellipsometer. The refractive indexes and extinction coefficients for

various SiNx compositions with different SiH4:NH3 gas ratio are plotted in Figure 6.1. The

measured refractive indexes and extinction coefficients increase with the SiH4:NH3 gas ratio

due to the increased absorption of the silicon-rich SiNx. The measured n and k of the third layer

(SiOx) are similar to previously published results [249].

400 600 800 10001.8

2.0

2.2

2.4

2.6

2.8

3.0

3.2

400 600 800 100010

-5

10-4

10-3

10-2

10-1

100

101

Re

fra

ctive

In

de

x (

n)

Wavelength (nm)

SiH4:NH

3 ratio

6:1

5:2

9:5

4:3

1:1

3:4

2:5

(b)

Extin

ctio

n c

oe

ffic

ien

t (k

)

Wavelength (nm)

SiH4:NH

3 ratio

6:1

5:2

9:5

4:3

(a)

1:1

3:4

2:5

Fig 6.1: Measured (a) refractive indexes and (b) extinction coefficients for PECVD SiNx with different SiH4:NH3 gas ratios deposited with an Oxford Plasmalab 100.

Substrate thickness – The substrate thickness affects the spectral absorption. For a

conventional silicon solar cell, the increased cell thickness improves the light absorption,

especially at longer wavelengths (>900 nm).

Light trapping model – Light trapping improves the light absorption because the optical path

length is increased to several times the actual device thickness. Light trapping is commonly

Page 99: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

86

achieved using a textured surface by coupling light obliquely into the silicon, which changes

the angle at which light travels. Figure 6.3 shows light travelling into two different cell structure.

In Figure 6.2a, the light ray only passes through once in the planar substrate with no light

trapping. However, in Figure 6.2b, the light ray bounces multiple times within the substrate,

increasing the optical light path with a front pyramidal texture and back reflector. The increased

optical light path can be defined using the optical pathlength enhancement factor (Z). The

optical pathlength enhancement factor for a solar cell can be determined by measuring its

external quantum efficiency (EQE) [249].

Fig 6.2: Optical pathlength with (a) no light trapping applied on planar silicon wafer and (b) improved light trapping with a pyramidal textured surface and a back reflector.

A Lambertian scatterer offers a high degree of randomisation to the incoming light rays and

significantly improves the light trapping capability of a solar cell. The ‘ideal’ cases for

Lambertian light models available in OPAL2 are shown in Table 6.1. The differences in the

models are due to the underlying assumptions. Yablonovitch et al., for example, accounted for

weak absorption within the silicon substrate [250], Tiedje et al. assumed no absorption within

the escape cones [251] and Green’s analytical solution considered that a fraction of 1/n2 of the

incident light is scattered into the loss cone [252]. These ‘ideal’ limits only present the

absorptance for Lambertian light paths and do not represent an upper limit.

These Lambertian light models provide a benchmark to evaluate the ‘ideal’ light trapping for

a solar cell. However, solar cells commonly use a model with a constant optical pathlength

enhancement of 2 ≤ Z ≤ 50. The absorption within the substrate for a lumped constant Z can

be re-written as Eq 6.2.

k = 1 − WA (6.2)

w: thickness of the substrate α: absorption coefficient

(a) (b)

Light

Page 100: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

87

Table 6.1: The absorption and optical pathlength enhancements for different Lambertian models [250-252]

Absorption in Substrate Optical Pathlength Enhancement

k = 1 − WA = 4=

k = 1 − WA1 − (1 − =)WA = ln(1 + 4=d)d

k = 4=d1 + 4=d = 4 + lnf= + (1 − =)WAgd

w: thickness of the substrate α: absorption coefficient n: refractive index

6.2.2 Simulation of ONO with OPAL2

The ARC simulation for the ONO stack was performed to determine the optimal thicknesses

for each layer of thermal SiO2, PECVD SiNx and SiOx. The parameters used for the simulation

in OPAL2 are listed in Table 6.2.

Table 6.2: Simulation parameters for the ONO stack using OPAL2.

Parameter Details

Surface morphology Random – Upright pyramids

Characterization angle 51˚

Planar fraction 0%

Film 1 PECVD SiOx [249]

Film 2 PECVD SiNx 1.94 (Fig 6.2)

Film 3 Thermal SiO2 [248]

Substrate Crystalline Si [242]

Substrate thickness 230 µm

Light trapping model = 4 + lnf= + (1 − =)WAgd

Simulated wavelength region 250–1200 nm

Two simulations were performed with a variety of thermal SiO2, PECVD SiNx and SiOx

thicknesses. The first simulation presented in Figure 6.3a varied the thermal SiO2 and SiNx

thicknesses while maintaining SiOx thickness at 90 nm. The second simulation had an optimum

Page 101: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

88

SiOx thicknesses ranging between 69–110 nm, for all simulated thermal SiO2 and SiNx

thicknesses.

Two trends were observed in Figures 6.3a-b. The first trend shows nearly identical counter

plots with various simulated thermal SiO2 and SiNx thicknesses. The largest difference in the

JGen for both plots of 0.23 mA/cm2 occurred at thermal SiO2 and SiNx thicknesses of 0 nm.

However, at thermal SiO2 and SiNx thicknesses above 6 nm, the differences in JGen reduced to

<0.08 mA/cm2. This demonstrates that a fixed SiOx thickness of 90 nm is nearly optimum for

the simulated range of thermal SiO2 and SiNx thicknesses above 6 nm.

The second trend observed in both graphs shows that for an optimum SiNx thickness, the JGen

reduces with the increasing thickness of the thermal SiO2. The initial thermal SiO2 layer on the

ONO stack behaves parasitically from an optical perspective. This is as expected since the

refractive index of thermal SiO2 is lower than that of simulated SiNx and is therefore

detrimental to the transmission of light into the silicon. Since the best optical performance is

achieved in the absence of the interfacial thermal SiO2, the optimal ARC is simulated with

thermal SiO2, PECVD SiNx and SiOx thickness of 0 nm, 65 nm and 90 nm (SiNx-SiOx),

respectively. However, as presented in Section 5.4.1, the surface passivation on the light

phosphorus diffused rantex wafer improved with the increasing oxide thickness, possibly

saturating above 12 nm. This creates an optimisation problem that balances the trade-off

between the optical transparency of the front ARC versus the passivation quality.

43.65 43.20

43.60

43.50

43.40

43.20

43.0042.80

43.65

43.60

43.50

43.40

43.20

43.00

0 5 10 15 20 25 300

10

20

30

40

50

60

Thic

kne

ss o

f S

iNx

(nm

)

Thickness of thermal SiO2 (nm)

42.00

42.40

42.80

43.20

43.60

JGen

(mA/cm2)(a)

0 5 10 15 20 25 300

10

20

30

40

50

60

(b)

Thic

kne

ss o

f S

iNx

(nm

)

Thickness of thermal SiO2 (nm)

42.00

42.40

42.80

43.20

43.60

JGen

(mA/cm2)

Fig 6.3: Simulation of ONO stack with varying thermal SiO2 and SiNx thicknesses using OPAL 2. (a) The SiOx thickness is 90 nm. (b) The SiOx thickness is optimum for each thermal SiO2 and SiNx thickness pair.

Page 102: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

89

While this balance introduces some complications, it coincides with the thickness for which

the passivation quality is nearly optimal for both the rear boron and phosphorus diffusions - a

thermal SiO2 thickness of 11–16 nm, as presented in Section 5.4.1 while an excellent front

optical performance with a JGen reduction of merely 0.15 mA/cm2 is expected.

0.0

0.5

1.0

1.5

2.0

Sp

ectr

al in

t. (

W/m

2/n

m)

400 600 800 10000

5

10

15

20

25

30

35

40 Simulated SiN

x

Simulated NO

Simulated ONO

AM1.5G

Rantex ONO

Wavelength (nm)

Re

flecta

nce (

%)

Fig 6.4: Reflection measurements of the ONO rantex wafer with layer thickness for the thermal SiO2 of 18 nm, PECVD SiNx of 48 nm and SiOx of 90 nm compared to the simulated optimised nitride-oxide and single layer SiNx ARC.

In Figure 6.4, the measured reflectance of an ONO rantex wafer is compared with a simulated

double layer SiNx-SiOx and single layer SiNx ARC. The input parameters used to fit the

measured reflectance of the fabricated ONO rantex in OPAL2 had a thermal SiO2 thickness of

18 nm and PECVD SiNx n632 of 1.92 and SiOx thicknesses of 48 and 90 nm, respectively. The

simulated double ARC from the SiNx-SiOx performed better at shorter wavelengths compared

to the ONO stack. Within the range of optimal thermal oxide thicknesses for passivation, the

ONO stack performed almost as well as the DARC SiNx-SiOx (NO), and both are capable of

achieving an overall lower reflection compared to a single layer SiNx ARC n632 of 1.92. The

measured JGen for each condition of NO, ONO and SiNx in Figure 6.5 with the parameters listed

in Table 6.2 were 43.71, 43.55 and 43.30 mA/cm2, respectively.

Page 103: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

90

6.2.3 Free carrier absorption

Free carrier absorption (FCA) occurs commonly within the highly doped regions of silicon

solar cells. The photon energy is absorbed by the free carriers, which are excited to higher

energy levels within a similar band state. The FCA does not generate additional electron-hole

pairs to assist in the current generation.

Transmission and reflectance measurements were performed on a heavily phosphorus diffused

polished silicon wafer with a resistivity of 14 Ω/. In Figure 6.5a, the FCA can be observed

for wavelengths greater than 1000 nm.

1000 1200 14000

10

20

30

40

50

60

70

80

800 1000 1200 140010-1

100

101

102

103

(b)

Reflectio

n /

1-t

ransm

issio

n (

%)

Wavelength (nm)

R Diffused

(1-T) Diffused

R Undiffused

(1-T) Undiffused

(a)

1020

cm-3

1019 cm-3

1018

cm-3

Absorp

tion C

off

icie

nt

α (c

m-1)

Wavelength (nm)

Intrinsic Si

Fig 6.5: (a) Measured transmission and reflectance of a heavily phosphorus diffused wafer with a resistivity of ~14 ohms/ and (b) the modelled absorption coefficient.

The parameterised absorption coefficient proposed by Baker-Finch et al. for phosphorus and

boron doping densities can be calculated using Eq 6.3 and 6.4 [253].

oe, = 1.68 × 10M_=. (6.3)

oe,p = 1.82 × 10>e=.9 (6.4)

αFCA,P: Absorption coefficient for phosphorus αFCA,B: Absorption coefficient for boron N: Dopant concentration

The FCA plotted in Figure 6.5b using Eq 6.3 shows the absorption coefficients for various

phosphorus concentrations. Increasing the phosphorus concentration increased the absorption

Page 104: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

91

at longer wavelengths and, therefore, fine tuning the diffusion profile to minimise FCA is

required to reduce the losses from photogeneration at these lower photon energies.

Substituting Eq 6.3 or 6.4 into Eq 6.5, the FCA for either phosphorus or boron diffusion profiles

can be calculated as a function of the wavelength. This provides the fraction of long wavelength

photons (≥1000 nm) absorbed by free carriers.

ko = 1 − exp(−o(3) J)320l39

(6.5)

imax: The node at which Ndop(z) falls below 1014 cm-3 z: diffusion profile depth θ: angle of transmission through the silicon

The simulations for the FCA in a cell can be accounted for by calculating the FCA absorption

for a specific diffusion profile and defining a wavelength-dependent reflection layer using the

SunSolve software package.

6.2.4 Simulation of FCA using SunSolve

Simulations for the FCA were performed using the IBC solar cell structure shown in Figure

6.6. The diffusion profiles were simulated as Gaussian functions with defined surface doping

peaks to achieve the intended sheet resistivity. Heavy phosphorus diffusion for contact dots

were simulated with a surface doping peak of 1.5 × 1020 cm-3 with a sheet resistivity of 15 Ω/.

Two other phosphorus diffusion profiles for the front and rear surface fields were simulated

with surface doping peaks of 1 × 1019 cm-3 to obtain resistivities of 450 Ω/ and 800 Ω/,

respectively. The diffusion profile of the boron diffusion with a surface doping peak of 2 × 1019

cm-3 was modelled using a Gaussian function to obtain a sheet resistivity of 160 Ω/.

The FCA absorption for the diffusions were calculated using Eq 6.5 with a wavelength from

1000–1300 nm. The calculated absorption was implemented into a wavelength-dependent

reflection, absorption and transmission layer, and simulated with the IBC solar cell structure

(Figure 6.6) using the SunSolve ray tracing simulator. SunSolve implements a Monte Carlo ray

tracing method to solve Jsc for various dielectric layers and solar cell surfaces. The simulations

with and without the FCA were performed on an IBC solar cell structure (front rantex, rear

planar) with ARC using the ONO stack on both front and rear surfaces.

Page 105: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

92

The calculated FCA in Table 6.3 for an IBC solar cell shows a total loss of ~0.06 mA/cm2. The

FCA for the defined IBC solar cell design (chapter 7) with ONO stack is negligible and

contributes to only ~0.14% of the total Jsc without FCA.

Table 6.3: Simulation for the FCA using the SunSolve ray tracing software.

6.3 Recombination losses of ONO IBC solar cell

The recombination losses component in a solar cell include those from the surface (surface

SRH), the emitter (surface SRH, emitter SRH and Auger), the bulk (bulk SRH, Auger and

radiative) and metal contact surfaces (surface SRH). One method to quantify each component

of the recombination losses is with the saturation current (J0), where J0 can be determined either

through experiments or control wafers during cell processing. The total cell recombination can

be calculated based on the total area weighted J0 for each component or region.

In this section, simulations for an ONO IBC solar cell were conducted using three-dimensional

modelling in Quokka3. The cross-section of a simulated IBC solar cell is presented in Figure

6.6, and the baseline simulation parameters are given in Table 6.3.

Fig 6.6: Schematic diagram for an IBC solar cell simulated using Quokka3 (not to scale).

Simulation of ONO IBC solar cell structure Simulated Jsc

(mA/cm2)

Weighted area

(%)

Reduction in Jsc

(mA/cm2)

No FCA 42.65 - -

Full front area phosphorus diffusion (450 Ω/) 42.61 100 0.04

Full rear area phosphorus diffusion (15 Ω/) 42.19 1.3 0.006

Full rear area boron diffusion (160 Ω/) 42.61 8.9 0.004

Full rear area phosphorus diffusion (750 Ω/) 42.64 89.8 0.006

Surface passivation and ARC

Boron diffusion

Heavy phosphorus diffusion

Light phosphorus diffusion

Metal Pitch

Page 106: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

93

Table 6.3: Simulation parameters for the Quokka3 simulation were based on experimental values from findings mostly presented in this thesis [14, 254]. Pitch size of cell is 500 µm. Optical simulations that were incorporated into Quokka3 were obtained using OPAL2. The path enhancement parameters used were from a previously characterised IBC solar cell [249].

Area Side Properties Reference Values

Rantex light phosphorus 100% FSF Sheet resistance 450 Ω/

J0e 10 fA/cm2

Planar light phosphorus BSF Sheet resistance 750 Ω/

J0e 2.5 fA/cm2

Localised phosphorus dot

contact

1.3% Rear Sheet resistance 15 Ω/

J0e 230 fA/cm2

Finger size 150 µm

Localised boron dot emitter 8.9% Rear Sheet resistance 160 Ω/

J0e 20 fA/cm2

Dot pitch 140 µm

Finger size 350 µm

Metallised contact 0.34% Rear Contact resistivity Chapter 7 2 × 10-5

Ω.cm2

Metallised phosphorus dot

contact

Rear J0e Chapter 7 500 fA/cm2

Metallised boron contact dot Rear J0e Chapter 7 1300

fA/cm2

Bulk 100 ms

Resistivity n-type 100 Ω.cm

Cell thickness 230 µm

Series resistance External circuit series

resistance

Chapter 7 0.21 Ω.cm2

Shunt resistance External circuit shunt

resistance

Chapter 7 10 kΩ.cm2

Ideality factor 1.0

Optical modelling Front ONO [19] OPAL2

Z parameterization [255] Z0, Zinf, Zp

Page 107: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

94

The subsequent subsections describe the performance of cell devices simulated using Quokka3

over a range of cell parameters. The performed simulations include:

- Front and rear surface recombination with light phosphorus diffusion.

- Boron emitter recombination and the dot area fraction.

- Heavy phosphorus recombination and the dot area fraction.

- Different pitch lengths while maintaining similar boron emitters, heavy phosphorus and

metal contact fraction.

- Bulk lifetime, bulk resistivity and edge recombination.

- Series and shunt resistances.

6.3.1 Simulation of front and rear surface recombinations with phosphorus

surface field

The implementation of light phosphorus diffusion as a surface field has shown to benefit cell

processing and the solar cell structure. This includes a reduced sensitivity for the front surface

recombination with the surface field, bulk gettering, contaminant inhibition during cell

processing and assistance in the lateral current transport [232, 256]. In Chapter 7, light

phosphorus diffusion was implemented on both the front and rear surfaces of the IBC solar cell

to inhibit contamination and reduce the sensitivity of the front surface recombination.

The phosphorus diffusion at the front surface creates an electric field that repeals minority

carriers (holes) from recombining at the surface [20]. However, the phosphorus diffusion

profile has to be sufficiently light and shallow to limit both Auger recombination and FCA.

Therefore, this section first investigates and simulates the different diffusion profiles and

surface recombination in terms of the performance for IBC solar cells. The cell performance

was investigated using various front and rear surface recombination, and, finally, the cell

performance was investigated by varying the rear surface recombination and sheet resistivity.

Simulation details:

The simulation details for the front and rear surface recombinations are divided into three

subsections.

Page 108: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

95

Different diffusion profiles – Four phosphorus diffusion profiles were simulated in EDNA2

using a Gaussian function with the first three surface dopant concentration of 4 × 1019 cm-3,

1019 cm-3 and 1018 cm-3 with a depth factor (Zf) of 0.117 µm, and the final diffusion profile with

a dopant concentration of 1019 cm-3 and a Zf of 0.354 µm. The J0e of the defined diffusion

profiles were simulated with a range of surface recombinations (Seff). The performance of IBC

solar cells was simulated using baseline parameters in Quokka3, however, the front J0e and

sheet resistivity were varied based on the EDNA2 simulation results.

Front and rear surface recombination – Simulations for an IBC solar cell were performed

using the baseline parameters, except the front and rear surface recombinations were varied

with J0e from 1 to 20 fA/cm2.

Rear surface recombination and sheet resistivity – Simulations for an IBC solar cell were

performed using the baseline parameters, except the rear surface recombination and sheet

resistivity were varied between 3 to 20 fA/cm2 and 200 to 2400 Ω/, respectively. A surface

recombination lower limit was defined with a J0e of 3 fA/cm2 (Seff ~0 cm/s) and sheet resistivity

of 200 Ω/. The lower limit was validated using a Gaussian function for the boron diffusion

profile using EDNA2 with a surface dopant concentration of 5 × 1018 cm-3.

Results and discussion:

The results for each condition are discussed in the following subsections.

Different diffusion profiles - The simulated J0e with different diffusion profiles and Seff are

plotted in Figure 6.7a. It was found that the J0e is dependent on the surface recombination,

surface dopant concentration and junction depth. The simulation shows that the surface

recombination becomes less substantive with an increased surface dopant concentration.

However, a sheet resistivity of 155 Ω/ with a shallower profile (smaller Zf) has higher J0e

recombination due to an increased Auger recombination within the diffusion profile.

The range of J0e for the phosphorus diffusion profiles from EDNA2 was utilised in Quokka3

simulations. The simulation results shown in Figure 6.7b show that the efficiency of the IBC

solar cell is increasingly sensitive to the quality of the front surface passivation as the surface

dopant concentration and sheet resistance decrease. However, the cell efficiency becomes

limited with an increasing surface dopant concentration and sheet resistance. Therefore, having

Page 109: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

96

a fine-tuned light diffusion can improve the robustness of the front surface passivation without

being detrimental to the overall cell performance.

100

101

102

103

0

10

20

30

40

50

60

100

101

102

103

23.0

23.5

24.0

24.5

25.0

25.5

26.0

155 Ω/

(a)

155 Ω/

1870 Ω/

468 Ω/

J0e (

fA/c

m2)

Seff

(cm/s)

Zf of 0.177 µm

Npeak

4×1019

Npeak

1×1019

Npeak

1×1018

Zf of 0.354 µm

Npeak

1×1019

Eff

icie

ncy (

%)

Seff

(cm/s)

1870 Ω/ 468 Ω/ 155 Ω/ 155 Ω/

No FSF

(b)

Fig 6.7: (a) Simulations for various diffusion profiles using EDNA2 as a function of Seff. The diffusion profile was simulated using a Gaussian function with a similar depth factor. (b) Simulations for cell efficiency using Quokka3 with a select range of Seff

Front and rear surface recombination - The simulation results of an IBC solar efficiency,

Voc, Jsc and fill factor for the front and rear surface recombinations are presented in Figure 6.8.

The cell efficiency results in Figure 6.8 shows that increasing the overall surface recombination

reduces the cell efficiency. It is observed that the effect of the cell efficiency is more prominent

for the front surface J0e as compared to the rear.

The main impact on the cell efficiency was due to the surface recombination, which is reflected

in the Voc. The efficiency and Voc both decrease as the surface J0e increases. The second

observed impact in cell efficiency is due to the decreased Jsc with increasing surface

recombination. Most carriers are generated closer to the front surface and high front surface

recombination can substantially reduce the amount of minority carriers collected at the rear

junction. These effects are more pronounced for the front J0e compared to the rear with

decreases in both Jsc and FF, as plotted in Figures 6.8c-d.

Page 110: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

97

82.20

82.40

82.60

82.80

83.00

42.50

42.52

42.54

712

718

724

728

732

736

742

24.80

25.00

25.20

25.40

25.60

25.80

26.00

5 10 15 20

5

10

15

20

Rea

r J

0e (

fA/c

m2)

Front J0e

(fA/cm2)

76.0

77.0

78.0

79.0

80.0

81.0

82.0

83.0

FF (%)(d) (c)

(b) (a)

5 10 15 20

5

10

15

20

Rea

r J

0e (

fA/c

m2)

Front J0e

(fA/cm2)

40.00

40.50

41.00

41.50

42.00

42.50

43.00J

sc (mA.cm

-2)

5 10 15 20

5

10

15

20

Re

ar

J0e (

fA/c

m2)

Front J0e

(fA/cm2)

700

705

710

715

720

725

730

735

740

745

Voc

(mV)

5 10 15 20

5

10

15

20R

ear

J0e (

fA/c

m2)

Front J0e

(fA/cm2)

22.00

22.50

23.00

23.50

24.00

24.50

25.00

25.50

26.00

26.50

Eff (%)

Fig 6.8: Simulation results for the (a) efficiency, (b) Voc, (c) Jsc and (d) FF for a range of J0e on both the front and rear surface passivations.

Rear surface recombination and sheet resistivity – Driving dopants deeper into the silicon

bulk requires relatively high temperatures and long annealing processing. In Chapter 7, which

discusses the fabrication process of IBC solar cells, the rear phosphorus diffusion for the back

surface field (BSF) is introduced at the beginning of the cell fabrication process. Introducing

the BSF at early stages of the fabrication process allows the phosphorus diffusion dopants to

be driven deeper into the silicon with subsequent high-temperature processing. Therefore, the

simulations presented in the following section investigate the different rear sheet resistivities

and J0e in terms of the IBC solar cell performance.

The simulation results for the IBC solar efficiency, Voc, Jsc and FF for the rear surface

recombination are presented in Figure 6.9. The simulation did not input specific diffusion

profiles, but employed lumped parameters for Rsheet and J0,skin to solve for the lateral conduction

Page 111: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

98

within the phosphorus diffusion [16, 257]. At a fixed rear J0e recombination value, the

simulation results in Figure 6.9 show improvements in the efficiency and FF with a heavier

rear sheet resistance.

82.30

82.40

82.50

82.60

82.70

714

718

722

726

25.00

25.10

25.20

25.30

25.40

25.50

42.526

42.530

500 1000 1500 2000

5

10

15

20

Re

ar

Ph

osp

horu

s J

0e (

fA/c

m2)

Sheet Resistance (Ω/ )

82.00

82.20

82.40

82.60

82.80

FF (%)

500 1000 1500 2000

5

10

15

20

Re

ar

Ph

osp

ho

rus J

0e (

fA/c

m2)

Sheet Resistance (Ω/ )

712

716

720

724

728

Voc

(mV)

500 1000 1500 2000

5

10

15

20

Re

ar

Ph

osp

ho

rus J

0e (

fA/c

m2)

Sheet Resistance (Ω/ )

24.80

25.00

25.20

25.40

25.60

25.80

Eff (%)

500 1000 1500 2000

5

10

15

20

Re

ar

Ph

osp

horu

s J

0e (

fA/c

m2)

Sheet Resistance (Ω/ )

42.40

42.44

42.48

42.52

42.56

42.60

Jsc

(mA.cm-2)

(a) (b)

(c) (d)

Fig 6.9: Simulation results for the (a) efficiency, (b) Voc, (c) Jsc and (d) FF for various sheet resistivities and J0e for a rear phosphorus diffused surface.

A heavier rear phosphorus sheet resistance is desirable to improve lateral conduction provided

that J0e remains sufficiently low, as shown in Figure 6.9a and 6.9d. Furthermore, a heavier rear

phosphorus diffusion is preferable to having one in the front as this minimizes the loss of short

wavelengths from Auger recombination.

Page 112: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

99

6.3.2 Simulation of rear boron-doped emitter

As discussed briefly in Chapter 5, the formation of a boron emitter forms a p-n junction, which

is needed for the fundamental operations of n-type IBC solar cells. The boron emitter provides

a region of high hole and low electron conductivity enabling it to selectively collect

photogenerated holes [258]. The minority carriers generated within the n-type bulk are

collected into the boron emitter during the n-type IBC solar cell operation. Chapter 5 showed

that the surface recombination of the ONO stack on boron diffusion is comparatively higher

than light phosphorus diffusion. The balance of the J0e and emitter fraction must be simulated

to optimise the collection efficiency and resistive effects with the other solar cell parameters.

The collection efficiency is defined as the number of minority carriers collected at the junction

to the number of photons entering the silicon.

This section investigates and simulates the performance of an IBC solar cell as a function of

J0e and boron emitter area fraction.

Simulation details

Simulations of an IBC solar cell were performed with the baseline parameters, except the J0e

for the boron diffusion emitter was varied between 6–100 fA/cm2 and the rear coverage of the

boron emitter was varied between 2–30%. The simulated J0e range is comparable to the

experimental surface passivation results of the ONO stack for boron diffused surfaces (Chapter

5).

Results and discussion

The simulation results for the IBC solar efficiency, Voc, Jsc and FF of the boron emitter

recombination are presented in Figure 6.10. Two trends were observed in Figure 6.10a. First,

the cell efficiency increased with a reduced J0e. Second, an increased boron emitter area with a

sufficiently low J0e was required to reach higher efficiencies. Moreover, the cell efficiency with

a smaller boron diffusion area (2–12%) was observed to have a higher tolerance to J0e as

compared to a larger rear boron diffusion area (>12%). In Figure 6.10c, the increased boron

diffusion area was observed to cause an increase in the Jsc, regardless of the J0e behaviour.

Page 113: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

100

As the J0e decreased and boron emitter area increased, Figure 6.10d shows that the fill factor

improved due to the lower resistive effects and improved collection of minority carriers. The

resistive effects observed in Figure 6.10d were verified using the Suns-Voc simulation with

Quokka3 (results not presented here), which showed the fill factor to be similar for a given J0e

across different boron diffusion areas from 2–30%.

82.80

82.60

82.70

82.60

42.54

42.52

710

715

720

722

726

728

25.00

25.60

25.20

25.50

25.40

5 10 15 20 25 30

20

40

60

80

100

J0e (

fA.c

m-2)

Rear Boron Area (%)

82.0

82.2

82.4

82.6

82.8

FF (%)

5 10 15 20 25 30

20

40

60

80

100

J0e (

fA.c

m-2)

Rear Boron Area (%)

40.00

40.50

41.00

41.50

42.00

42.50

43.00J

sc (mA.cm

-2)

5 10 15 20 25 30

20

40

60

80

100

J0e (

fA.c

m-2)

Rear Boron Area (%)

700

704

708

712

716

720

724

728

732

Voc

(mV)

(d)(c)

(b)(a)

5 10 15 20 25 30

20

40

60

80

100

J0e (

fA.c

m-2)

Rear Boron Area (%)

24.80

25.00

25.20

25.40

25.60

25.80

Eff (%)

Fig 6.10: Simulations for the (a) efficiency, (b) Voc, (c) Jsc and (d) fill factor as functions of the boron emitter area fraction (%) and boron emitter J0e values.

The simulated J0e for the boron diffused region contributes to much higher recombination (>6

fA/cm2) as compared to most of the other defined cell region parameters given in Table 6.3.

Therefore, the simulation results presented in Figure 6.10b show that an increasing boron

diffusion area causes a decreased Voc.

Page 114: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

101

6.3.3 Simulations for rear phosphorus-doped contact region

The heavily doped phosphorus diffusion on the n-type IBC solar cell is primarily to achieve a

low contact resistivity between the silicon and metal contacts and improve the collection of

majority carriers. It is more difficult to achieve ohmic contact towards an n-type silicon wafer

due to the potential barrier from the bandgap offset and high defect interface states from the

silicon-metal interface (see Appendix B). Heavily doped phosphorus diffusion reduces the

potential barrier between metals and silicon to more easily allow electrons to tunnel through.

Thus, this section shows the investigations and simulations for the IBC solar cell performance

over a range of J0e and area fractions for heavily diffused phosphorus region.

Simulation details

Simulations for an IBC solar cell were performed with the baseline parameters, except the J0e

for heavy phosphorus diffusion was varied between 80–300 fA/cm2, and the rear coverage of

the phosphorus diffusion area was varied between 0.4–5%. A lower area fraction of 0.4% was

not simulated due to the limitations of the fabrication process that limit achieving such a

resolution.

Results and discussion

The simulation results for the IBC solar efficiency, Voc, Jsc and FF of the heavy phosphorus

recombination are presented in Figure 6.11. It was observed that the efficiency in Figure 6.11a

decreased with an increased J0e and rear phosphorus diffusion area.

The Voc decreased with an increasing heavy phosphorus diffusion area, as presented in Figure

6.11b. The fill factor shown in Figure 6.11d for a given J0e may improve marginally with an

increasing phosphorus area due to an increased majority carrier collection. However, the

recombination of heavy phosphorus diffusion dominates with simulated J0e >80 fA/cm2, which

causes the efficiency to decrease with the increased heavy phosphorus diffusion area. Only a

small fraction of the heavy phosphorus diffusion area (0.4–0.8%) is needed to achieve high

efficiency.

Page 115: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

102

82.60

82.75

82.7042.528

42.532

720

722

724

726

728

730

25.30

25.40

25.50

25.60

25.65

1 2 3 4 5

100

150

200

250

300

J0e (

fA.c

m-2)

Rear Phos Area (%)

82.0

82.2

82.4

82.6

82.8

FF (%)

1 2 3 4 5

100

150

200

250

300

J0e (

fA.c

m-2)

Rear Phos Area (%)

42.00

42.20

42.40

42.60

42.80

Jsc

(mA.cm-2)

1 2 3 4 5

100

150

200

250

300

J0e (

fA.c

m-2)

Rear Phos Area (%)

710

714

718

722

726

730

Voc

(mV)

1 2 3 4 5

100

150

200

250

300J

0e (

fA.c

m-2)

Rear Phos Area (%)

24.00

24.50

25.00

25.50

26.00

Eff (%)(a) (b)

(c) (d)

Fig 6.11: Simulations for the (a) efficiency, (b) Voc, (c) Jsc and (d) fill factor on heavy phosphorus area fraction (%) with a range of phosphorus J0e values.

6.3.4 Simulation of different pitch sizes for the ONO IBC solar cell

The pitch of an ONO IBC solar cell in this work is defined between the adjacent phosphorus

contact dots, as illustrated in Figure 6.6. The pitch length of an IBC solar cell affects the

efficiency of the carriers collected. A larger pitch would require the carriers to travel further

before collection.

In this section, different IBC solar cell pitches were simulated while maintaining the area ratio

of all other parameters (boron emitter, heavy phosphorus and metal dot opening).

Page 116: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

103

Simulation details

Simulations for an IBC solar cell were performed with the baseline parameters, except the pitch

of the cell was varied between 400 to 1000 µm while maintaining a similar area ratio for all

other parameters.

Results and discussion

The simulation results for the IBC solar efficiency, Voc, Jsc and FF of different pitch sizes are

presented in Figure 6.12. The plotted cell efficiencies in Figure 6.12a decreased as the pitch

size increase.

The decreased cell efficiency is mostly due to the decreased fill factor, as shown in Figure

6.12b. The Voc of the cell is maintained for different pitch sizes due to the similar fractional

area of the defined parameters. However, the fill factor and Jsc decreased due to increased series

resistance. As the pitch size increased, the carriers that are generated at the front surface have

to travel further before being collected at the rear junction of the device [259].

400 500 600 700 800 900 100081

82

83

Jsc

FF

(b)

FF

Jsc

Pitch (µm)

Fill

Facto

r (%

)

42.40

42.44

42.48

42.52

42.56

42.60

Jsc (

mA

/cm

2)

400 500 600 700 800 900 1000724

726

728

730

732

Eff

Voc

Eff

Pitch (µm)

Voc

(mV

)

(a)

Voc

25.0

25.1

25.2

25.3

25.4

25.5

25.6

25.7

Eff

icie

ncy(

(%)

Fig 6.12: Simulations for the (a) efficiency and Voc and the (b) Jsc and fill factor as functions of the pitch size while maintaining similar features for the area ratio.

Reducing the pitch size improved the cell efficiency due to lower series resistance effect.

However, limitations in the cell fabrication process place a lower limit on the pitch size as this

requires greater precision in the cell alignment process.

Page 117: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

104

6.3.5 Simulation of bulk lifetime, resistivity and edge recombination

High bulk lifetimes in wafers are required for high-efficiency IBC solar cells. The carriers

generated at the front are must travel through the thickness of the wafer to be collected at the

rear junction. High resistivity wafers can readily achieve higher bulk lifetimes due to the lower

Auger recombination. However, edge recombination is reported to have a larger effect for

wafers with higher resistivities [11].

In this section, n-type 1–100 Ω.cm resistivity wafers with various bulk lifetime were

investigated with and without edge recombination.

Simulation details

Two sets of simulations were performed for IBC solar cells in this section. The first simulation

was performed using the baseline parameters, except the wafer resistivity and bulk lifetime

were varied between 1–100 Ω.cm and 1–100 ms, respectively. No edge recombination was

considered in the simulation.

The second simulation was performed with a quarter of a full area IBC solar cells accompanied

by two extended regions away from the active cell area of 3 cm. The extended region was

shaded to simulate edge recombination. The simulations performed were similar to the baseline

parameters; however, three changes were made to the cell design. First, the diffusion and metal

opening contacts were defined as a line as compared to dots before. Second, the defined lines

had similar area ratio as the defined dots and open dot contacts. Third, for each bulk resistivity,

the bulk lifetime was set using the parametrised Auger limit [17].

Results and discussion

The first simulation results for the IBC solar efficiency, Voc, Jsc and FF of different wafer

resistivities and bulk lifetimes are presented in Figure 6.13. The wafer resistivities in Figure

6.13 are given in terms of the doping concentration. The simulated results in Figure 6.13a show

two trends. First, the efficiency improved with higher bulk lifetimes. Second, wafers with

higher resistivities were capable of achieving a higher cell efficiency, Voc and Jsc.

Wafers with higher resistivities can achieve higher lifetimes and longer diffusion lengths due

to the lower Auger recombination within the bulk. The longer diffusion lengths increased the

Page 118: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

105

Jsc, and the higher bulk lifetimes increased the Voc. The simulation results show that higher

bulk lifetimes for wafers with higher resistivities should improve cell efficiency. However, the

simulation results did not include edge recombination of the cell.

Fig 6.13: Simulations for the (a) efficiency, (b) Voc, (c) Jsc and (d) fill factor as functions of the wafer resistivity and bulk lifetime.

The simulation results for IBC solar cells with defined edge recombination at different wafer

resistivities are plotted in Figure 6.1a. The results show that efficiency was reduced from

between 0.28–0.40%. The differences in efficiency were observed to decreases as the bulk

resistivity decreased. This is due to the fill factor having less of an influence on the edge

recombination and the carriers are more easily collected with lower bulk resistivities, as shown

with simulation results in Figure 6.14b. The bulk lifetime for the parameterised Auger limit

and a specific cell thickness of 230 µm allowed the high resistivity wafers of 100 Ω.cm to

achieve higher efficiency for the IBC solar cell as compared to lower resistivities [17].

Page 119: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

106

1014

1015

81.5

82.0

82.5

83.0(b)

FF no edge

Jsc

no edge

FF edge

Jsc

edge

ND Doping Density (cm

-3)

Fill

Fa

cto

r (%

)

41.5

42.0

42.5

Jsc (

mA

/cm

2)

1014

1015

715

720

725

730

Eff with edge

Eff no edge

Voc

no edge

Eff edge

Voc

edge

Eff no edge

ND Doping Density (cm

-3)

Vo

c (

mV

)(a)

24.5

25.0

25.5

Eff

icie

ncy(

(%)

Fig 6.14: Simulations for the (a) efficiency and Voc and the (b) Jsc and fill factor as functions of the wafer resistivity with and without edge recombination.

6.3.6 Simulation of series and shunt resistance

Series and shunt resistance in a solar cell reduce the efficiency by dissipating power in the

resistances. The total series resistance within the cell can be ascertained from the current

resistance through the emitter and bulk, the contact resistance of the metal contact on silicon

(reviewed in Appendix B) and the grid resistance of the metal fingers and busbar. The shunt

resistance persists due to current leakage within the device from fabrication defects, such as

overlapping metallisation on the diffusion overlap regions.

In this section, a range of series and shunt resistances were simulated to evaluate the

performance of an IBC solar cell. As part of the solar cell design, only contributions to the

series resistance from the fingers and busbar were assessed.

Simulation and calculation details

Simulations for the IBC solar cell were performed using the baseline parameters, except the

series and shunt resistances were varied between 0.1 to 2.0 Ω.cm2 and 100 to 100k Ω.cm2,

respectively.

The rear metal grid resistance of an IBC solar cell was calculated using Eq 6.6.

z. = 13 oo uo= (6.6)

Page 120: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

107

WF: Width of metallised finger HF: Thickness of metallised finger LF: Length of metallised finger P: Pitch size of the cell

Results and discussion

The simulation results for the IBC solar efficiency, Voc, Jsc and FF with different series and

shunt resistances are presented in Figure 6.15. The increased series resistance and reduced

shunt resistance caused a decrease in cell efficiency. The series and shunt resistances affected

both the Jsc and fill factor, as shown in Figures 6.15c-d, respectively. However, the series

resistance did not affect the Voc as no net current flow was obtained during the operation of the

cell at Voc. The cause of the high series and low shunt resistances in a solar cell are normally

due to poor cell design or fabrication defects.

The grid resistance from fingers and busbar contributes in part to the total series resistance of

a solar cell. The calculated metal grid resistance shown in Figure 6.15a with a ~92% rear

coverage on a 4 cm2 cell and 3 µm thickness of Al on the rear is ~0.07 Ω.cm2, which contributes

to a loss in fill factor of ~0.35% according to Eq 6.7. Other metals, either Ag or Cu, with similar

area coverages and thicknesses show a grid resistance of ~0.04 Ω.cm2, with a lower loss in the

fill factor of ~0.2% as plotted in Figure 6.15.

∆ = ; z.c.1n@1 (6.7)

The use of Ag for metallisation at the rear is preferred when compared to Al and Cu due to the

low resistivity and lower absorption at longer wavelengths when used concurrently as a back

reflector. However, the adhesion of specific metals to silicon or dielectrics must be considered

during the cell fabrication process.

Page 121: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

108

83.00

82.50

82.00

81.00

80.00

75.00

42.50

25.50

25.00

24.50

23.50

0.1 1.0102

103

104

105

Sh

un

t R

esis

tan

ce

.cm

2)

Series Resistance (Ω.cm2)

65

70

75

80

(d) (c)

(b)(a)

0.1 1.0102

103

104

105

42.30

Sh

un

t R

esis

tan

ce

.cm

2)

Series Resistance (Ω.cm2)

40.00

41.00

42.00

43.00

0.1 1.0102

103

104

105

725

726Sh

un

t R

esis

tan

ce

.cm

2)

Series Resistance (Ω.cm2)

710

715

720

725

730

727

0.1 1.0102

103

104

105S

hu

nt

Re

sis

tan

ce

.cm

2)

Series Resistance (Ω.cm2)

19.00

21.00

23.00

25.00

Fig 6.15: Simulations for the (a) efficiency, (b) Voc, (c) Jsc and (d) fill factor as functions of the series and shunt resistances.

10-1 100 101

10-2

10-1

100

101

10-1 100 101

75

76

77

78

79

80

81

82

83(a)

Meta

llisa

tion s

erie

s r

esis

tance

.cm

2)

Metallisation finger thickness (µm)

Aluminium Silver

Copper

Fill

facto

r (%

)

Metallisation finger thickness (µm)

Aluminium

Silver

Copper

(b)

Fig 6.16: (a) Series resistance as calculated for an IBC solar cell with a pitch size of 500 µm for a cell size of 4 cm2, and (b) the behaviour of the fill factor as a function of the metallisation finger thickness.

Page 122: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

109

6.4 Chapter Summary

This chapter covers the detailed effects of the optical losses, recombination losses, series

resistances and shunt resistances. The calculated and modelled solar cell parameters present an

in-depth analysis for the ways to improve cell performance by individually reducing each of

the losses.

Thermal SiO2 within the ONO stack behaves parasitically from an optical perspective due to

the thermal SiO2 refractive index mismatch between the silicon and the PECVD SiNx. However,

considering the passivation of the ONO stack on diffused surfaces, the optimum ONO stack

for an ARC with a thermal SiO2 thickness of 11–16 nm has a PECVD SiNx n632 of ~1.94 with

a thickness between 53–48 nm and a PECVD SiOx thickness of ~92 nm.

Device simulations for the ONO IBC solar cell were performed over a range of surface and

bulk recombination, diffusion fractions, pitch sizes, bulk resistivities and series and shunt

resistances. The information obtained was applied to the cell design and fabrication processes,

which are discussed in Chapter 7. The contact resistivity of metal to silicon contacts was not

covered in this chapter, but is reviewed in Appendix B.

Page 123: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

110

Chapter 7

7.1 Introduction

Building on the simulation results presented in Chapter 6, the fabrication, characterization and

analysis of high-efficiency n-type ONO IBC solar cells are detailed and discussed in this

chapter. The fabrication of ONO IBC solar cells involves the use of photolithography and

numerous high-temperature processing steps. Four key aspects of the cell design and

fabrication are also investigated, including the implementation of the boron-rich layer gettering,

design of the ONO surface passivation on the IBC solar cell, effects of positive corona charging

on the completed cells and an alternative metallisation method based on lift-off processing.

Lastly, the improvements culminated in a champion ONO IBC solar cell with a certified cell

efficiency of 25.0%.

7.2 Fabrication sequence

The ONO IBC cells were fabricated using n-type FZ 100 Ω.cm silicon wafers. The fabrication

process involved multiple processes, including chemical cleaning, pre-oxidation of the wafer,

gettering of the metal contaminants, diffusion, chemical masking, photolithography patterning

of localised diffusions, surface texturing, ONO surface passivation and metallisation of the

fingers and busbar. Each of the processing steps are discussed in the following section. An

overall visual illustration and flow chart of the ONO IBC solar cell fabrication process are

presented in Figures 7.1 and 7.2.

Fabrication and analysis of ONO IBC solar

cells

Page 124: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

111

Fig 7.1: Visual illustration of the fabrication sequence for an ONO IBC solar cell with rear localised n+ and p+ emitters with localised contacts.

7.2.1 Chemical cleaning

The fabrication of IBC cells involves multiple high-temperature steps (>700 ˚C). Subjecting

the wafers to repeated high-temperature furnaces and photoresist spinning steps poses the risk

of metal and organic contamination on the wafer. Metal contaminants unintentionally diffuse

into the bulk and act as active recombination centres to reduce the bulk lifetime of the wafer

[260, 261]. The exposure of organic contaminants can potentially reduce the electronic

performance on the silicon surface [262]. To reduce the contamination risk of both organic and

metallic impurities, RCA cleaning is commonly used within the silicon semiconductor industry

to chemically clean silicon wafers [263].

The RCA cleaning consists of two chemical cleaning stages. The first, commonly known as

RCA1, removes organics and several species of metal contaminants. Silicon wafers are

subjected to a RCA1 solution, which consists of DI H2O, electronic grade NH3OH and H2O2.

The H2O2 grows a thin layer of oxide on the silicon and removes organic surface films, while

the NH3OH dissolves and removes metal impurities such as gold, copper and chromium [263].

The wafers were cleaned in the RCA1 solution for approximately ten minutes at 55–65 ˚C. The

1) Pre-oxidation

2) POCl3 gettering

3) Etch wafer

4) L POCl3 diffusion

5) SiO2-Si3N4 mask

6) Photolithography

7) H POCl3 diffusion

8) SiO2-Si3N4 mask

9) Photolithography

10) BBr3 diffusion

11) SiO2-Si3N4 mask

12) Random Texture

13) L POCl3 diffusion

14) ONO passivation

15) Photolithography

16) Al evaporation

17) Photolithography and

Al etch

Page 125: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

112

silicon wafers were then rinsed and dipped into a mixture of HCl, HF and H2O to remove the

sacrificial oxide layer [264].

Fig 7.2: Flow diagram of the fabrication sequence for an ONO IBC solar cell.

The second cleaning stage, commonly known as RCA2, removes the remaining metal

contaminants [263]. The wafers are subjected to a RCA2 solution, which consists of DI H2O,

electronic grade HCl and H2O2. Similar to the first cleaning, the H2O2 grows a thin layer of

oxide, while the HCl removes the alkali ions (Al+3, Fe+3 and Mg+2), which were insoluble in

the first NH3OH solution. The wafers were cleaned in RCA2 for approximately ten minutes at

55–65 ˚C. Finally, the wafers were subjected to another mixture of HF, DI H2O and HCl to

remove the second sacrificial oxide [264].

Photolithography

2. Pre-oxidation

of wafer

1. Damage etch

4. Thinning of

wafer by etching

5. Light

phosphorus

diffusion

RCA Clean

RCA Clean

3. Phosphorus

gettering

6. Masking

RCA Clean

RCA Clean

7. Define

phosphorus

dots

Photolithography

8. Heavy

phosphorus

diffusion

TMAH RCA Clean

9. Masking

RCA Clean

10. Define boron

dots opening

11. Boron

diffusion

TMAH RCA Clean

12. Masking

13. Random

Texturing

TMAH RCA Clean

14. Light

phosphorus

diffusion

RCA Clean

RCA Clean

15. Thermal

oxidation

16. PECVD SiNx

17. PECVD SiOx

RCA Clean

18. Define metal

dot opening

19. Metal

evaporation

Photolithography

RIE

20. Metal

separation

Photolithography

21. Sintering of

metal contacts

Phosphorus Dots

Boron Dots ONO

Metallisation

Page 126: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

113

7.2.2 Pre-oxidation of wafers

FZ wafers may have defects within the bulk as either interstitials or vacancies, which develop

during the ingot growth. These point defects are dependent on the growth rate (V) and thermal

gradient (G) of the ingot. Silicon interstitials are formed with a slow growth rate (below the

critical value of V/G), while vacancies are formed at a fast growth rate (above the critical value

of V/G) [265].

(a) (b)

Fig 7.3: Photoluminescence (PL) images of ring defects on FZ wafers during the IBC solar cell fabrication. (a) Lifetime calibrated PL image of a wafer taken after phosphorus diffusion. (b) Uncalibrated PL image of a wafer with ring defect taken after completing high temperatures wafer processing.

To suppress the interstitial and vacancy-related defects and strengthen the ingot, wafer

manufacturers introduce N2 doping during the growth of the silicon ingot crystal [266, 267],

with most of the N2 retaining its molecular form because it is a more favourable configuration

within the crystal [266]. However, N2 doping also introduces electrically active point defects

at deep levels, which lowers the bulk lifetime of the wafers and often exhibits radially-

symmetric low lifetime regions on the wafer [268]. The PL imaging for these wafer defects are

presented in Figure 7.3 as taken during the IBC cell fabrication process. Abe et al. and Grant

et al. reported that annealing FZ wafers at temperatures above 1000 ˚C for 10 minutes or more

in O2 can permanently deactivate these defects and restore the bulk lifetime [199, 200, 266].

Therefore, the FZ silicon wafers selected for cell processing were pre-oxidized to prevent these

deep level active point defects from forming during cell processing [200]. The established

process for cell fabrication in this work was to perform a silicon etch in TMAH for ten minutes

at 85 ˚C, followed by RCA cleaning and loading into the oxidation furnace at 900 ˚C for

Page 127: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

114

oxidation at 1050 ˚C for an hour in O2 ambient with a 3% mixture of Trans 1,2-

dicholoroethylene (TLC).

7.2.3 Gettering

Gettering is the process of removing impurities or contaminants from an active region to a pre-

defined region in a silicon wafer. The basic gettering mechanisms can be defined as either

‘relaxation’ or ‘segregation’ gettering [269].

Relaxation gettering involves the precipitation of impurities at regions with high precipitation

site densities, such as surfaces or dislocations. As relaxation gettering is limited by the

solubility limit of the contaminants, lower annealing temperatures are required to reduce the

solubility limit and increase the maximum gettering effectiveness. However, the gettering

process may then require longer anneal times at these lower temperatures due to the reduced

diffusivity of the contaminants.

Segregation gettering is due to differences in the solubility of the impurities present within a

region of the bulk. For example, contaminants are driven towards the heavily phosphorus

diffusion due to higher solubility limit for metal contaminants within the diffused region.

Segregation gettering is not limited by the solubility limit of the wafer bulk and, therefore, can

reduce the contaminant’s concentration below its solubility limits, even at high temperatures.

Phosphorus diffusion gettering may involve a much more complicated mechanism depending

on the processing conditions. High phosphorus concentrations that are insoluble at the surface

of the wafers during diffusion causes the formation of SiP precipitates. These precipitates have

been reported to be very effective gettering sites for Ni and Pt [270, 271]. Schröter et al.

demonstrated that phosphorus diffusion gettering is dependent on the formation of PSG and

the plateau region of the diffusion [272]. The injection of silicon interstitials is thought to be

the mechanism by which the gettering occurs. Syre et al. reported that another possible

phosphorus gettering mechanism during phosphorus diffusion is due to the growth of SiO2

precipitates at vacancy rich regions, in which the SiO2 precipitates act as gettering centres [273].

The conditions for phosphorus gettering may consist of multiple complex mechanisms that are

dependent on the diffusion process. However, the general consensus is that heavy phosphorus

diffusion can effectively getter metal contaminants [272, 274].

Page 128: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

115

The fabrication of IBC solar cell wafers has two phosphorus diffusion gettering steps. The first

gettering was performed after the pre-oxidation, which was subsequently etched in TMAH.

The second gettering was implemented during the formation of the heavy phosphorus diffusion.

The front surface of the wafers, which was intentionally exposed to heavy phosphorus diffusion,

had a sheet resistivity of ~15 Ω/. The heavy phosphorus diffused layers on the front surfaces

were retained until the cells reached the front surface texturing stage. During the boron emitter

formation on the rear patterned dots, BRL were maintained for gettering purposes [275]. The

implementation and analysis of the BRL gettering are discussed in Section 7.3.1.

7.2.4 Masking layer

Thermal SiO2 is commonly used as a masking layer for both boron and phosphorus diffusions

due to the slow diffusivity of the dopants. A 10 nm thermal SiO2 is reportedly required to mask

the boron diffusion for approximately 2 hours at 1000 ˚C [276]. However, it was found that

such a thin thermal SiO2 was unable to mask the boron diffusion required for cell fabrication.

A thermal SiO2 thickness of ~120 nm grown in the lab was required to mask the boron diffusion

performed at 940 C for 40 minutes. Two possible explanations for the required thicker thermal

oxide may be due to the presence of H2 within the SiO2. The literature has shown that dry SiO2

containing H2 increases the diffusivity of both boron and phosphorus [277]. Another possible

reason may be that the boron reacts with SiO2 to form B2O3, which causes boron to diffuse into

the silicon.

As the fabrication of an ONO IBC solar cell requires three masking steps, repeated growths of

multiple thermal dry or wet SiO2 layers for long durations at high temperatures would

drastically change the diffusion profiles. To minimise the thermal budget required for growing

the masking layers, a thin SiO2 and LPCVD Si3N4 stack was implemented to mask the thermal

diffusions.

Thermally grown Si3N4 has been reported to be an effective pinhole-free barrier against boron

diffusion [278]. The combination of thermal SiO2 capped with LPCVD Si3N4 was tested and

found to mask both boron and phosphorus diffusions more effectively than the oxide alone. A

thin thermal SiO2 layer of 2–5 nm was initially grown on the wafer to prevent contamination

from the LPCVD SiN3N4 deposition while minimising changes to the diffusion profile. The

thermal SiO2 was grown with cell wafers loaded into the furnace at 700 ˚C and ramped up to

850 ˚C in O2. As soon as the temperature reached the set point of 850 ˚C, the wafers were

Page 129: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

116

ramped down in O2 and unloaded at 700 ˚C. The oxidized wafers then had LPCVD Si3N4

deposited to thicknesses of 50–70 nm. The wafers loaded into the LPCVD chamber were

deposited at 780 ˚C with a DCS and NH3 flow of 40 and 120 sccm, respectively, and a chamber

pressure of 450 mTorr. In addition to testing against boron and phosphorus diffusions, the

stacks were subjected to further testing with TMAH chemical etching. The SiO2-Si3N4 stack

was etched in TMAH at 85 ˚C for an hour and showed no etched pits, indicating that the SiO2-

Si3N4 stack was free of pinholes and capable of masking the TMAH etch and texturing solution.

7.2.5 Photolithography patterning

MicroChemicals positive photoresists of AZ1518 and AZ4562 were used to pattern the rear

dots and protect the surface passivation from exposure to buffer HF (BHF). The differences

between the two photoresists are in their viscosity and thickness after spin coating. The AZ1518

is a thin photoresist capable of achieving thicknesses between 1.5–2.5 µm, whereas the AZ4562

is a thicker photoresist capable of achieving thicknesses between 6–15 µm. The AZ1518

photoresist was used specifically to define the rear patterning, whereas the thicker AZ4562

photoresist was used to protect the peaks and troughs of the textured surfaces from chemical

etching.

A successful spin coating process for the photoresist requires the following: good adhesion

between the photoresist and the dielectric surface, the photoresist is free of particles and air

bubbles, a clean formation and opening for the patterns and clean removal of the photoresist.

The photolithography was performed using the sequence of steps described below:

1. First prebake: prior to spinning, the wafers were baked in an oven at 100–110 ˚C for

10–15 minutes to remove any moisture.

2. Photoresist spin cast: a slow continuous flow of photoresist was squeezed from a pipet

to avoid the formation of bubbles. The wafers with photoresist were spun at 2500 –

3000 rpm for 45 seconds. After spinning, the uniformly coated wafers were rested in a

fume hood for ten minutes.

3. Second Prebake: the wafers were baked at 90 ˚C for 10–20 minutes in an oven to reduce

the solvent concentration. After unloading, the cells rested for an additional ten minutes

at room temperature.

4. UV exposure: the wafers were loaded onto the mask aligner and a chrome photomask

was used to define the regions for UV exposure for 45 seconds.

Page 130: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

117

5. Development: The AZ326 MIF developer was used to develop the exposed regions.

The wafers were agitated in the developer for 90 seconds and spin rinsed at 2000 rpm.

A visual inspection was performed to ensure the adhesion of the photoresist to the

wafers.

6. Hard bake: the wafers were hard baked in an oven at 110 ˚C for 20 minutes to increase

the chemical and physical stability of the photoresist.

7. Etching of the exposed dielectric layers: RIE was performed on the wafers to remove

the exposed SiO2-Si3N4 layer. The wafers were then submerged in a 25% BHF solution

until the features became hydrophobic.

8. Removal of the photoresist: a sequence of three acetone baths and spin rinsing were

used to remove the photoresist layers.

In a high humidity environment conditions (rainy day), the photoresist was found to adhere

poorly onto the SiOx layer. In such cases, the MicroChemicals TI PRIME adhesion promoter

was spun on to the SiOx prior to the photoresist to mitigate the poor adhesion.

7.2.6 Diffusions process

All diffusions were performed in a Tempress 6-inch tube furnace with boron tribromide (BBr3)

and phosphorus oxychloride (POCl3) as the dopant sources for boron and phosphorus

diffusions, respectively. As indicated in Figure 7.2, the diffusions required for the cell

fabrication included heavy phosphorus diffusion for gettering, heavy phosphorus diffusion for

the localised contact, localised boron diffusion for the emitter formation, and light phosphorus

diffusion for the front and back surface fields.

Prior to thermal diffusion, all wafers were cleaned with freshly prepared RCA and HF dips.

The wafers were loaded at 700 ˚C in O2 and ramped up to the set temperatures at a rate of 10

˚C/min in ambient O2. Phosphorus diffusions and the phosphorus gettering process were

performed with gas flow rates of 2.5 slm of N2, 130 sccm of the N2 carrier gas with POCl3 and

65 sccm of O2. The boron emitter diffusion had a gas flow rate of 2.5 slm N2, 18 sccm for N2

as the carrier gas for BBr3 and 18 sccm of O2. After the diffusion, the furnaces were ramped

down to 700 ˚C in O2 at a rate of 10 ˚C/min, except for the boron and phosphorus gettering

processes. Phosphorus gettering diffusions were ramped down in ambient N2 at a rate of 5

˚C/min and unloaded at 600 ˚C, whereas the boron diffusions were ramped down in either

Page 131: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

118

ambient O2 or N2 at a rate of 10 ˚C/min. Figure 7.4 plots the thermal profiles for each of the

performed diffusions.

0 20 40 60 80 100 120 140500

600

700

800

900

1000

780 °C

850 °C

860 °C

Te

mp

era

ture

(°C

)

Time (min)

Boron

POCl Gettering

Heavy POCl Light POCl

940 °C

Fig 7.4: Thermal profiles for the heavy phosphorus, front surface field (FSF) and back surface field (BSF) with light phosphorus, boron emitter and phosphorus gettering.

7.2.7 Texturing

Random alkaline texturing can be performed using a mixture of TMAH, H2O, orthosilicic acid

and isopropanol (IPA) [279-281]. During the initial development of the ONO IBC solar cells,

a series of experiments were performed to optimise the texturing process over a range of

solution temperatures, IPA surfactants, orthosilicic acid crystals, TMAH concentrations and

silicon surface conditions. However, no conclusive outcomes were drawn from the various

experiments. Repeating similar texturing conditions on the IBC cells had poor consistency and

occasionally incomplete texturing with the observed planar regions examined under the

microscope. The average variation in the reflection measured across 600–900 nm varies from

10–15%. The poor consistency and texturing uniformity were postulated to be due to IPA

evaporation and the detachment of hydrogen bubbles.

The formation of hydrogen in the TMAH texturing solution has been reported to be both

important for and a hindrance to the formation of randomly textured pyramids. The formation

of pyramids is thought to be due to the generation of hydrogen bubbles that act as a shield

during the TMAH-Si etching and with the preferential etching of the 100 plane over the 111

plane [282, 283]. However, there are reports that suggest the accumulation of large hydrogen

Page 132: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

119

bubbles (2–3 mm) adhering to the surface of the wafer for too long may inhibit pyramid

nucleation and uniformity [284, 285]. Therefore, adding an appropriate amount of IPA has

been understood to improve the wettability of silicon [283], hinder the etching rates of certain

crystallographic planes [286] and remove the hydrogen bubbles from the surface during

texturing to improve the uniformity with alkaline textured solutions [284, 287, 288].

Moreover, Amouzgar et al. reported that the additional gas lift effect, which removes hydrogen

bubbles on the surface, reduces the required etch time and improves the uniformity of the

random texture [285]. However, the addition of IPA does not guarantee a uniformly random

texture. The evaporation of IPA over the duration of the texturing process may cause uniformity

issues [289]. As the boiling temperature of IPA is ~82 ˚C, Jiang et al. observed non-uniformity

of a planarized surface when texturing was performed at 80 ˚C [289]. They therefore suggested

a two-step temperature process to avoid IPA evaporation; loading the wafers at 75 ˚C followed

by a slow increase at 1 ˚C/min to 80 ˚C. They proposed that a low temperature of 75 ˚C

promotes a large amount of pyramid nucleation sites, and increasing the etching temperature

to 80 ˚C improves the growth of the formed pyramids [289]. However, no satisfactory and

consistent texturing results were achieved after considering and implementing most of the

procedures suggested in the literature. To improve the texturing repeatability, the IPA

surfactant was substituted for an industrial additive, monoTEX-F with TMAH.

(a) (b)

Fig 7.5: SEM images taken from Liang et al. [290] of the pyramids formed using the TMAH texturing solution with the (a) monoTEX addictive and the (b) IPA surfactant.

Page 133: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

120

The industrial monoTEX-F additive has been reported to provide repeatable and consistent

textured reflectances of ~11% at wavelengths from 400–1100 nm [243]. The surface

preparation for texturing with monoTEX-F requires a native RCA1 chemical oxide, whereas

the conventional IPA texturing solution requires wafers to be hydrophobic. The pyramidal

formation in Figure 7.5 shows that smaller pyramids are formed with the monoTEX-F and that

the texturing process etches less silicon relative to the IPA surfactant. The measured reflectance

of the monoTEX-F wafers was on average of 10–12% at wavelengths from 400–900 nm with

no flat surfaces observed under the microscope.

7.2.8 Metal contact and finger patterning

The baseline fabrication process to form metal contact and finger patterns on the ANU IBC

solar cell are based on the etching of thermally evaporated Al layers. First, the dielectric layers

are locally etched using a photolithography mask. After removing the photoresist, Al is

thermally evaporated onto the rear of the cell, followed by photoresist patterning for the p and

n fingers and a final etch of the exposed Al using a mixture of phosphoric acid to form the

defined p and n fingers.

Although this same process was successfully implemented for previous record-breaking cells

[14], the process was marred by the occasional formation of poor contacts. The prevailing

hypothesis for the source of randomly occurring poor contacts are the formation of native

oxides prior to metal evaporation or the surface contamination of the contact surfaces from

exposure to the photoresist and organic solvents during resist removal immediately prior to

metal evaporation. In an attempt to resolve this ongoing issue, two different metallisation

processes were tested for the fabrication of the ONO IBC solar cells.

The first method is the baseline method. Prior to metal evaporation, all wafers were subjected

to either a BHF or a quick diluted HF-dip to remove the native oxides that formed on the

exposed Si surfaces. The metallisation was performed in a modified Varian thermal bell jar

evaporator, capable of evaporating three boats individually in a single run without breaking the

chamber vacuum. Al pellets with a purity of 99.999% (5N) were loaded onto the three boats

and wafers placed one meter above the metal source to avoid unnecessary heating from the

evaporation. A supplied power of 550–650 W was applied across the metal boat after the

chamber was pumped down to 7.5 × 10-5 Torr. An initial pre-melt of the metal was performed

while the shutter was closed. The pre-melt stage is important as it has been observed to achieve

Page 134: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

121

a low contact resistivity [254]. This is thought to be due to contaminants from the metal source

or tungsten boat and surrounding crucible shield that evaporates before the targeted metal has

melted [254]. Therefore, the shutter remains closed to prevent these initial evaporations with

possible contaminants from being deposited onto the wafer. After the metal on the boat had

melted sufficiently over a defined period, the shutters were opened and deposition began with

an average rate of 6–10 Å/sec. This process was similarly performed for the remaining two

boats with the other materials. The temperature measured at the control wafer placed nearly a

meter high from the sources did not exceed 80 ˚C after all three depositions were performed.

The chamber was vented in N2 gas after reaching the set targeted thickness of ~1 µm. Two

subsequent runs were performed on the cell to reach a total Al thickness of ~3 µm.

The second “lift-off” metallisation method, which attempts to resolve the contact problem, is

discussed in a later section.

7.3 Improvements to the ONO IBC cells fabrication process

This section describes the process and performance of the various fabricated IBC solar cells.

Firstly, we examine BRL as a source for gettering during cell processing. Previously developed

surface passivations for the front SiNx-SiOx and rear SiO2-Si3N4 stacks on IBC solar cells were

analysed and compared with the developed ONO surface passivation [14]. The effect of corona

charging on the completed cells was examined and is presented. Finally, the performance of

different metallisation schemes on similar ONO IBC solar cells was evaluated.

7.3.1 Boron-rich layer gettering

The boron-rich layer (BRL) gettering process is less established compared with phosphorus

diffusion gettering for cell fabrication. However, the implementation of a BRL gettering

sequence in this IBC cell fabrication process substantially reduces the amount of additional

high-temperature processing steps required for phosphorus gettering and removes the slow

diffused contaminants within the boron diffused region.

The BRL is a silicon boride layer with electrically inactive boron atoms that form between the

boron silicate glass (BSG) and the heavily boron-doped region in the silicon. The mechanism

of BRL gettering was suggested by Phang et al. as being caused by the segregation effect of

Page 135: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

122

impurities into the BRL [275]. A total of 99.9% of Fe defects in intentionally contaminated

monocrystalline FZ silicon wafers were reported to be gettered into the BRL without significant

bulk degradation [274, 275]. However, the gettered impurities were released back into the

silicon bulk if the BRL was oxidised into BSG at high temperatures [275]. Moreover, the

surface passivation quality was limited due to the high surface concentration of boron, which

is coupled with the formation of the BRL. To retain the gettering effectiveness of the BRL and

improve the surface passivation quality, the BRL can be chemically removed using boiling

HNO3. The surface concentration can then be lowered for surface passivation by driving the

diffusion profile deeper into the bulk.

In this section, the lifetime of the wafers with the BRL and phosphorus gettering was

investigated for various type of wafers and resistivities.

Experimental details

To investigate the impact of BRL gettering, different dopant types (n- and p-type), growth

conditions (FZ and Cz), and resistivities (1 and 100 Ω.cm) were subjected to (i) a boron

diffusion without a BRL, (ii) a boron diffusion with a BRL and (iii) a heavy phosphorus

diffusion. A set of control wafers were considered without any diffusion. The conditions under

which the experiments for boron diffusion with and without a BRL and phosphorus diffusion

were performed similar to the cell fabrication process (Section 7.2.6). The ramp down from

940 to 700 ˚C was performed in either pure ambient N2 to retain the BRL or in pure ambient

O2 to oxidise the BRL. The measured sheet resistivity of the boron diffusions with and without

the BRL were approximately 60 and 120 Ω/, respectively, and sheet resistivity for the

phosphorus diffusion was ~30 Ω/.

All diffused and control wafers were chemically oxidised with boiling nitric acid and etched in

TMAH. Thereafter, they were passivated with an identical ONO dielectric stack, and the

injection carrier lifetime was measured at ∆n = 1014 cm-3.

Results and discussion

The results presented in Figure 7.6a show that regardless of the doping concentration and

growth method of the wafers, the boron diffusion without the BRL leads to a lower lifetime

Page 136: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

123

than with the BRL, heavy phosphorus diffusion or for the control wafer without diffusion. All

wafers with the ONO surface passivation had a J0s ≤ 1 fA/cm2. Hence, the lifetimes reported

here can be considered as representative of the bulk lifetime. In the case of the n-type 100 Ω·cm

FZ wafers used for the IBC cells, boron diffusion without BRL degraded the lifetime from the

58 ms measured in the control wafer to 15 ms. In contrast, boron diffusion with BRL improved

the lifetime to 84 ms. No bulk degradation was observed for the boron diffusions with a BRL

(contrary to other reports [291, 292]), possibly due to differences in the deposition conditions.

1013 1014 1015 1016

0

10

20

30

40

50

60

70

No BRL BRL Phos Control10-1

100

101

102

Life

tim

e a

t ∆n

= 1

01

4 c

m-3

(m

s)

n-type FZ 100 Ω.cm

n-type CZ 1 Ω.cm

p-type FZ 1 Ω.cm(a)

No BRL

BRL

Total cell J0e

~10 fA/cm2

τbulk

~35ms

Eff

ective

Life

tim

e (

ms)

Excess Carrier Density (cm-3)

Batch 4

Batch 3

Fitting

Total cell J0e

~10 fA/cm2

τbulk

~63 ms

(b)

Fig 7.6: (a) Boron diffusion with and without a BRL and phosphorus diffusion gettering were compared for various samples with identical ONO surface passivations with J0s ≤1 fA/cm2. (b) Photoconductance measurements on the two of the best performing ONO IBC solar cell from batch 3 and batch 4 (details of batches in Section 7.3.2). The cell from batch 3 had no BRL gettering whereas the cell from batch 4 had BRL gettering. An implied Voc at 10 fA/cm2 was simulated to be >740 mV.

On a batch of IBC cells incorporating BRL gettering (batch 4, which will be discussed in

Section 7.3.2), very high bulk lifetimes were achieved at the end of the cell fabrication process.

The lifetimes as measured using a Sinton PCD system with the coil over an active cell area are

presented in Figure 7.6b, which demonstrates a τeff above 50 ms at excess carrier density (∆n)

approximately 1014 cm3. The average bulk lifetime measured across 6 completed cell wafers

before metallisation was ~46 ms. Such high bulk lifetimes were not observed in previous cell

fabrication batches. This strongly suggests that the BRL gettering in combination with heavy

phosphorus diffusion gettering layers, which were maintained throughout most of the high-

temperature fabrication processes, played a major role in maintaining and improving the bulk

lifetimes.

Page 137: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

124

7.3.2 Surface passivation on IBC solar cell

The ONO dielectric stack provided improved passivation on the light phosphorus diffusion

rantex, as presented in Chapter 5. However, the fabricated IBC solar cells have boron diffusion

rear area coverage of 5–17%. Four batches of cells were fabricated to evaluate the performance

of the ONO surface passivation and ARC on the IBC solar cells.

The first and second batches of cells were passivated using a technique previously developed

at ANU for IBC solar cells where a front surface was passivated with a PECVD SiNx-SiOx

stack and the rear passivation thermal SiO2 and LPCVD Si3N4 stack [14]. The first batch had

no light phosphorus diffusion on either the front or rear surfaces. The second batch had a light

phosphorus diffusion only at the rear surface. The third batch had similar ONO passivation on

both the front and rear surfaces, with a SiNx refractive index of ~1.93. The fourth batch had

ONO passivation on both the front and rear; however, the front SiNx had a refractive index n632

of ~1.93 and the rear had a SiNx refractive index n632 of ~2.10. Positive charges largely come

from SiNx at low refractive indexes of 1.85–1.93, which may cause poor passivation on the

boron diffused region (Chapter 5). Therefore, the higher SiNx refractive index is utilised at the

rear to improve the passivation on the boron diffused region.

Experimental details

The first batch had a similar fabrication sequence as that shown in Figure 7.2; however, light

phosphorus diffusion (steps 5 and 14) and ONO (steps 15–17) were not implemented.

Texturing was performed with TMAH and IPA solution mixture. The front SiNx-SiOx and rear

SiO2-Si3N4 surface passivation stacks were generated as described in a previous publication

[14].

The fabrication of the second batch had a similar fabrication sequence as that shown in Figure

7.2, however, the rear had a full area heavy phosphorus diffusion, the front had a light

phosphorus diffusion (step 14) and the ONO was not implemented (steps 15–17). The rear

heavy phosphorus diffusion was etched back with a single-sided etching mask (between steps

12–13) while retaining the heavy diffusion dots. Texturing was performed with TMAH and

IPA solution mixture. The front SiNx-SiOx and rear SiO2-Si3N4 surface passivation stacks were

generated similarly to a previous publication [14].

Page 138: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

125

The fabrication of the third and fourth batches had similar fabrication sequences as shown in

Figure 7.2. However, there are three differences between the two batches in the rear SiNx on

the ONO stack and the texturing surfactant that was used. The third batch used TMAH with

IPA for the texturing solution, a SiNx refractive index n632 of ~1.93 was used for the rear ONO

stack and no BRL gettering was performed. The fourth batch used TMAH with MonoTex

additives for the texturing solution, a SiNx refractive index n632 of ~2.10 for the rear ONO stack

and BRL gettering was performed.

None of the cells described in this section had corona charging, and differences in the cell

fabrication processes are summarised in Table 7.1.

Table 7.1: The different fabrication processes and cell designs from the baseline sequence given in Section 7.2

Batches Differences in the cell process and design

Batch 1 No light phosphorus diffusion on the front or rear surfaces of the cell

Front passivation and ARC: PECVD SiNx-SiOx

Rear passivation: Thermal SiO2 – LPCVD Si3N4

Batch 2 No light phosphorus diffusion on the front surface of the cell

Front passivation and ARC: PECVD SiNx-SiOx

Rear passivation: Thermal SiO2 - LPCVD Si3N4

Batch 3 Front and rear passivation ONO with SiNx n ~1.93

Batch 4 BRL gettering

Texturing solution using MonoTex and TMAH

Front ONO passivation with SiNx n ~1.93

Rear ONO passivation with SiNx n ~2.10

The fabricated cells measured with dark and light I-V were analysed and simulated with an

‘enhanced’ equivalent circuit to measure the non-ideal recombination. In an ideal solar cell, the

basic equivalent circuit consists of a diode, which represents the p-n junction, a current

generation component and series and shunt resistances. The basic equivalent circuit model of

a solar cell was modified to describe the non-ideal recombination using two diodes with an

occasional third diode and a resistor in series, which is defined as the resistance-limited

enhanced recombination [293].

Page 139: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

126

Results and discussion

Test wafers were processed simultaneously for each cell batches. The area weighted J0 of the

cell’s diffusions and surfaces were calculated based on the J0 of test wafers and area coverage

summarised in Tables 7.2–7.5.

Table 7.2: Measured surface passivation for an IBC solar cell using test wafers from batch 1. Front surface with SiNx-SiOx and rear surface with SiO2-Si3N4. Cell wafer thicknesses of ~220 µm.

Region Passivation Coverage (%)

Sheet Resistivity (Ω/)

J0 (fA/cm2)

Area Weighted J0 (fA/cm2)

Boron Planar SiO2-Si3N4 Rear ~9.5 80 37.5 3.6

Heavy Phosphorus Planar

SiO2-Si3N4 Rear ~1.6 11 194 3.1

Undiffused Planar SiO2-Si3N4 Rear ~88.9 - 16.5 14.7

Undiffused Texture SiNx-SiOx Front 100 - 8.3 8.3

Table 7.3: Measured surface passivation for an IBC solar cell using test wafers from batch 2. Front surface with SiNx-SiOx and rear surface with SiO2-Si3N4. Cell wafer thicknesses of ~220 µm.

Region Passivation Coverage (%)

Sheet Resistivity (Ω/)

J0 (fA/cm2)

Area Weighted J0 (fA/cm2)

Boron Planar SiO2-Si3N4 Rear ~9.5 82 35.8 3.4

Heavy Phosphorus Planar

SiO2-Si3N4 Rear ~1.6 37 80 1.3

Light Phosphorus Planar

SiO2-Si3N4 Rear ~88.9 ~310 8.0 7.1

Undiffused Texture SiNx-SiOx Front 100 - 6.7 6.7

Table 7.4: Measured surface passivation for an IBC solar cell using test wafers from batch 3. Front and rear surfaces with ONO (n SiNx ~1.93). Cell wafer thicknesses of ~270 µm.

Region Passivation Coverage (%)

Sheet Resistivity (Ω/)

J0 (fA/cm2)

Area Weighted J0 (fA/cm2)

Boron Planar ONO Rear ~9.5 140 21 2.0

Heavy Phosphorus Planar

ONO Rear ~1.6 15 194 3.1

Light Phosphorus Planar

ONO Rear ~88.9 ~440 3.8 3.4

Light Phosphorus Texture

ONO Front 100 ~466 10 10*

*Measurement prior to corona charging

Page 140: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

127

Table 7.5: Measured surface passivation for an IBC solar cell using test wafers from batch 4. Front surface with ONO (n SiNx ~1.93) and rear surface with ONO (n SiNx ~2.10). Cell wafer thicknesses of ~235 µm.

Region Passivation Coverage (%)

Sheet Resistivity (Ω/)

J0 (fA/cm2)

Area Weighted J0 (fA/cm2)

Boron Planar ONO Rear ~9.5 150 20 1.9

Heavy Phosphorus Planar

ONO Rear ~1.6 15 219 3.5

Light Phosphorus Planar

ONO Rear ~88.9 ~680 4.4 3.9

Light Phosphorus Texture

ONO Front 100 ~550 11.8 11.8*

*Measurement prior to corona charging

Differences in the diffusions and sheet resistivities make it difficult to have a fair comparison

between the J0 from all batches. Comparing the experimental data of the ONO presented in

Chapters 4 and 5 to SiO2-Si3N4, the ONO passivation performed better in most cases. No

experimental data for ONO surface passivation are available for comparison in the case of

heavily doped phosphorus diffusion with resistivities lower than 60 Ω/. However, the J0 for

both ONO and SiO2-Si3N4 stacks are thought to be similar as the quality of surface passivation

for such heavily diffused regions becomes less critical. The PECVD SiNx-SiOx on an

undiffused rantex surface had a much lower J0 compared to the lightly diffused phosphorus

texture surface with the ONO stack. Therefore, a large contribution to the improvement of J0

on ONO cell batches is considered to be due to the rear surface passivation.

The ONO stack has better passivation results on undiffused (Chapter 4) and light phosphorus

diffused surfaces compared with the SiO2-Si3N4 stack, which is probably due to the differences

in the hydrogen concentration of the PECVD SiNx and LPCVD Si3N4. The higher hydrogen

concentration commonly found in PECVD SiNx may have hydrogenated the Si-SiO2 interface

[294]. Moreover, a lower J0 with the ONO surface passivation was obtained with SiNx at a

higher refractive index of ~2.10 on a boron diffusion surface due to the lower positive charges

in the SiNx layer. The measured efficiency, Voc, Jsc, FF and shunt and series resistances for

batches 1 to 4 are plotted in Figure 7.7. The batches had efficiencies ranging between 22.8–

24.8%. It is observed from Figure 7.7a that the ONO with rear SiNx n ~1.93 had a higher

efficiency on average compared to the other batches. However, the highest cell efficiency was

found in batch with the ONO having an SiNx with n632 ~2.10.

Page 141: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

128

Batch 1 uses the previous passivation scheme and has a Voc ranging between 708–716 mV. The

Jsc was measured to be much lower as compared to passivating both surface with the ONO

stack with an average of ~40.95 mA/cm2. The low Jsc was caused by incomplete texturing at

the front surface. Microscope imagery captured for one of the cells from batch 1 in Figure 7.8

shows incomplete texturing at the front surface with a high fraction of planar regions. This

coincides with the lower average Jsc observed for batch 1 as compared to both ONO batches.

The filtered image in Figure 7.8b was processed using ImageJ. The calculated white region in

the filtered image shows the planar regions covers ~12%. Simulation using OPAL2 show that

the ~12% planar region can lead to a JGen loss of 0.4–0.5 mA/cm2, assuming identical

enhancements to the path length. The large distribution for Voc can be explained by the non-

uniform surface passivation from the PECVD SiNx on the undiffused front surface. As

discussed in Section 6.3.1, the surface recombination is much more sensitive to the passivation

quality without a diffusion.

Fig 7.8: (a) Microscope image was taken with one of the cells from batch 1. The planarized region is presented as bright blue areas. (b) Processed images used to measure the planarised region were created using ImageJ.

The performance of cells in batch 2 are similar to those of batch 1. Batch 2 uses the previous

passivation scheme and has a large range for Voc from 698–716 mV, whereas Jsc had an average

of ~41.23 mA/cm2. The low Jsc is due to the high planar region (similar issue to batch 1). The

lower fill factor observed as compared to Batch 1 may be due to the low bulk lifetimes of 6–9

ms as measured from the photoconductance at the cell region. The bulk lifetime is thought to

have degraded after the high-temperature processing, as other similar wafers prior to multiple

high-temperature processing achieved bulk lifetimes that were on average ~30 ms.

(a) (b)

Page 142: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

129

As previously stated, batch 3 had the highest average cell efficiency among all the batches,

however, batch 4 contained a single cell with the highest overall cell efficiency. Batch 3 had a

Voc range of 710–714 mV with an average Jsc of ~41.98 mA/cm2. Batch 4 had a higher Voc

range of 715–720 mV with an average Jsc of ~42.19 mA/cm2. Comparing both ONO batches

shows that batch 4 had both a higher Voc and Jsc compared to the cells in batch 3. This was

mainly due to the higher bulk lifetimes and thinner cell wafers for batch 4 as well as the

planarised region on the textured surface covering ~5% in batch 3. The average bulk lifetime

measured across the 6 wafers in batches 3 and 4 were 36 and 46 ms, respectively. The lower

efficiency of most cells in batch 4 was largely due to the poor fill factor with an average of

~78%. The low fill factor was due to non-ideal recombination, which is described later in this

section.

The series and shunt resistances given in Figures 7.7e-f were obtained using the dark-light I-V

method discussed in Chapter 2. Relatively similar series resistances were obtained from all four

batches using conventional metallisation of Al. However, the shunt resistances were much

lower for batch 4.

All four batches of cells were further analysed using the dark I-V measurements, which

provides the local ideality factor at Vmpp. The local ideality factor (m) is defined as the slope

of the ln(I)-V curve described in Eq 7.1. The different recombination types that contribute to

the m make it difficult to associate its value with a single specific recombination mechanism.

However, plotting an m-V curve allows the detection of small changes within the I-V curve to

indicate resistivity effects or non-ideal recombination within the operation of the cell. Batches

1, 2 and 3 had an ideal m of <1.4 at the maximum power point (Mpp). The cells in batch 4 had

a wide range of m from 1.7–3.0, which correlates to the lower fill factor observed within the

batch from Figure 7.7d.

K = 1nV Y (n((ln )\ (7.1)

The best performing cells within each batch were modelled using an equivalent circuit model

with either two diodes or resistance-limited enhanced recombination [293]. The recombination

ideality factors are normally associated with a specific recombination mechanism. The values

presented in Table 7.5 show the fit to the dark I-V curve for the best performing cell from each

batch.

Page 143: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

130

Batch 1 Batch 2 Batch 3 Batch 4

695

700

705

710

715

720

725

(b)

VO

C (

mV

)

Batch 1 Batch 2 Batch 3 Batch 440.4

40.8

41.2

41.6

42.0

42.4

42.8

43.2

JS

C (

mA

/cm

2)

(c) (d)

(a)

(e) (f)

Batch 1 Batch 2 Batch 3 Batch 474

76

78

80

82

84

FF

(%)

Batch 1 Batch 2 Batch 3 Batch 4

23

24

25E

ffic

ien

cy (

%)

Batch 1 Batch 2 Batch 3 Batch 4

103

104

105

106

107

Sh

unt

Resis

tan

ce

.cm

2)

Batch 1 Batch 2 Batch 3 Batch 4

0.20

0.24

0.28

0.32

0.36

0.40

Se

ries R

esis

tan

ce

.cm

2)

Fig 7.7: The performance of four batches of IBC solar cells. Batch 1 used a previous passivation scheme of rear SiO2-Si3N4 and front SiNx-SiOx [14]. Batch 2 had a similar passivation scheme as batch 1 with rear phosphorus diffusion. Batch 3 had an ONO cell with a SiNx n ~1.93 and is passivated both at the front and rear surfaces of the cell. Batch 4 had an ONO cell with SiNx n ~1.93 on the front and n ~2.10 at the rear. The (a) open circuit voltage, (b) short circuit current, (c) fill factor, (d) efficiency, (e) shunt resistance and (f) series resistance were measured at 25 ± 1 ˚C using a combination of dark and light I-V curves.

Front SiNx-SiOx

Rear SiO2-Si3N4

Front Rear ONO

Page 144: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

131

There are a few other reasons the cells may exhibit non-ideal recombination and low fill factors,

including pin holes in the ONO dielectric stack, metal spiking through the p-n junction,

metallisation on the diffusion overlap region, a contiguous boron and phosphorus region,

change in injection dependence of wafer surface, etc. [189, 293, 295]. Depending on the type

of recombination, the equivalent circuit with a second diode associated with either the SRH or

Auger recombination at high injection levels exhibited m2 values within the range of 1–2 [293,

295].

No clear conclusions from Table 7.5 could be drawn on which specific loss mechanism may

be involved in the non-ideal recombination and low fill factors, as this would require a more

detailed experimental design. Three changes to the cell fabrication process may overcome the

non-ideal recombination and improve the fill factor and m, which are discussed in the following

sections. The ONO IBC solar cell has a very light and shallow boron diffusion of ~150 Ω/. A

simulation performed by Edler et al. showed that increasing the depth of metal spiking through

the p-n junction drastically increases the metal recombination [296]. Moreover, metallisation

with Al is known to form an alloy at temperatures from 200˚C to the eutectic point of 577 ˚C

[297]. The annealing of Al onto silicon may cause its dissolution and create voids within the

crystal to allow Al to precipitate and form Al spikes into the silicon bulk. Therefore, having a

heavier and deeper boron diffusion junction depth may prevent metal spiking through the p-n

junction. This may additionally reduce the metal recombination by providing a ‘field-effect’

on the minority carriers that reach the metal contacts [293, 295, 296].

0 200 400 600 80010-8

10-6

10-4

10-2

100

102

104

Batch 1 Batch 2 Batch 3 Batch 4

0.5

1.0

1.5

2.0

2.5

3.0

3.5

(a)

Idealit

y F

acto

r (m

)

Batch 1 Mpp

~620 mV

Batch 2 Mpp

~618 mV

Batch 3 Mpp

~624 mV

Batch 4 Mpp

~628 mV

Mpp

Batch 1

Batch 2

Batch 3

Batch 4

Simulation

Curr

ent

Density (

mA

/cm

2)

Voltage (mV)

(b)

Fig 7.9: The measured (a) local ideality factor and (b) dark I-V curve obtained from the best performing cells within the three batches without any corona charging.

Front SiNx-SiOx

Rear SiO2-Si3N4

Front Rear ONO

Page 145: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

132

Table 7.5: Parameters used to fit the dark I-V curve with the two diode model and resistance-limited enhanced recombination [293]

The adjacent boron-doped dots and light phosphorus BSF region can lead to tunnelling

recombination, which causes current leakage and a non-ideality factor [298, 299]. Isolating the

diffusions with additional photolithography steps may eliminate such recombinations. In

addition, metallisation with different metal doped silicon approaches or stacks may prevent

spiking, such as Al doped silicon or TiN as diffusion barriers [300].

7.3.3 Positive corona charging on the front surface of ONO IBC solar cells

Positive corona charging reduces the surface recombination on undiffused and light phosphorus

rantex surfaces, as presented in Chapters 4 and 5. This implementation can potentially increase

the efficiency of ONO IBC solar cells. In this section, ONO IBC solar cells were measured

before and after positive corona charging on the front surfaces and were further annealed to

render the charge stable.

Experimental details

Two fabricated ONO IBC solar cells were positively corona charged at the front rantex surface

for 100 seconds at 5 kV. The corona charged cells were subsequently annealed at either 250 ˚C

Conditions Batch 1 Batch 2 Batch 3 Batch 4

Saturation Current J0 32 fA/cm2 32 fA/cm2 42 fA/cm2 35 fA/cm2

Ideality factor m2 2 2 - -

Saturation Current J02 5.7 nA/cm2 8.4 nA/cm2 - -

Ideality factor mH - - 1.38 1.55

Saturation Current J0H - - 1000 pA/cm2 4300 pA/cm2

Resistance H - - 1000 Ω.cm2 180 Ω.cm2

Dark I-V Shunt Resistance 55 kΩ.cm2 19 kΩ.cm2 50 MΩ.cm2 9.0 kΩ.cm2

Dark I-V Series Resistance 0.06 Ω.cm2 0.1 Ω.cm2 0.1 Ω.cm2 0.1 Ω.cm2

Local Ideality Factor at Vmpp 1.20 1.20 1.32 1.76

R2 0.9982 0.9994 0.9986 0.9973

Page 146: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

133

or 400 ˚C with forming gas for ten minutes to embed the charges. Dark and light I-V

measurements were performed on the cells before and after the corona charge and FGA.

Results and discussion

Sintering ONO IBC solar cells at 400 ˚C - The dark and light I-V measurement of the first

ONO IBC cell before and after corona charging are provided in Table 7.6. Corona charging on

the front surface of the IBC cell improved the efficiency from 23.6 to 24.2%. The increased

efficiency was mainly due to the improvement in the Voc from 720 mV to 723 mV, and fill

factor from 78.2 to 79.6%. The dark I-V measurement performed on the cell in Figure 7.10a

reveal that the improvements occurred mostly at lower voltages after corona charging due to

the accumulation of majority carriers on the front surface. This is similarly reflected with the

corona charged ONO high resistivity rantex wafers with improvement to the lifetime, which

were much more significant at lower injection levels, as described in Section 4.5. However,

further annealing at 400 ˚C in ambient forming gas, unfortunately, degraded the cell. The dark

I-V measurements revealed a much higher shunt resistance after annealing. However, the

efficiency of the cell drastically decreased, with both the Voc and the fill factor, which declined

to 683 mV and 74%, respectively. The increased shunt resistance indicates that the

improvements are observed at a lower ∆n injection level. Nevertheless, the drop in efficiency

is due to an increase in the non-ideal recombination that occurs at higher voltages, higher J0,

and higher series resistances, as presented in Table 7.6.

0 100 200 300 400 500 600 700 800

10-4

10-3

10-2

10-1

100

101

102

103

0 100 200 300 400 500 600 700 800

0

2

4

6

8

10

12

14

Curr

ent

(mA

/cm

2)

Voltage (mV)

Mpp

Before CC

After CC

After CC & Annealing

Simulation

(a) (b)

Local Id

ea

lity F

acto

r

Voltage (mV)

Fig 7.10: The measured (a) dark I-V and (b) local ideality factor for the ONO IBC solar cell before and after corona charging and annealing at 400 ˚C in FGA for ten minutes.

Page 147: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

134

Table 7.6: Dark and light I-V measurements before and after corona charging, and with further annealing at 400 ˚C to embed charges into the ONO IBC solar cell.

Sintering ONO IBC solar cells at 250 ˚C - The dark and light I-V measurement results for

the second ONO IBC cell before and after corona charging are provided in Table 7.7. The

efficiency after corona charging for the second cell increased from 24.3% to 24.5%. The

improvement was mainly due to an increased Voc from 711 to 714 mV and a fill factor that

improved from 81.5% to 81.9%. Unfortunately, dark I-V measurements were not recorded after

corona charging.

Instead of annealing at 400 ˚C, the cell was subsequently annealed at 250 ˚C for ten minutes.

The additional low-temperature annealing did not degrade the Voc with an extracted dark J0 of

35 fA/cm2. Moreover, the cell efficiency improved further and increased to 24.7% due to the

higher shunt resistance, higher fill factor and lower series resistance, as shown in Table 7.7 and

Figure 7.11. This suggests that the lower annealing temperature of 250 ˚C is suitable for our

cell design without causing additional increases in the non-ideal recombination at the Mpp. It is

noted that the non-ideal recombination, which was once hidden for lower shunt resistances,

was revealed after corona charging and annealing.

Dark I-V Conditions Before corona After corona After corona and annealing

Saturation current J0 26 fA/cm2 24 fA/cm2 86 fA/cm2

Ideality factor m2 2.0 2.0 2.0

Saturation current J02 6.22 nA/cm2 6.77 nA/cm2 36 nA/cm2

Ideality factor mH - - 1.8

Saturation current J0H - - 82 nA/cm2

Resistance H - - 180 Ω.cm2

Shunt resistance 1 kΩ.cm2 3.5 kΩ.cm2 7 kΩ.cm2

Series resistance 0.3 Ω.cm2 0.3 Ω.cm2 0.35 Ω.cm2

Local ideality Factor at Mpp 1.97 1.82 2.17

R2 0.9992 0.9965 0.9997

Light I-V Conditions

Voc 720.4 mV 723.4 mV 683.34 mV

Jsc 41.95 mA/cm2 42.05 mA/cm2 41.78 mV

FF 78.15% 79.55% 73.86%

Eff 23.6% 24.2% 21.1%

Light and Dark Rs 0.415 0.396 0.603

Page 148: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

135

0 100 200 300 400 500 600 700 800

10-6

10-5

10-4

10-3

10-2

10-1

100

101

102

103

0 100 200 300 400 500 600 700 800

0

2

4

6

8

10

12

14C

urr

ent

(mA

/cm

2)

Voltage (mV)

Mpp

Before CC

After CC & Annealing

Simulation

(a) (b)

Local Id

ea

lity F

acto

r

Voltage (mV)

Fig 7.11: The measured (a) dark I-V and (b) local ideality factor of the ONO IBC solar cell before and after corona charging with an additional anneal at 250 ˚C in FGA for ten minutes.

Table 7.7: Dark and light I-V measurement before and after corona charging and with further annealing at 250 ˚C to embed charges into the ONO IBC solar cell.

Dark I-V Conditions Before corona After corona After corona and annealing

Saturation current J0 38 fA/cm2 - 35 fA/cm2

Ideality factor m2 2.0 - 2.0

Saturation current J02 6.55 nA/cm2 - 6 nA/cm2

Ideality factor mH - - 1.4

Saturation current J0H - - 1.6 nA/cm2

Resistance H - - 1500 Ω.cm2

Shunt resistance 2 kΩ.cm2 - 0.4 MΩ.cm2

Series resistance 0.08 Ω.cm2 - 0.05 Ω.cm2

Local ideality factor at Mpp 1.27 - 1.14

R2 0.9992 - 0.9941

Light I-V Conditions

Voc 711.47 mV 713.62 mV 714.77 mV

Jsc 42.0 mA/cm2 42.0 mA/cm2 42.0 mA/cm2

FF 81.46% 81.88% 82.10%

Eff 24.34% 24.54% 24.65%

Light and Dark Rs 0.231 - 0.189

Page 149: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

136

Initial CC FGA 250 °C21

22

24

25

Initial CC FGA 250 °C40

41

42

Initial FGA 250 °C Initial FGA 250 °C0.8

0.9

1.0

1.1

1.2

1.3

2.0

2.1

2.2

21

22

24

25E

ffic

iency (

%)

Initial CC FGA 250 °C 708

710

712

714

716

(d)

(a)

Voc (

mV

)

(b)

40

41

42

Jsc (

mA

/cm

2)

Initial CC FGA 250 °C 73

77

78

79

80

81

82

83(c)

FF

(%

)

0.0

0.2

0.4

0.6

0.8

1.0

1.6

1.8

2.0(e) (f)

Fig 7.12: ONO IBC solar cells were measured before and after positive corona charging with further annealing in forming gas at 250 ˚C. The (a) efficiency, (b) open circuit voltage, (c) short circuit current, (d) fill factor, (e) series resistance and (f) local ideality factor were measured at 24 ± 1 ˚C.

Page 150: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

137

A batch of cells was fabricated, corona charged and annealed at 250 ˚C. The measured

efficiency, Voc, fill factor and series and shunt resistances are plotted in Figure 7.12. All I-V

curves in this batch had their Jsc corrected due to fluctuations in the light intensity from an

outdated halogen light source. The Jsc for a calibrated reference cell was simultaneously

monitored during the IBC solar cell measurements. However, even after correction, there is

still uncertainty within the corrected Jsc due to the poor light uniformity across the measurement

area. The voltages for all cells improved after the corona charging and further annealing.

However, the efficiency only improved for some of the cells. The improvements in the fill

factors for cells 38A, 38B, 38C and 38D were observed due to the lower series resistance and

increased Voc. The degradations for cells 38E and 38F were due to sudden increases in the

series resistance (from ~0.2 to 0.8–1.0 Ω.cm2), as caused by the metallised fingers on the cells

peeling off after the subsequent annealing in the forming gas. The poor metal and dielectric

adhesions were postulated to be due to residue from the acetone rinses before the metal

evaporation.

Page 151: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

138

7.3.4 Lift-off metallisation on ONO IBC solar cells

The formation of Al contacts on a cell have occasionally been imperfect, which causes the cell

to suffer from a high contact resistivity, low fill factor and reduced efficiency. The lift-off

metallisation was developed to lower the contact resistivity during cell processing and reduce

the potential for metal spiking by using different metallisation stacks [301].

The second metallisation method involved a lift-off process and is as shown in Figure 7.13b.

The potential advantage of the lift-off process is the avoidance of both native oxide growth and

organic contaminants between the evaporated metal and silicon dot contacts. Conventional

metal evaporation has metal contact dots that are exposed to acetone and photoresist prior to

metal evaporation. The ‘lift-off’ process avoids such exposures and the metal evaporation can

proceed immediately after the BHF. However, for the lift-off process to be successful, the

outermost layer of the metal stack deposited prior to the photoresist lift-off must either be stable

at room temperature (non-oxidising) or must form a low resistivity conductive oxide.

A thin metal stack is evaporated onto the open metal dot contacts with the patterned photoresist

that adheres to the dielectric layer. The wafers were then submerged in an ultrasonic bath to

remove the evaporated metal together with underlying photoresist while retaining the metal

deposited on the silicon wafers. Subsequently, Al was evaporated over the entire surface,

stacking onto and connecting the individual metal dots.

In this section, the transfer length measurement (TLM) structures were fabricated on boron

diffused wafers to measure the different contact resistivities with various metallisation stacks.

In addition, dark and light I-V measurements were performed on the IBC solar cells with a

selection of metallisation stacks developed using these lift-off process and were compared to

conventional metallisation.

Experimental details

To fabricate the TLM structures, high resistivity 100 Ω.cm n-type silicon wafers were boron

diffused at 910 ˚C for 40 minutes to achieve a sheet resistivity of ~210 Ω/. The BSG was then

removed and the wafer was oxidised at 1000 ˚C and annealed in N2 for 45 minutes.

Photolithography was used to define the TLM patterns for the lift-off metallisation. Seven

metal stacks, which include chromium-palladium-silver (CrPdAg), aluminium-silver-

aluminium (AlAgAl), aluminium-palladium-aluminium (AlPdAl), palladium-aluminium

Page 152: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

139

(PdAl), aluminium-palladium (AlPd), silver-aluminium (AgAl) and chromium-aluminium

(CrAl) were selected for metal evaporation. The metallisation was performed as described in

Section 7.2.8 with a boat/rod power supplied at 550–800 W onto the patterned photoresist. The

total thickness of each stack on the TLM pads were ~165 nm. The metal evaporated onto the

exposed silicon and AZ4562 photoresist was subjected to an ultrasonic bath in acetone to lift-

off the undesired metallised area. Two TLM pads were fabricated for each metal stack. The

TLM structures were cut with a dicing saw, sintered in forming gas at 250 ˚C for 20 minutes

and measured with a Keithly source meter ranging from -0.3 to 0.3 V.

Two batches of cells were also fabricated using lift-off metallisation of chromium-palladium-

silver-aluminium (CrPdAg-Al) and aluminium-palladium-aluminium (AlPd-Al). The selected

metallisation stack for the cells were based on low contact resistivity for AlPd-Al and

frequently used metallisation stack at the ANU with CrPdAg. The cell fabrication process is

similar to the baseline; however, both metal stacks were performed using lift-off metallisation.

For example, the CrPdAg-Al stack used a Cr rod that was loaded onto the first boat followed

by Pd and Ag pellets into the second and third tungsten boats, respectively. The metal purity

selected was in the range of 3–5N. The wafers were placed one metre above the metal source

to avoiding unnecessary heating from the evaporation. As the chamber was pumped down to

7.5 × 10-5 Torr, a supplied power of 600–800W was applied across the metal boat/rod. An

initial pre-melt of the metal was performed with the shutter closed. After the metal on the boat

melted over a defined period, the shutters were opened with a deposition rate of ~6 Å/sec. This

was similarly performed for the other two metals, Pd and Ag. The chamber was vented in N2

gas after reaching the targeted thicknesses of Cr ~15 nm, Pd ~15 nm and Ag ~40 nm.

The wafers for the first stack of metal evaporation (CrPdAg) with the thin AZ1518 photoresist

were submerged in acetone and subjected to an ultrasonic bath for 10–15 minutes. After

visually inspecting that all of the photoresist was removed, the wafers were rinsed with DI H2O

and fresh acetone once more. The wafers were spin rinsed, dried and loaded into the chamber

for the thermal evaporation of ~3 µm of Al as previously described.

For both metallisation methods, the photoresist was then spun onto the front and rear of the

wafers for metal separation after the evaporation of the blanket Al layer. After exposing the

gaps between the p and n fingers, the wafers were submerged into an Al etch solution, which

consisted of H3PO4, H2O and HNO3 with a volume ratio of 20:4:1 at 40–50 ˚C. The separation

Page 153: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

140

of the metal was checked using a microscope and PL imaging to ensure the cells were not

shunted by fine metal bridges that connected the p and n fingers.

Similarly, for AlPd-Al, lift-off metallisation was performed on the AlPd stack followed by a

full blanket Al metal evaporation. The evaporation of the Al and Pd stack had thicknesses of

~50 nm and ~20 nm, respectively, followed by the final Al evaporation of ~3 µm. The I-V

measurement for both the lift-off metallisation cell results was compared to cells from batch 4.

Fig 7.13: Metallisation process for the ONO IBC solar cell: (a) conventional and (b) lift-off process

(a) Conventional process (b) Lift-off process

ONO Dielectric Stack

Photoresist

Lift-Off Metal

Aluminium

Silicon

Photoresist

Develop

RIE - BHF

Page 154: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

141

Results and discussion

The average contact resistivity from TLM structures prepared using the lift-off processes is

plotted in Figure 7.14a. The TLM measurements obtained with AlPdAl showed the lowest

contact resistivity of 1.55 × 10-6 Ω.cm2, whereas the Al and CrPdAg contact resistivity were

measured to be 1.24 × 10-4 Ω.cm2 and 9.4 × 10-5 Ω.cm2, respectively. The electrochemical

capacitance-voltage (ECV) measurements performed on the boron diffused wafers had a

surface concentration of ~9 × 1018 cm-3 with sheet resistivity of 210 Ω/. The measured contact

resistivities for all the metal stacks were considerably lower as compared to the Al with the lift-

off processing.

50 100 150 200 250 30010

-9

10-8

10-7

10-6

10-5

10-4

10-3

10-2

0 100 200 300 400 500

0

5

10

15

20

Al (Franklin et al.)

Cr-Pd-Ag

Pd-Al

Al-Pd-Al

Al-Ag-Al

Cr-Al

Al

Al-Pd

Al-Ag

Co

nta

ct

Re

sis

tivity (

Ω.c

m2)

Sheet Resistivity (Ω/sq)

(a)

y = 0.0371x + 0.083

R2 = 0.9999

Rsheet ~223 Ω/

(b)

y = 0.0369x + 0.511

R2 = 0.9999

Rsheet ~215 Ω/

Al

Al-Pd-Al

Cr-Pd-Ag

Resis

tan

ce

)

TLM spacing (µm)

y = 0.0359x + 0.517

R2 = 0.9999

Rsheet ~215 Ω/

Fig 7.14: (a) The measured contact resistivity for various metal stacks on light boron diffusion. The metallisation was performed with lift-off except for the Al performed by Franklin et al. [14]. (b) The TLM measurement for the Al, AlPdAl and CrPdAg structures.

Photoconductance measurements and PL imaging were performed on a cell after the lift-off

metallisation for CrPdAg. Photoconductance measurements on the cell showed an increased J0

of ~5 fA/cm2 whereas the PL imaging measured a lifetime drop of ~1.6 ms at 3 × 1015 cm-3

across the cell region.

The cell efficiencies, Voc, Jsc and fill factors for both the lift-off metallisations were plotted

together with conventional Al evaporation in Figure 7.15. Metallisation with AlPd-Al has the

lowest average efficiency of ~22.14% among the three batches, whereas Al and CrPdAg-Al

had average efficiencies of ~23.6% and ~23.7%, respectively. All three batches of cells were

measured to have relatively similar Voc and Jsc. Moreover, the photoconductance measurements

Page 155: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

142

performed on these cells before metallisation had a J0 of 7–10 fA/cm2. The comparatively

similar Voc observed across the batches suggests that the metal contact recombination for all

three metallisations are too small to cause any differences. The similar Jsc observed across the

batches could be justified with the relatively small difference in the overall rear reflectivity, as

the back reflector of the cell consists mostly of Al of ~99.65% and the metal stack of only

~0.35%.

A much lower fill factor of the AlPd-Al cells was observed as compared to the CrPdAg-Al and

Al cells. Unfortunately, the lower fill factor is due to a higher series resistance that ranged from

0.6–2.3 Ω.cm2. Moreover, the AlPd-Al cells have large variations in their shunt resistances.

The contributions of both the low shunt and high series resistances were the main cause for the

low fill factor, causing the low efficiency of the cells. There are a few plausible reasons for the

high series resistances caused by the AlPd-Al stack. Firstly, the evaporation of palladium had

a much slower deposition rate compared to the TLM test structures. The TLM test structures

had palladium evaporation recorded at ~6 Å/sec as compared to the cells with deposition rates

recorded at ~2 Å/sec just after the shutter was opened. The low evaporation rate suggests that

the Pd in the boat may not have pre-melted fully and that other contaminates with similar and

lower melting points as Pd may have been deposited onto the sample as well [254]. Secondly,

contamination of the metal contacts between the AlPd and Al after the lift-off processes could

be another plausible cause for the high contact resistivities. The measured metallised stacks on

the TLM structure were performed with a single run without breaking the chamber vacuum.

However, the cell structure for the lift-off metallisation required a total of two metal depositions,

exposing the metal to air and wet chemicals before evaporating the final Al layer. Moreover,

photoresist and acetone residue may have resided on the AlPd metal dots during the lift-off

process, which could contribute to increased contact resistance. Of the three metals used for

the cell fabrication process, the baseline metallisation process with Al remains the best method

to make good contact with silicon.

The cells metallised with CrPdAg-Al had a slightly higher series resistance of 0.34–0.60 Ω.cm2

compared to the cells with Al metallisation. Due to the number of cells fabricated for CrPdAg-

Al, it is inconclusive but plausible that CrPdAg-Al may prevent spiking through the p-n

junction due to the observed higher shunt resistance compared to the Al and AlPd-Al

metallisation stack in Figure 7.15e. However, the conventional metallisation process with Al

achieves lower contact resistivities as compared to the fabricated metallisation stacks.

Page 156: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

143

AlPd-Al Al CrPdAg-Al710

715

720

725

VO

C (

mV

)

(b)

AlPd-Al Al CrPdAg-Al40

41

42

43(c)

JS

C (

mA

/cm

2)

AlPd-Al Al CrPdAg-Al66

68

70

72

74

76

78

80

82(d)

FF

(%)

AlPd-Al Al CrPdAg-Al

19

20

21

22

23

24

25(a)

Eff

icie

ncy (

%)

AlPd-Al Al CrPdAg-Al

103

104

Sh

un

t R

esis

tan

ce

.cm

2)

(e)

AlPd-Al Al CrPdAg-Al0.0

0.5

1.0

1.5

2.0

2.5(f)

Se

rie

s R

esis

tan

ce

.cm

2)

Fig 7.15: Three batches of ONO IBC solar cells were created that are identical except for the rear metallisation on the diffusion region. The cells have metallisation stacks of aluminium-palladium-aluminium (AlPd-Al), aluminium (Al) and chromium-palladium-silver-aluminium (CrPdAg-Al). The (a) Voc, (b) Jsc, (c) fill factor and (d) efficiency were measured at 25 ± 1 ˚C. The box represents the range of data between 25 – 75% with the line showing the mean, and the whiskers represent the maximum and minimum values obtained.

Page 157: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

144

7.4 ONO IBC cell with certified conversion efficiency of 25%

The best performing ONO IBC solar cell was achieved using the batch 4 cell fabrication

process (Section 7.3.2) and the conventional Al metallisation. The champion cell fabricated at

ANU as part of this work was independently certified at the Commonwealth Scientific and

Research Organisation (CSIRO) in Newcastle using our custom designed measurement jig and

an aperture mask.

Fig 7.18: (a) Rear image of champion ONO IBC solar cell (b) Calibrated lifetime image of ONO IBC solar cell measured at carrier injection ∆n of ~2.5 × 1015 cm-3 before metallisation.

Prior to corona charging, in house measurements from batch 4 reached an efficiency of 24.8%.

However, the similar cell measured at CSIRO achieved a certified efficiency of 25.0 ± 0.6%

(report no: A0014-ANU-D02_001). The certified cell attained a Voc of 716 ± 11 mV, fill factor

of 81.0 ± 1.9% and Jsc of 43.0 ± 0.78 mA/cm2, by considering the spectral mismatch of the

current as defined under a solar spectral irradiance at AM1.5G and an aperture mask area of ~4

cm2 [302]. The discrepancy in cell efficiency measured at ANU and CSIRO was largely due to

the underestimation in Jsc.

Due to the annealing limitations at CSIRO, the cell was only corona charged on the front rantex

surface without further annealing. The cell efficiency was measured as having a similar value

of 24.95 ± 0.61% after corona charging, as indicated by the I-V curve plotted in Figure 7.17a.

Although the cell had a slight increase in the Voc of 717 ± 11 mV and fill factor of 81.1 ± 1.9%,

the measured Jsc was slightly lower at 42.9 ± 0.75 mA/cm2. Figure 7.17b shows the normalised

reflectance and external and internal quantum efficiency curves of the ONO IBC solar cell. The

(a) (b)

Page 158: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

145

measured reflectance and external quantum efficiency (EQE) show an excellent internal

quantum efficiency (IQE) with an average of 99.1% across the 600–900 nm wavelength range.

0 200 400 600 800

0

10

20

30

40

(a)

Curr

ent

(mA

/cm

2)

Voltage (mV)

Cell aperture area: 3.99 ± 0.02 cm2

JSC

: 42.9 ± 0.75 mA/cm2

VOC

: 717 ± 1.1 mV

FF: 81.1 ± 1.9 %

η: 25.0 ± 0.6 %0.0

0.5

1.0

1.5

2.0

Spectr

al in

t. (

W/m

2/n

m)

AM 1.5 G

400 600 800 10000.0

0.2

0.4

0.6

0.8

1.0(b)

Measured ONO

Simulated ONO

ONO EQE

ONO IQE

Wavelength (nm)

Reflecta

nce, E

QE

, IQ

E

Fig 7.17: Measurements of the (a) I-V curve and (b) reflectance, EQE, and IQE of the champion ONO IBC solar cell after corona charging. The independent measurements were performed at CSIRO in Newcastle.

Simulations were performed using Quokka3, OPAL2 and SunSolve on the champion cell

before corona charging to quantify the primary loss mechanism by utilising the free energy loss

analysis (FELA). The thicknesses of the ONO simulated in the SunSolve ray tracing and

OPAL2 packages show the thermal SiO2, SiNx and SiOx were approximately 17, 53 and 100

nm, respectively. A similar cell structure was simulated using SunSolve with an ONO stack on

both front and rear surfaces, a rear Al of 1000 nm and a Lambertian scattering fraction of 0.63

to achieve a Jsc of ~43.0 mA/cm2. The generation profile simulated using SunSolve was loaded

into Quokka3 together with the J0 recombination as measured from test structures (shown in

Table 7.5 - metallisation recombination J0 of 5 fA/cm2 as discussed in section 7.3.4) and series

and shunt resistances of 0.21 Ω.cm2 and 9000 Ω.cm2, respectively.

The simulated “25.0% cell” modelled before corona charging had a Voc, Jsc, fill factor and

efficiency of 719 mV, 42.98 mA/cm2, 81.38% and 25.17%, respectively, which is well within

the measurement errors. The loss analysis breakdown of the champion cell is shown in Figure

7.18.

Page 159: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

146

25.17

0.41

0.140.19

0.69

0.20

0.78

23.5

24.0

24.5

25.0

25.5

26.0

26.5

27.0

27.5

28.0

28.5C

ell

Effic

ien

cy a

nd

Effic

ien

cy L

osse

s (

%)

Resistive loss - Internal, 0.78%

Resistive loss - Contacts, 0.045%

Resistive loss - Fingers, 0.11%

Resistive loss - Shunt, 0.04%

Recombination - Contacts, 0.20%

Recombination - Bulk, 0.09%

Recombination - Front pass. light phos diff, 0.69%

Recombination - Rear pass. light phos diff, 0.19%

Recombination - Rear pass. heavy phos, 0.14%

Recombination - Rear pass. boron diff, 0.04%

Perimeter/edge losses, 0.41%

Modelled Cell Efficiency, 25.17%

Fig 7.18: Modelled cell efficiency and breakdown of losses for the champion cell. Modelling was performed using Sunsolve and Quokka3 to simulate the free energy loss analysis method.

Excluding the optical losses, adding the modelled losses in Figure 7.18 yields an efficiency of

27.8% for the 230 µm thick 100 Ω.cm n-type cell. The resistive losses accounted for

approximately 0.98%, which is mostly dominated by the bulk resistance. Recombination within

the cell accounted for approximately 1.35%, where a large portion was due to the front surface

passivation and the rear heavy phosphorus diffusion dots. The simulated edge recombination

contributed to approximately 0.41% of loss in the efficiency. The cell performance has the

potential to further improve by reducing the recombination on the front rantex surface.

Simulation of champion IBC solar cell structure with ONO stack rantex surface J0 of ~4 fA/cm2

(attainable J0 value as presented in Chapter 5) without non-ideal recombination shows

simulated efficiency of 25.6% is achievable.

Page 160: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

147

7.5 Chapter Summary

High-efficiency IBC solar cells were achieved by introducing numerous improvements over

previous cell fabrication processes.

High bulk lifetimes for the cells was maintained by having a heavy phosphorus diffusion layer

at the front surface and BRL gettering. The gettering steps were simplified as the process was

implemented during the formation of the p-n junction and diffusion for the contact dots without

requiring additional processing steps.

The ONO surface passivation has been demonstrated to provide excellent passivation on both

the boron and phosphorus diffusion regions. However, having a higher SiNx refractive index

layer of 1.93 to 2.10, as achieved in batches 3 and 4, did not provide a significant improvement

in the boron diffused surface passivation.

Corona charging at the front surface improved cell efficiency. However, annealing the cells at

400 ˚C to trap the charges decreased the cell performance due to an increase of the non-ideal

recombination. A lower annealing temperature of 250 ˚C could trap the charges into the ONO

stack while preventing non-ideal recombination.

Lift-off metallisation did not achieve a lower contact resistivity on the fabricated IBC solar

cells as compared with conventional metallisation using Al. However, the lift-off process has

the potential to achieve consistently low contact resistivities. A much more rigorous experiment

should be conducted to evaluate the potential of such a process.

By incorporating all the processing improvements, an independently measured efficiency of

25.0% was achieved on a ~4 cm2 aperture mask with a custom-designed measurement jig.

Direct metal contact to the silicon absorber and front surface recombination remains one of the

major recombination centres to the cell design. Further reducing the front rantex surface

passivation to J0 of ~4 fA/cm2 and without non-ideal recombination can further increase the

cell efficiency to 25.6%.

Page 161: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

148

Chapter 8

This thesis investigated the application and optimisation of ONO dielectric stacks onto IBC

solar cells at one-sun illumination. The focus was on improving the fabrication process to

develop high-efficiency IBC solar cells in conjunction with the optimised ONO dielectric

stacks. The specific conclusions for each chapter from this thesis are summarised below.

In Chapter 2, various types of surface passivation commonly used in silicon solar cells were

reviewed. Thermal SiO2, PECVD SiNx and ALD Al2O3 surface passivation dielectrics were

evaluated and compared to the optimised respective surface passivations achieved at ANU.

Chapter 3 presents the different measurements employed to quantify the quality of the surface

passivation, dielectric charge and the performance of the solar cell. The measured effective

lifetime and J0 on undiffused and diffused wafers were detected primarily using

photoconductance measurements. The effective charges were determined using C-V

measurements with a MOS structure. Stringent light I-V measurements were performed on

solar cells to accurately measure the performance and the extraction of the series resistance,

which were measured and compared to the light and dark I-V measurements.

In Chapter 4, the conditions presented on each dielectric layer of the ONO stack were assessed.

The main conclusions were as follows:

- Thermal SiO2 provides chemical passivation, SiNx hydrogenates the Si-SiO2 interface

and induces a field-effect with positive charges and the SiOx traps corona charges and

improves the overall stack for the ARC.

- An ONO stack with a thermal SiO2 thickness greater than 7 nm without POA provided

excellent surface passivation on undiffused planar n-type wafers.

- The SiNx deposition temperature and SiH4:NH3 gas ratio were identified as the most

sensitive parameters for surface passivation.

- Corona charging of the ONO stack reduces the surface recombination substantially for

planar 111 and rantex surfaces as compared to planar 100 surfaces.

- Further annealing of positive corona charges at 400 ˚C into the ONO stack traps the

charges and renders it stable over a two-year period.

Conclusion

Page 162: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

149

- High record lifetimes on various n-type wafers were achieved, which surpassed those

of recent parameterisation Auger lifetimes [17].

In Chapter 5, the conditions for each dielectric layer of the ONO stack for phosphorus and

boron diffused surfaces were investigated. The main conclusions were as follows:

- The ONO stack on the light phosphorus diffused wafers showed excellent surface

passivation without the need for POA. The phosphorus diffused rantex wafers required

an overall thicker thermal SiO2 compared to the planar wafers were likely due to the

concave and convex corners on its surface.

- The ONO stack on the boron diffused wafers improved with an increased POA of N2

due to the reduced positive charges within the SiO2 after annealing, lower surface

concentration and deeper diffusion profile with a prolong POA N2 drive-in. An

increased thermal SiO2 thickness improved the surface passivation due to a lower Auger

recombination within the diffusion and the charge centroid of the SiNx being further

from the surface.

- ONO passivation improved on the boron diffused surface with higher refractive indexes

n632 from 1.94–2.10 and the lower positive charges measured within the SiNx.

- Corona charging can both improve and damage the Si-SiO2 interface. The deposition

of corona charging improved with the field-effect passivation; however, ion

bombardment of charges may cause dangling bonds towards the interface.

In Chapter 6, the three common losses in a solar cell were studied, including optical,

recombination and resistance losses. Simulations were performed using Quokka3 to review the

efficiency, Voc, Jsc and fill factor. Numerous design improvements were made over the 24.4%

IBC cell design based on the findings from the simulation and modelling work [14].

In Chapter 7, the cell fabrication processes were determined based on the simulation results in

Chapter 6 while also considering the inevitable fabrication limitations. The main conclusions

of Chapter 7 are summarised below.

The improvement in the cell fabrication processes compared to previous IBC cell designs

include:

- Loading n-type 100 Ω.cm FZ wafers at high-temperature of 900 ˚C and oxidizing at

1000 ˚C was performed for one hour to permanently remove any low lifetime defects

caused by N2 doping.

Page 163: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

150

- A thinner AZ1518 photoresist offers limited protection to the ONO rantex surfaces as

the tip of the pyramids are occasionally etched after BHF. A thicker photoresist was

used to protect the surface passivation at the front rantex surface from exposure to both

BHF and RIE.

- BRL gettering was implemented during the formation of the boron dot emitter to

maintain the bulk lifetime of wafers.

- TMAH with MonoTEX additives have shown consistent random texture etching results

across multiple test runs and cell wafers.

The following conclusions were made about the fabrication of the ONO IBC solar cells:

- No significant improvements to the surface passivation were observed on the boron

diffused test wafers as the SiNx refractive index n632 increased from 1.93 to 2.10.

- Positive corona charging on the front surface improved the cell Voc by 1–3 mV.

Annealing to embed the charges at 400 ˚C increased the local ideality factor and non-

ideal recombination within the cell. The annealing temperature was lowered to 250 ˚C

to embed the charges and improve the performance of the solar cell. This additional

annealing at 250 ˚C marginally affected the local ideality factor and non-ideal

recombination.

- In our experiments, the lift-off metallisation did not provide any means of lowering the

contact resistivity for ONO IBC solar cells. Conventional metallisation using thermal

evaporated Al on an ONO IBC solar cell achieved a low Rs of ~0.2 Ω.cm2.

The optimisation of the ONO dielectric stack, including the improved design from the previous

cell structure and refined fabrication process on an IBC solar cell, achieved an independent

measure efficiency of 25.0%. The ONO dielectric stack together with current cell designs is

capable of achieving an even higher cell efficiency as simulation results demonstrated an

efficiency reaching potentially 25.6% by reducing the J0 of the front surface recombination and

without non-ideal recombination.

Page 164: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

151

Page 165: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

152

A - Reflection of symmetrical pyramids

Symmetrical pyramids were investigated by defining the limits of light ray bounces within the

pyramid. The light ray for two identical pyramids with an angle of α can be calculated as shown

in Figure A.1 together with Eq A.1 and A.2, for ω between 31˚ and 54˚. For the rays to be

reflected twice, the minimum required angle of σ would have to be above 30˚. As σ increases

to above 45˚, all the rays are reflected twice. Table A.1 shows the reflected rays for the

symmetrical pyramids and the conditions required to attain lower reflections. As the angle of

the pyramid increased, the number of reflected rays on the pyramid increased, which reduced

the overall surface reflectance.

Fig A.1: Symmetrical pyramids with two rays reflecting at an angle of ω

= sin(3ω − 90)sin(90 − ω) (A.1)

180˚ − 2ω = 3ω − 90˚ (A.2)

The two-dimensional reflection calculations overlook the polarisation effects in a three-

dimensional textured surface. Moreover, most lab scale and industrial solar cells have

randomly textured surfaces.

The simulation for three-dimensional and irregular random textured surfaces requires ray

tracing to determine the possible path and fractions of light. As a result, the reflectivity can be

simulated using the optical simulator OPAL2, which considers various possible ray paths on

the random pyramids and accounts for the polarisation of the rays [19, 303].

Appendix

ω ω

90˚ - ω

360˚ - 6ω

180˚- 2ω

3ω - 90˚

Light ray

Page 166: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

153

Table A.1: The reflection conditions and calculations for symmetrical pyramids in two-dimensions

Reflected Rays

(bounces)

Condition Fraction Reflectance Equation

1 ω≤30˚ 1 z4@40, 1:22

30˚<ω ≤ 45˚45˚<ω ≤ 54˚

(1-σ):σ1

z4@40, = (1 − )z + z=z4@40, = (1 − )z= + z=

= sin(3ω − 90)sin(90 − ω)

2:3 54˚<ω≤60˚

(1-β):β

z4@40, = (1 − ª)z= + ªz « = sin(5ω − 270)sin(270 − 3ω)

ª = « sin(3ω − 90)sin(90 − ω)

Page 167: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

154

B - Metal-Semiconductor Contact

Metal contacts towards the silicon can be described as either ohmic or rectifying contacts. An

Ohmic contact has a linear response to the voltage and current, allowing charges to flow at

negative and positive bias polarities. In contrast, a rectifying contact allows either the current

to flow mostly in one direction with some leakage, or no current in the opposite direction.

Figure B.1 shows the band diagram for metal deposited onto a clean silicon surface under ideal

conditions. The Ohmic contact on an n-type silicon wafer can be achieved by selecting a metal

with a small ϕm, which reduces the barrier height of ϕBn. However, the opposite (a large ϕm)

would be required to make good Ohmic contact to the p-type silicon.

Fig B.1: Band diagram of the (a) rectifying and (b) Ohmic metal contact on an n-type silicon wafer.

A thin layer of native oxide commonly exists during the deposition of metal onto silicon, as

shown in Figure B.2. The thin oxide causes the interface density states (Dit) to exist due to

dangling bonds between the native oxide and silicon. The relationship between the Dit on the

surface and its effect on the pinning of the metal is described by Eq 6.17.

As Dit approaches infinite, γ becomes 0 (Bradeen limit) [304]. In this condition, the observed

barrier height of the ϕBn to n-type silicon is approximately 2/3 of the bandgap, while the ϕBp is

1/3 of the bandgap to p-type silicon [305], regardless of the ϕm. Turner et al. showed the barrier

height to be independent of the metal work function on a cleaved surface, which has a large Dit

[306]. However, when Dit approaches zero, and γ becomes 1 (Schottky-Mott limit) with δ being

0, the ϕBn and ϕBp become very dependent on ϕm [306]. Tao et al. showed that Cr and Al have

Ef,m

ϕm

Ef,Si

n-type

Metal

Evac

EC

EV

(a) Rectifying contact

χ Vbi ϕBn ϕBp

ϕs E

f,m ϕm

Ef,S

n-type

Metal

Evac

EC

EV

(b) Ohmic contact

χ ϕs

Page 168: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

155

barrier heights that are closely related to the ideal conditions [307] when metallised onto an n-

type Si surface with a monolayer of selenium passivation with silicon χ at 4.29 eV [308].

ϕp = γ(ϕ2 − x) + (1 − «) Z± − ϕ;# − γm_i3 (6.17)

« = 11 + ²b34v;v/

(6.18)

QD: Charge in depleted region Ci: Capacitance of interface layer of native oxide δ: Native oxide thickness εr: Dielectric constant of native oxide

Fig B.2: Band diagram for the MOS on an n-type silicon wafer.

Another way to achieve Ohmic contacts is by contacting metal onto the heavily doped silicon.

There are three known conduction mechanisms on a diffused region: thermionic emission,

thermionic-field emission and field emission [309]. Thermionic emission occurs when the

silicon is lightly doped. The conduction processes will require the electrons to jump across the

potential barrier due to the large depletion region. The fourth characteristic for thermionic

emission is found to be similar to a rectifying contact. However, when the silicon is heavily

doped, the potential barrier is narrowed, and the electrons can tunnel through. This mode is

defined as ‘field emission’ and creates an Ohmic contact with the metal. When the silicon is

moderately doped, only a certain number of electrons can tunnel through the narrower region

of the potential barrier (such as the midsection of the potential barrier). This conduction is

Ef,m

ϕmE

f,Si

n-type Si

Metal

EC

EV

Metal Contact with interface states

χ Vi ϕBnϕ0E0

Qit Depletion

region

Page 169: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

156

defined as ‘thermionic-field emission.’ The contact resistivities of these mechanisms are

modelled using different potential barriers, as plotted in Figure B.3 [309].

1018 1019 1020 102110

-8

10-6

10-4

10-2

100

102

104 TE 0.8 φBn

TE 0.7 φBn

Co

nta

ct R

esis

tan

ce

/sq

)

Carrier Concentration (cm-3)

φBn

0.6

φBn

0.7

φBn

0.8

TE 0.6 φBn

Fig B.3: Model of emissions from Xu et al. with different potential barriers [309]. The model is generated using Al and Pt with ϕBn assumed at be 0.6–0.8 eV. The thermionic-field emission region is modelled at 1018 to 1020 cm-

3. The field emission dominates with carrier concentrations above 1020 cm-3. The thermionic emission is not dependent on the carrier concentration, which occurred at lower carrier concentrations, as shown by the dotted lines.

The various contact resistivities for the direct metal contacts and passivated contacts with TiO2

on heavy phosphorus diffusion are plotted in Figure B.4. Thermal evaporation of Al performed

in our laboratory is known to achieve the lowest contact resistivity by far. Hao et al. compared

the Ti deposition using a low-bias physical vapour deposition with the TiO2 passivated contacts

using ALD [310]. The passivated contacts with a TiO2 of 1.4 nm have shown lower contact

resistivities at low surface concentrations compared to direct Ti metallisation. This is due to

the Fermi level pinning with the surface passivation (Dit < infinite) [310]. Hao et al. found that

at lower surface concentrations (ND <4×1019 cm-3), the thermionic-field emission dominates,

with ϕBn on the Ti-TiO2-Si as measured to be ~0.14 eV compared to ~0.5 eV with Ti-Si [310].

Therefore, an Ohmic contact can be achieved either by lowering the Dit interface between the

metal and silicon by selecting an appropriate work function for both the dielectric and metal,

or by direct metal contact onto the heavily diffused regions [311].

Page 170: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

157

1018

1019

1020

10-9

10-8

10-7

10-6

10-5

10-4

10-3

10-2

10-1

Model

φBn

0.7

φBn

0.6 φ

Bn 0.5

Al - Kean et al.

Al - Schroder et al.

Al - Yu et al.

Ti - Ashish et al.

TiO2-x

/ Ti - Ashish et al.

Ti - Hao et al.

TiO2 / Ti - Hao et al.

Conta

ct

Resis

tivity (

Ω.c

m2)

Surface Carrier Concentration (cm-3)

Fig B.4: Contact resistivity of the direct metal contact and passivated contacts with TiO2 on phosphorus-doped silicon.

Page 171: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

158

Refereed journal papers:

1. Teng Choon Kho, Kean Fong, Keith McIntosh, Evan Franklin, Nicholas Grant,

Matthew Stocks, Sieu Pheng Phang, Yimao Wan, Er-Chien Wang, Kaushal Vora, Zin

Ngwe, Andrew Blakers. ‘Exceptional silicon surface passivation by an ONO dielectric

stack’, Volume 189, Pages 245–253, Solar Energy Materials and Solar Cells, 2019.

2. Kean Chern Fong, Teng Choon Kho, WenSheng Liang, Teck Kong Chong, Marco Ernst,

Daniel Walter, Matthew Stocks, Evan Franklin, Keith McIntosh, Andrew Blakers.

‘Phosphorus diffused LPCVD polysilicon passivated contacts with in-situ low pressure

oxidation’, Volume 186, Pages 236–242, Solar Energy Materials and Solar Cells, 2018.

3. T Niewelt, A Richter, TC Kho, NE Grant, RS Bonilla, B Steinhauser, J-I Polzin, F

Feldmann, M Hermle, JD Murphy, SP Phang, W Kwapil, MC Schubert. ‘Taking

monocrystalline silicon to the ultimate lifetime limit’, Volume 185, Pages 252–259,

Solar Energy Materials and Solar Cells, 2018.

4. Katherine A Collett, Ruy S Bonilla, Phillip Hamer, Gabrielle Bourret-Sicotte, Richard

Lobo, Teng Kho, Peter R Wilshaw. ‘An enhanced alneal process to produce SRV<1

cm/s in 1 Ω cm n-type Si’, Volume 173, Pages 50–58, Solar Energy Materials and Solar

Cells, 2017.

International conference papers:

5. Kean Chern Fong, Teng Choon Kho, WenSheng Liang, Teck Kong Chong, Marco

Ernst, Daniel Walter, Matthew Stocks, Evan Franklin, Keith McIntosh, Andrew

Blakers. ‘Optimization and Characterization of Phosphorus Diffused LPCVD

List of publications

Page 172: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

159

Polysilicon Passivated Contacts with Low Pressure Tunnel Oxide’, Pages 2002–2005,

IEEE 7th World Conference on Photovoltaic Energy Conversion, 2018.

Page 173: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

160

[1] H. Gloystein, "Large-scale solar power set for double-digit growth: Goldman Sachs," ed:

Reuters, 2019.

[2] G. Carr, "Sunny Uplands: Alternative energy will no longer be alternative," in The Economist,

ed, 2012.

[3] S. Philipps, "Photovoltaics Report," presented at the PSE Conferences & Consulting GmbH,

2019.

[4] I. PVPS, "2018 Snapshot of Global Photovoltaics Markets," T1-33:2018, 2018.

[5] I. PVPS, "2016 Snapshot of Global Photovoltaic Markets," T1-31:2017, 2016.

[6] I. PVPS, "2014 Snapshot of Global PV Markets," T1-26:2015, 2014.

[7] V. Henze, "Clean Energy Investment Exceeded $300 Billion Once Again in 2018," ed:

BloombergNEF, 2019.

[8] M. Schmela, G. Masson, and N. N. T. Mai, "Global Market Outlook For Solar Power / 2016 -

2020," 2016.

[9] C. Philibert, "Technology Roadmap - Solar Photovoltaic Energy 2014," 2014.

[10] W. Xiaoting and B. Allen, "The Evolving Value of Photovoltaic Module Efficiency," Applied

Sciences, vol. 9, no. 6, p. 1227, 2019.

[11] C. Hollemann, F. Haase, S. Schäfer, J. Krügener, R. Brendel, and R. Peibst, "26.1%-efficient

POLO-IBC cells: Quantification of electrical and optical loss mechanisms," Progress in

Photovoltaics: Research and Applications, 2018.

[12] D. Levi and D. Friedman, "Conversion efficiencies of best research solar cells worldwide from

1976 through 2019 for various photovoltaic technologies. Efficiencies determined by certified

agencies/laboratories.," ed: National Renewable Energy Laboratory (NREL), 2019.

[13] K. Yoshikawa et al., "Silicon heterojunction solar cell with interdigitated back contacts for a

photoconversion efficiency over 26%," Nature Energy, Article vol. 2, p. 17032, 2017.

[14] E. Franklin et al., "Design, fabrication and characterisation of a 24.4% efficient interdigitated

back contact solar cell," Progress in Photovoltaics: research and applications, vol. 24, no. 4,

2014.

[15] K. Masuko et al., "Achievement of More Than 25% Conversion Efficiency With Crystalline

Silicon Heterojunction Solar Cell," IEEE Journal of Photovoltaics, vol. 4, no. 6, pp. 1433-1435,

2014.

[16] A. Fell, J. Schön, M. C. Schubert, and S. W. Glunz, "The concept of skins for silicon solar cell

modeling," Solar Energy Materials and Solar Cells, vol. 173, pp. 128-133, 2017.

[17] A. Richter, S. W. Glunz, F. Werner, J. Schmidt, and A. Cuevas, "Improved quantitative

description of Auger recombination in crystalline silicon," Physical Review B, vol. 86, no. 16, p.

165202, 2012.

[18] K. R. M. a. P. P. Altermatt, "A freeware 1D emitter model for silicon solar cells," presented at

the 35th IEEE Photovoltaic Specialists Conference, Honolulu, 20-25 June, 2010.

[19] K. R. McIntosh and S. C. Baker-Finch, "OPAL 2: Rapid optical simulation of silicon solar cells,"

presented at the Photovoltaic Specialists Conference (PVSC), 2012.

[20] J. G. Fossum, "Physical operation of back-surface-field silicon solar cells," IEEE Transactions on

Electron Devices, vol. 24, no. 4, pp. 322-325, 1977.

[21] R. S. Bonilla, B. Hoex, P. Hamer, and P. R. Wilshaw, "Dielectric surface passivation for silicon

solar cells: A review," physica status solidi (a), vol. 214, no. 7, p. 1700293, 2017.

[22] A. M. M., T. E., and S. E. J., "Stabilization of Silicon Surfaces by Thermally Grown Oxides*," Bell

System Technical Journal, vol. 38, no. 3, pp. 749-783, 1959.

Bibliography

Page 174: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

161

[23] A. W. Blakers and M. A. Green, "20% efficiency silicon solar cells," Applied Physics Letters, vol.

48, no. 3, pp. 215-217, 1986.

[24] J. G. Fossum and E. L. Burgess, "High‐efficiency p+‐n‐n+ back‐surface‐field silicon solar cells,"

Applied Physics Letters, vol. 33, no. 3, pp. 238-240, 1978.

[25] C. J. Frosch and L. Derick, "Surface Protection and Selective Masking during Diffusion in

Silicon," Journal of The Electrochemical Society, vol. 104, no. 9, pp. 547-552, September 1,

1957 1957.

[26] E. A. Lewis and E. A. Irene, "The Effect of Surface Orientation on Silicon Oxidation Kinetics,"

Journal of The Electrochemical Society, vol. 134, no. 9, pp. 2332-2339, 1987.

[27] C.-H. Lin, "Oxidation (of Silicon)," in Encyclopedia of Microfluidics and Nanofluidics, D. Li, Ed.

Boston, MA: Springer US, 2013, pp. 1-11.

[28] B. E. Deal, "The Oxidation of Silicon in Dry Oxygen, Wet Oxygen, and Steam," Journal of The

Electrochemical Society, vol. 110, no. 6, pp. 527-533, 1963.

[29] C. Y. Chen, M. J. Jeng, and J. G. Hwu, "Rapid thermal postoxidation anneal engineering in thin

gate oxides with Al gates," IEEE Transactions on Electron Devices, vol. 45, no. 1, pp. 247-253,

1998.

[30] R. R. Razouk and B. E. Deal, "Dependence of Interface State Density on Silicon Thermal

Oxidation Process Variables," Journal of The Electrochemical Society, vol. 126, no. 9, pp. 1573-

1581, 1979.

[31] N. F. Mott, S. Rigo, F. Rochet, and A. M. Stoneham, "Oxidation of silicon," Philosophical

Magazine B, vol. 60, no. 2, pp. 189-212, 1989.

[32] W. Weber and M. Brox, "Physical Properties of SiO2 and Its Interface to Silicon in

Microelectronic Applications," MRS Bulletin, vol. 18, no. 12, pp. 36-42, 1993.

[33] I. Takahashi, T. Shimura, and J. Harada, "X-ray diffraction evidence for epitaxial

microcrystallinity in thermally oxidized SiO 2 thin films on the Si(001) surface," Journal of

Physics: Condensed Matter, vol. 5, no. 36, p. 6525, 1993.

[34] A. Munkholm, S. Brennan, F. Comin, and L. Ortega, "Observation of a Distributed Epitaxial

Oxide in Thermally Grown SiO2 on Si(001)," Physical Review Letters, vol. 75, no. 23, pp. 4254-

4257, 1995.

[35] A. Munkholm and S. Brennan, "Ordering in Thermally Oxidized Silicon," Physical Review Letters,

vol. 93, no. 3, p. 036106, 2004.

[36] P. M. Lenahan and J. F. C. Jr., "What can electron paramagnetic resonance tell us about the

Si/SiO2 system?," Journal of Vacuum Science & Technology B: Microelectronics and

Nanometer Structures Processing, Measurement, and Phenomena, vol. 16, no. 4, pp. 2134-

2153, 1998.

[37] W. Futako, N. Mizuochi, and S. Yamasaki, "In situ ESR Observation of Interface Dangling Bond

Formation Processes During Ultrathin SiO2 Growth On Si(111)," Physical Review Letters, vol.

92, no. 10, p. 105505, 2004.

[38] N. Yoshio, "Study of Silicon-Silicon Dioxide Structure by Electron Spin Resonance I," Japanese

Journal of Applied Physics, vol. 10, no. 1, p. 52, 1971.

[39] N. Yoshio, T. Kiyoshi, and O. Atsushi, "Study of Silicon-Silicon Dioxide Structure by Electron

Spin Resonance II," Japanese Journal of Applied Physics, vol. 11, no. 1, p. 85, 1972.

[40] T. Dittrich, Materials Concepts for Solar Cells. World Scientific Publishing Company, 2014.

[41] E. H. Poindexter, P. J. Caplan, B. E. Deal, and R. R. Razouk, "Interface states and electron spin

resonance centers in thermally oxidized (111) and (100) silicon wafers," Journal of Applied

Physics, vol. 52, no. 2, pp. 879-884, 1981.

[42] A. Stesmans and V. V. Afanas’ev, "Electron spin resonance features of interface defects in

thermal (100)Si/SiO2," Journal of Applied Physics, vol. 83, no. 5, pp. 2449-2457, 1998.

[43] J. P. Campbell and P. M. Lenahan, "Density of states of Pb1 Si/SiO2 interface trap centers,"

Applied Physics Letters, vol. 80, no. 11, pp. 1945-1947, 2002.

Page 175: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

162

[44] N. H. Thoan, K. Keunen, V. V. Afanas’ev, and A. Stesmans, "Interface state energy distribution

and Pb defects at Si(110)/SiO2 interfaces: Comparison to (111) and (100) silicon orientations,"

Journal of Applied Physics, vol. 109, no. 1, p. 013710, 2011.

[45] L. M. Terman, "An investigation of surface states at a silicon/silicon oxide interface employing

metal-oxide-silicon diodes," Solid-State Electronics, vol. 5, no. 5, pp. 285-299, 1962.

[46] R. E. Mikawa and P. M. Lenahan, "Structural damage at the Si/SiO2 interface resulting from

electron injection in metal‐oxide‐semiconductor devices," Applied Physics Letters, vol. 46, no.

6, pp. 550-552, 1985.

[47] Y. Y. Kim and P. M. Lenahan, "Electron ‐ spin ‐ resonance study of radiation ‐ induced

paramagnetic defects in oxides grown on (100) silicon substrates," Journal of Applied Physics,

vol. 64, no. 7, pp. 3551-3557, 1988.

[48] E. H. Poindexter, G. J. Gerardi, M. E. Rueckel, P. J. Caplan, N. M. Johnson, and D. K. Biegelsen,

"Electronic traps and Pb centers at the Si/SiO2 interface: Band‐gap energy distribution,"

Journal of Applied Physics, vol. 56, no. 10, pp. 2844-2849, 1984.

[49] P. V. Gray and D. M. Brown, "DENSITY OF SiO2–Si INTERFACE STATES," Applied Physics Letters,

vol. 8, no. 2, pp. 31-33, 1966.

[50] S. K. Lai, D. R. Young, J. A. Calise, and F. J. Feigl, "Reduction of electron trapping in silicon

dioxide by high‐temperature nitrogen anneal," Journal of Applied Physics, vol. 52, no. 9, pp.

5691-5695, 1981.

[51] J. C. King and C. Hu, "Effect of low and high temperature anneal on process-induced damage

of gate oxide," IEEE electron device letters, vol. 15, no. 11, pp. 475-476, 1994.

[52] A. Stesmans, "Passivation of Pb0 and Pb1 interface defects in thermal (100) Si/SiO2 with

molecular hydrogen," Applied Physics Letters, vol. 68, no. 15, pp. 2076-2078, 1996.

[53] Y. C. Cheng, "Electronic states at the silicon-silicon dioxide interface," Progress in Surface

Science, vol. 8, no. 5, pp. 181-218, 1977.

[54] V. V. Afanas’ev and A. Stesmans, "Positive charging of thermal SiO2/(100)Si interface by

hydrogen annealing," Applied Physics Letters, vol. 72, no. 1, pp. 79-81, 1998.

[55] B. E. Deal, M. Sklar, A. S. Grove, and E. H. Snow, "Characteristics of the Surface‐State Charge

(Qss) of Thermally Oxidized Silicon," Journal of The Electrochemical Society, vol. 114, no. 3, pp.

266-274, 1967.

[56] B. E. Deal, "The Current Understanding of Charges in the Thermally Oxidized Silicon Structure,"

Journal of The Electrochemical Society, vol. 121, no. 6, pp. 198C-205C, 1974.

[57] D. R. Lamb and F. R. Badcock, "The effect of ambient, temperature and cooling rate, on the

surface charge at the silicon/ silicon dioxide interface†," International Journal of Electronics,

vol. 24, no. 1, pp. 11-16, 1968.

[58] B. E. Deal and A. S. Grove, "General Relationship for the Thermal Oxidation of Silicon," Journal

of Applied Physics, vol. 36, no. 12, pp. 3770-3778, 1965.

[59] C. J. Han and C. R. Helms, "Parallel Oxidation Mechanism for Si Oxidation in Dry O2," Journal

of The Electrochemical Society, vol. 134, no. 5, pp. 1297-1302, 1987.

[60] E. H. Nicollian and A. Reisman, "A new model for the thermal oxidation kinetics of silicon,"

Journal of Electronic Materials, journal article vol. 17, no. 4, pp. 263-272, 1988.

[61] H. Z. Massoud, J. D. Plummer, and E. A. Irene, "Thermal Oxidation of Silicon in Dry Oxygen:

Accurate Determination of the Kinetic Rate Constants," Journal of The Electrochemical Society,

vol. 132, no. 7, pp. 1745-1753, 1985.

[62] H. R. Philipp, "Silicon Dioxide (SiO2) (Glass) A2 - PALIK, EDWARD D," Handbook of Optical

Constants of Solids, pp. 749-763, 1985.

[63] R. Kitamura, L. Pilon, and M. Jonasz, "Optical constants of silica glass from extreme ultraviolet

to far infrared at near room temperature," Applied Optics, vol. 46, no. 33, pp. 8118-8133, 2007.

Page 176: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

163

[64] M. J. Kerr and A. Cuevas, "Very low bulk and surface recombination in oxidized silicon wafers,"

Semiconductor Science and Technology, vol. 17, no. 1, p. 35, 2002.

[65] T. C. Kho, S. C. Baker-Finch, and K. R. McIntosh, "The study of thermal silicon dioxide electrets

formed by corona discharge and rapid-thermal annealing," Journal of Applied Physics, vol. 109,

no. 5, p. 053108, 2011.

[66] J. Zhao, A. Wang, P. P. Altermatt, S. R. Wenham, and M. A. Green, "24% efficient perl silicon

solar cell: Recent improvements in high efficiency silicon cell research," Solar Energy Materials

and Solar Cells, vol. 41-42, pp. 87-99, 1996.

[67] L. Do Thanh and P. Balk, "Elimination and Generation of Si ‐ SiO2 Interface Traps by Low

Temperature Hydrogen Annealing," Journal of The Electrochemical Society, vol. 135, no. 7, pp.

1797-1801, 1988.

[68] S. Glunz and F. Feldmann, SiO2 surface passivation layers – a key technology for silicon solar

cells. 2018.

[69] K. A. Collett et al., "An enhanced alneal process to produce SRV < 1cm/s in 1Ωcm n-type Si,"

Solar Energy Materials and Solar Cells, 2017.

[70] J. Benick, K. Zimmermann, J. Spiegelman, M. Hermle, and S. W. Glunz, "Rear side passivation

of PERC-type solar cells by wet oxides grown from purified steam," Progress in Photovoltaics:

Research and Applications, vol. 19, no. 3, pp. 361-365, 2011.

[71] T. Mueller, S. Schwertheim, M. Scherff, and W. R. Fahrner, "High quality passivation for

heterojunction solar cells by hydrogenated amorphous silicon suboxide films," Applied Physics

Letters, vol. 92, no. 3, p. 033504, 2008.

[72] P. Hamer, G. Bourret-Sicotte, G. Martins, A. Wenham, R. S. Bonilla, and P. Wilshaw, "A novel

source of atomic hydrogen for passivation of defects in silicon," physica status solidi (RRL) –

Rapid Research Letters, vol. 11, no. 5, 2017, Art. no. 1600448.

[73] A. G. Aberle, S. Glunz, and W. Warta, "Impact of illumination level and oxide parameters on

Shockley–Read–Hall recombination at the Si‐SiO2 interface," Journal of Applied Physics, vol.

71, no. 9, pp. 4422-4431, 1992.

[74] H. F. Sterling and R. C. G. Swann, "Chemical vapour deposition promoted by r.f. discharge,"

Solid-State Electronics, vol. 8, no. 8, pp. 653-654, 1965.

[75] A. G. Aberle and R. Hezel, "Progress in Low-temperature Surface Passivation of Silicon Solar

Cells using Remote-plasma Silicon Nitride," Progress in Photovoltaics: Research and

Applications, vol. 5, no. 1, pp. 29-50, 1997.

[76] S. Duttagupta, "Advance Surface Passivation of Crystalline Silicon for Solar Cell Application,"

Department of Electrical and Computer Engineering, NATIONAL UNIVERSITY OF SINGAPORE,

2014.

[77] Y. Wan, "Highly Transparent and Highly Passivating Silicon Nitride for Solar Cells," The

Australian National University, 2014.

[78] M. J. Kerr, "Surface, Emitter and Bulk Recombination in Silicon and Development of Silicon

Nitride Passivated Solar Cells," The Australian National Univerisity, 2002.

[79] T. Lauinger, J. Moschner, A. G. Aberle, and R. Hezel, "Optimization and characterization of

remote plasma-enhanced chemical vapor deposition silicon nitride for the passivation of p-

type crystalline silicon surfaces," Journal of Vacuum Science & Technology A: Vacuum, Surfaces,

and Films, vol. 16, no. 2, pp. 530-543, 1998.

[80] H. Mäckel and R. Lüdemann, "Detailed study of the composition of hydrogenated SiNx layers

for high-quality silicon surface passivation," Journal of Applied Physics, vol. 92, no. 5, pp. 2602-

2609, 2002.

[81] D. Jousse, J. Kanicki, and J. H. Stathis, "Observation of multiple silicon dangling bond

configurations in silicon nitride," Applied Physics Letters, vol. 54, no. 11, pp. 1043-1045, 1989.

[82] J. Schmidt, F. M. Schuurmans, W. C. Sinke, S. W. Glunz, and A. G. Aberle, "Observation of

multiple defect states at silicon–silicon nitride interfaces fabricated by low-frequency plasma-

Page 177: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

164

enhanced chemical vapor deposition," Applied Physics Letters, vol. 71, no. 2, pp. 252-254,

1997.

[83] J. Schmidt and A. G. Aberle, "Carrier recombination at silicon–silicon nitride interfaces

fabricated by plasma-enhanced chemical vapor deposition," Journal of Applied Physics, vol.

85, no. 7, pp. 3626-3633, 1999.

[84] A. Stesmans, "The Si identical to Si3 defect at various (111)Si/SiO2 and (111)Si/Si3N4

interfaces," Semiconductor Science and Technology, vol. 4, no. 12, p. 1000, 1989.

[85] S. Garcia, I. Martil, G. G. Diaz, E. Castan, S. Dueñas, and M. Fernandez, "Deposition of SiNx:H

thin films by the electron cyclotron resonance and its application to Al/SiNx:H/Si structures,"

Journal of Applied Physics, vol. 83, no. 1, pp. 332-338, 1998.

[86] G. Kovacevic and B. Pivac, "Reactions in silicon-nitrogen plasma," Physical Chemistry Chemical

Physics, vol. 19, no. 5, pp. 3826-3836, 2017.

[87] D. B. Beach and J. M. Jasinski, "Excimer laser photochemistry of silane-ammonia mixtures at

193 nm," The Journal of Physical Chemistry, vol. 94, no. 7, pp. 3019-3026, 1990.

[88] H. F. W. Dekkers et al., "Investigation on mc-Si bulk passivation using deuterated silicon-

nitride," in 3rd World Conference onPhotovoltaic Energy Conversion, 2003. Proceedings of,

2003, vol. 1, pp. 983-986.

[89] S. P. P. H. Sio, H.T. Nguyen, D. Yan, T. Trupke, D. Macdonald, "Comparison of Recombination

Activity of Grain Boundaries in Various Multicrystalline Silicon Materials," 31st European

Photovoltaic Solar Energy Conference and Exhibition, pp. 328 - 333, 2015.

[90] M. L. Reed and J. D. Plummer, "Chemistry of Si-SiO2 interface trap annealing," Journal of

Applied Physics, vol. 63, no. 12, pp. 5776-5793, 1988.

[91] J. Schmidt, M. Kerr, and A. Cuevas, "Surface passivation of silicon solar cells using plasma-

enhanced chemical-vapour-deposited SiN films and thin thermal SiO2/plasma SiN stacks,"

Semiconductor Science and Technology, vol. 16, no. 3, p. 164, 2001.

[92] H. F. W. Dekkers, G. Beaucarne, M. Hiller, H. Charifi, and A. Slaoui, "Molecular hydrogen

formation in hydrogenated silicon nitride," Applied Physics Letters, vol. 89, no. 21, p. 211914,

2006.

[93] M. Ikeda, H. Nagayoshi, Y. Onozawa, T. Saitoh, and K. Kamisako, "Analysis of the effect of

hydrogen-radical annealing for SiO2 passivation," Solar Energy Materials and Solar Cells, vol.

48, no. 1, pp. 109-115, 1997.

[94] N. Hiroshi et al., "Effect of Hydrogen-Radical Annealing for SiO2 Passivation," Japanese Journal

of Applied Physics, vol. 35, no. 8B, p. L1047, 1996.

[95] F. E. Rougieux, N. E. Grant, C. Barugkin, D. Macdonald, and J. D. Murphy, "Influence of

Annealing and Bulk Hydrogenation on Lifetime-Limiting Defects in Nitrogen-Doped Floating

Zone Silicon," IEEE Journal of Photovoltaics, vol. 5, no. 2, pp. 495-498, 2015.

[96] M. Sheoran et al., "Hydrogen diffusion in silicon from PECVD silicon nitride," in Photovoltaic

Specialists Conference, 2008. PVSC'08. 33rd IEEE, 2008, pp. 1-4: IEEE.

[97] W. L. Warren, P. M. Lenahan, and S. E. Curry, "First observation of paramagnetic nitrogen

dangling-bond centers in silicon nitride," Physical Review Letters, vol. 65, no. 2, pp. 207-210,

1990.

[98] W. L. Warren, F. C. Rong, E. H. Poindexter, G. J. Gerardi, and J. Kanicki, "Structural

identification of the silicon and nitrogen dangling‐bond centers in amorphous silicon nitride,"

Journal of Applied Physics, vol. 70, no. 1, pp. 346-354, 1991.

[99] D. T. Krick, P. M. Lenahan, and J. Kanicki, "Electrically active point defects in amorphous silicon

nitride: An illumination and charge injection study," Journal of Applied Physics, vol. 64, no. 7,

pp. 3558-3563, 1988.

[100] D. T. Krick, P. M. Lenahan, and J. Kanicki, "Nature of the dominant deep trap in amorphous

silicon nitride," Physical Review B, vol. 38, no. 12, pp. 8226-8229, 1988.

Page 178: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

165

[101] V. Sharma, C. Tracy, D. Schroder, S. Herasimenka, W. Dauksher, and S. Bowden, "Manipulation

of K center charge states in silicon nitride films to achieve excellent surface passivation for

silicon solar cells," Applied Physics Letters, vol. 104, no. 5, p. 053503, 2014.

[102] S. D. Wolf, G. Agostinelli, G. Beaucarne, and P. Vitanov, "Influence of stoichiometry of direct

plasma-enhanced chemical vapor deposited SiNx films and silicon substrate surface roughness

on surface passivation," Journal of Applied Physics, vol. 97, no. 6, p. 063303, 2005.

[103] R. Hezel, K. Blumenstock, and R. Schörner, "Interface States and Fixed Charges in MNOS

Structures with APCVD and Plasma Silicon Nitride," Journal of The Electrochemical Society, vol.

131, no. 7, pp. 1679-1683, 1984.

[104] D. Landheer, G. C. Aers, G. I. Sproule, R. Khatri, P. J. Simpson, and S. C. Gujrathi, "Positron

study of plasma‐enhanced chemical vapor deposited silicon nitride films," Journal of Applied

Physics, vol. 78, no. 4, pp. 2568-2574, 1995.

[105] P. E. Bagnoli, A. Piccirillo, A. L. Gobbi, and R. Giannetti, "Electrical characteristics of silicon

nitride on silicon and InGaAs as a function of the insulator stoichiometry," Applied Surface

Science, vol. 52, no. 1, pp. 45-52, 1991.

[106] J. F. Lelièvre, E. Fourmond, A. Kaminski, O. Palais, D. Ballutaud, and M. Lemiti, "Study of the

composition of hydrogenated silicon nitride SiNx:H for efficient surface and bulk passivation

of silicon," Solar Energy Materials and Solar Cells, vol. 93, no. 8, pp. 1281-1289, 2009.

[107] M. W. P. E. Lamers, K. T. Butler, J. H. Harding, and A. Weeber, "Interface properties of a-

SiNx:H/Si to improve surface passivation," Solar Energy Materials and Solar Cells, vol. 106, pp.

17-21, 2012.

[108] S. Dauwe, "Low-Temperature Surface Passivation of Crystalline Silicon and its Application to

the Reasr Side of Solar Cells," Hannover Universitat, 2004.

[109] S. Helland, "Electrical Characterization of Amorphous Silicon Nitride Passivation Layers for

Crystalline Silicon Solar Cells," Department of Materials Science and Engineering, Norwegian

University of Science and Technology, 2011.

[110] T. Makino, "Composition and Structure Control by Source Gas Ratio in LPCVD SiNx," Journal of

The Electrochemical Society, vol. 130, no. 2, pp. 450-455, 1983.

[111] E. Bustarret, M. Bensouda, M. C. Habrard, J. C. Bruyère, S. Poulin, and S. C. Gujrathi,

"Configurational statistics in a-SixNyHz alloys: A quantitative bonding analysis," Physical

Review B, vol. 38, no. 12, pp. 8171-8184, 1988.

[112] V. Verlaan et al., "The effect of composition on the bond structure and refractive index of

silicon nitride deposited by HWCVD and PECVD," Thin Solid Films, vol. 517, no. 12, pp. 3499-

3502, 2009.

[113] W. M. M. Kessels et al., "High-rate deposition of a-SiNx:H for photovoltaic applications by the

expanding thermal plasma," Journal of Vacuum Science & Technology A: Vacuum, Surfaces,

and Films, vol. 20, no. 5, pp. 1704-1715, 2002.

[114] P. Doshi, G. E. Jellison, and A. Rohatgi, "Characterization and optimization of absorbing

plasma-enhanced chemical vapor deposited antireflection coatings for silicon photovoltaics,"

Applied Optics, vol. 36, no. 30, pp. 7826-7837, 1997.

[115] S. D. Gupta, B. Hoex, F. Lin, T. Mueller, and A. G. Aberle, "High-quality surface passivation of

low-resistivity p-type C-Si by hydrogenated amorphous silicon nitride deposited by industrial-

scale microwave PECVD," in 2011 37th IEEE Photovoltaic Specialists Conference, 2011, pp.

001421-001423.

[116] A. J. M. v. Erven, R. C. M. Bosch, and M. D. Bijker, "Textured silicon surface passivation by high‐rate expanding thermal plasma deposited SiN and thermal SiO2/SiN stacks for crystalline

silicon solar cells," Progress in Photovoltaics: Research and Applications, vol. 16, no. 7, pp. 615-

627, 2008.

[117] Y. Wan, K. R. McIntosh, and A. F. Thomson, "Characterisation and optimisation of PECVD SiNx

as an antireflection coating and passivation layer for silicon solar cells," AIP Advances, vol. 3,

no. 3, p. 032113, 2013.

Page 179: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

166

[118] T. Jingnan, T. Alexander, L. Alison, and H. Bram, "Unintentional consequences of dual mode

plasma reactors: Implications for upscaling lab-record silicon surface passivation by silicon

nitride," Japanese Journal of Applied Physics, vol. 56, no. 8S2, p. 08MB12, 2017.

[119] Y. Wan, K. R. McIntosh, A. F. Thomson, and A. Cuevas, "Low Surface Recombination Velocity

by Low-Absorption Silicon Nitride on c-Si," IEEE Journal of Photovoltaics, vol. 3, no. 1, pp. 554-

559, 2013.

[120] S. Duttagupta, F. Lin, M. Wilson, M. B. Boreland, B. Hoex, and A. G. Aberle, "Extremely low

surface recombination velocities on low-resistivity n-type and p-type crystalline silicon using

dynamically deposited remote plasma silicon nitride films," Progress in Photovoltaics:

Research and Applications, vol. 22, no. 6, pp. 641-647, 2014.

[121] J. Schmidt and M. Kerr, "Highest-quality surface passivation of low-resistivity p-type silicon

using stoichiometric PECVD silicon nitride," Solar Energy Materials and Solar Cells, vol. 65, no.

1, pp. 585-591, 2001.

[122] Y. Wan and K. R. McIntosh, "On the Surface Passivation of Textured C-Si by PECVD Silicon

Nitride," IEEE Journal of Photovoltaics, vol. 3, no. 4, pp. 1229-1235, 2013.

[123] A. Pakkala and M. Putkonen, "Chapter 8 - Atomic Layer Deposition," in Handbook of

Deposition Technologies for Films and Coatings (Third Edition), P. M. Martin, Ed. Boston:

William Andrew Publishing, 2010, pp. 364-391.

[124] J. Benick et al., "Effect of a post-deposition anneal on Al2O3/Si interface properties," in 35th

IEEE Photovoltaic Specialists Conference, 2010, pp. 000891-000896.

[125] K. Kimoto et al., "Coordination and interface analysis of atomic-layer-deposition Al2O3 on

Si(001) using energy-loss near-edge structures," Applied Physics Letters, vol. 83, no. 21, pp.

4306-4308, 2003.

[126] F. Werner et al., "Electronic and chemical properties of the c-Si/Al2O3 interface," Journal of

Applied Physics, vol. 109, no. 11, p. 113701, 2011.

[127] J. L. Cantin and H. J. von Bardeleben, "An electron paramagnetic resonance study of the

Si(100)/Al2O3 interface defects," Journal of Non-Crystalline Solids, vol. 303, no. 1, pp. 175-178,

2002.

[128] A. Focsa, A. Slaoui, H. Charifi, J. P. Stoquert, and S. Roques, "Surface passivation at low

temperature of p- and n-type silicon wafers using a double layer a-Si:H/SiNx:H," Materials

Science and Engineering: B, vol. 159-160, pp. 242-247, 2009.

[129] G. Dingemans, M. C. M. van de Sanden, and W. M. M. Kessels, "Influence of the Deposition

Temperature on the c-Si Surface Passivation by Al2O3 Films Synthesized by ALD and PECVD,"

Electrochemical and Solid-State Letters, vol. 13, no. 3, pp. H76-H79, 2010.

[130] D. K. Simon, P. M. Jordan, T. Mikolajick, and I. Dirnstorfer, "On the Control of the Fixed Charge

Densities in Al2O3-Based Silicon Surface Passivation Schemes," ACS Applied Materials &

Interfaces, vol. 7, no. 51, pp. 28215-28222, 2015.

[131] B. Hoex, J. J. H. Gielis, M. C. M. v. d. Sanden, and W. M. M. Kessels, "On the c-Si surface

passivation mechanism by the negative-charge-dielectric Al2O3," Journal of Applied Physics,

vol. 104, no. 11, p. 113703, 2008.

[132] A. Youngkyoung et al., "Estimation of Interfacial Fixed Charge at Al2O 3 /SiO2 Using Slant-

Etched Wafer for Solar Cell Application," Japanese Journal of Applied Physics, vol. 50, no. 7R,

p. 071503, 2011.

[133] N. M. Terlinden, G. Dingemans, M. C. M. Sanden, and W. M. M. Kessels, Role of field-effect on

c-Si surface passivation by ultrathin (2–20 nm) atomic layer deposited Al2O3. 2010, pp.

112101-112101.

[134] V. Naumann, M. Otto, R. B. Wehrspohn, and C. Hagendorf, "Chemical and structural study of

electrically passivating Al2O3/Si interfaces prepared by atomic layer deposition," Journal of

Vacuum Science & Technology A: Vacuum, Surfaces, and Films, vol. 30, no. 4, p. 04D106, 2012.

Page 180: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

167

[135] G. Dingemans and W. M. M. Kessels, "Status and prospects of Al2O3-based surface passivation

schemes for silicon solar cells," Journal of Vacuum Science & Technology A: Vacuum, Surfaces,

and Films, vol. 30, no. 4, p. 040802, 2012.

[136] P. Kumar, M. K. Wiedmann, C. H. Winter, and I. Avrutsky, "Optical properties of Al2O3 thin

films grown by atomic layer deposition," Applied Optics, vol. 48, no. 28, pp. 5407-5412, 2009.

[137] C. Barbos et al., "Characterization of Al2O3 Thin Films Prepared by Thermal ALD," Energy

Procedia, vol. 77, pp. 558-564, 2015.

[138] J. Benick, A. Richter, M. Hermle, and S. W. Glunz, "Thermal stability of the Al2O3 passivation

on p-type silicon surfaces for solar cell applications," physica status solidi (RRL) – Rapid

Research Letters, vol. 3, no. 7-8, pp. 233-235, 2009.

[139] G. Dingemans, R. Seguin, P. Engelhart, M. C. M. Sanden, and W. M. M. Kessels, Silicon surface

passivation by ultrathin Al2O3 films synthesized by thermal and plasma atomic layer

deposition. 2009, pp. 10-12.

[140] B. A. Veith-Wolf and J. Schmidt, "Unexpectedly High Minority-Carrier Lifetimes Exceeding

20 ms Measured on 1.4-Ω cm n-Type Silicon Wafers," physica status solidi (RRL) vol. 11, no.

11, 2017.

[141] D. Yan, S. P. Phang, Y. Wan, C. Samundsett, D. Macdonald, and A. Cuevas, "High efficiency n-

type silicon solar cells with passivating contacts based on PECVD silicon films doped by

phosphorus diffusion," Solar Energy Materials and Solar Cells, vol. 193, pp. 80-84, 2019.

[142] L. E. Black, T. Allen, K. R. McIntosh, and A. Cuevas, "Effect of boron concentration on

recombination at the p-Si–Al2O3 interface," Journal of Applied Physics, vol. 115, no. 9, p.

093707, 2014.

[143] M. K. Stodolny et al., "n-Type polysilicon passivating contact for industrial bifacial n-type solar

cells," Solar Energy Materials and Solar Cells, vol. 158, pp. 24-28, 2016.

[144] H. Steinkemper, F. Feldmann, M. Bivour, and M. Hermle, "Numerical Simulation of Carrier-

Selective Electron Contacts Featuring Tunnel Oxides," IEEE Journal of Photovoltaics, vol. 5, no.

5, pp. 1348-1356, 2015.

[145] P. Stradins et al., Passivated Tunneling Contacts to N-Type Wafer Silicon and Their

Implementation into High Performance Solar Cells. 6th World Conference on Photovoltaic

Energy Conversion: WCPEC-6, 2014.

[146] B. Stegemann et al., "Ultra-thin silicon oxide layers on crystalline silicon wafers: Comparison

of advanced oxidation techniques with respect to chemically abrupt SiO2/Si interfaces with

low defect densities," Applied Surface Science, vol. 395, pp. 78-85, 2017.

[147] Y. Larionova et al., "On the recombination behavior of p+-type polysilicon on oxide junctions

deposited by different methods on textured and planar surfaces," physica status solidi (a), pp.

1700058-n/a, 2017, Art. no. 1700058.

[148] D. Yan, A. Cuevas, Y. Wan, and J. Bullock, "Passivating contacts for silicon solar cells based on

boron-diffused recrystallized amorphous silicon and thin dielectric interlayers," Solar Energy

Materials and Solar Cells, vol. 152, no. Supplement C, pp. 73-79, 2016.

[149] D. Yan, A. Cuevas, J. Bullock, Y. Wan, and C. Samundsett, "Phosphorus-diffused polysilicon

contacts for solar cells," Solar Energy Materials and Solar Cells, vol. 142, pp. 75-82, 2015.

[150] H. Felix et al., "Interdigitated back contact solar cells with polycrystalline silicon on oxide

passivating contacts for both polarities," Japanese Journal of Applied Physics, vol. 56, no. 8S2,

p. 08MB15, 2017.

[151] R. Peibst et al., "Working principle of carrier selective poly-Si/c-Si junctions: Is tunnelling the

whole story?," Solar Energy Materials and Solar Cells, vol. 158, pp. 60-67, 2016.

[152] E. Yablonovitch, T. Gmitter, R. M. Swanson, and Y. H. Kwark, "A 720 mV open circuit voltage

SiOx:c‐Si:SiOx double heterostructure solar cell," Applied Physics Letters, vol. 47, no. 11, pp.

1211-1213, 1985.

Page 181: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

168

[153] G. Yang, A. Ingenito, O. Isabella, and M. Zeman, "IBC c-Si solar cells based on ion-implanted

poly-silicon passivating contacts," Solar Energy Materials and Solar Cells, vol. 158, pp. 84-90,

2016.

[154] G. L. Patton, J. C. Bravman, and J. D. Plummer, "Physics, technology, and modeling of

polysilicon emitter contacts for VLSI bipolar transistors," IEEE Transactions on Electron Devices,

vol. 33, no. 11, pp. 1754-1768, 1986.

[155] P. A. Potyraj, D. L. Chen, M. K. Hatalis, and D. W. Greve, "Interfacial oxide, grain size, and

hydrogen passivation effects on polysilicon emitter transistors," IEEE Transactions on Electron

Devices, vol. 35, no. 8, pp. 1334-1343, 1988.

[156] T. N. Truong et al., "Hydrogenation of Phosphorus-Doped Polycrystalline Silicon Films for

Passivating Contact Solar Cells," ACS Applied Materials & Interfaces, vol. 11, no. 5, pp. 5554-

5560, 2019.

[157] B. Gruska, "Ellipsometric analysis of polysilicon layers," Thin Solid Films, vol. 364, no. 1, pp.

138-143, 2000.

[158] Y. Laghla and E. Scheid, "Optical study of undoped, B or P-doped polysilicon," Thin Solid Films,

vol. 306, no. 1, pp. 67-73, 1997.

[159] F. Feldmann, M. Nicolai, R. Müller, C. Reichel, and M. Hermle, "Optical and electrical

characterization of poly-Si/SiOx contacts and their implications on solar cell design," Energy

Procedia, vol. 124, pp. 31-37, 2017.

[160] K. C. Fong et al., "Phosphorus diffused LPCVD polysilicon passivated contacts with in-situ low

pressure oxidation," Solar Energy Materials and Solar Cells, vol. 186, pp. 236-242, 2018.

[161] K. C. Fong et al., "In-situ Low Pressure Thermal Oxidation for Polysilicon Passivated Contacts,"

presented at the Asia-Pacific Solar Research Conference, 2018.

[162] J. Y. Gan and R. M. Swanson, "Polysilicon emitters for silicon concentrator solar cells," in IEEE

Conference on Photovoltaic Specialists, 1990, pp. 245-250.

[163] F. Feldmann, M. Bivour, C. Reichel, H. Steinkemper, M. Hermle, and S. W. Glunz, "Tunnel oxide

passivated contacts as an alternative to partial rear contacts," Solar Energy Materials and

Solar Cells, vol. 131, pp. 46-50, 2014.

[164] S. W. Glunz, D. Biro, S. Rein, and W. Warta, "Field-effect passivation of the SiO2-Si interface,"

Journal of Applied Physics, vol. 86, no. 1, pp. 683-691, 1999.

[165] J. Skalný, G. Hortváth, and N. J. Mason, "Spectra of Ions Produced by Corona Discharges," AIP

Conference Proceedings, vol. 876, no. 1, pp. 284-293, 2006.

[166] K. J. Weber, H. Jin, C. Zhang, N. Nursam, W. E. Jellett, and K. R. McIntosh, "Surface passivation

using dielectric films: how much charge is enough?," 24th European Photovoltaic Solar Energy

Conference, pp. 534-537, 2009.

[167] T. C. Kho, S. C. Baker-Finch, and K. R. McIntosh, "Towards a Silicon Dioxide Electret for Silicon

Solar Cells," 25th European Photovoltaic Solar Energy Conference and Exhibition, 2010.

[168] W. Olthuis and P. Bergveld, "On the charge storage and decay mechanism in silicon dioxide

electrets," IEEE Transactions on Electrical Insulation, vol. 27, no. 4, pp. 691-697, 1992.

[169] A. J. Sprenkels, W. Olthuis, and P. Berveld, "The application of silicon dioxide as an electret

material," in 6th International Symposium on Electrets,(ISE 6) Proceedings., 1988, pp. 165-169.

[170] T. Minami, T. Utsubo, T. Yamatani, T. Miyata, and Y. Ohbayashi, "SiO2 electret thin films

prepared by various deposition methods," Thin Solid Films, vol. 426, no. 1, pp. 47-52, 2003.

[171] R. Kressmann, G. M. Sessler, and P. Gunther, "Space-charge electrets," IEEE Transactions on

Dielectrics and Electrical Insulation, vol. 3, no. 5, pp. 607-623, 1996.

[172] J. A. Voorthuyzen, W. Olthuis, P. Bergveld, and A. J. Sprenkels, "Research and development of

miniaturized electrets," IEEE Transactions on Electrical Insulation, vol. 24, no. 2, pp. 255-266,

1989.

[173] R. B. M. Schofthaler, G. Langguth and J. H. Werner, "High-quality surface passivation by

corona-charged oxides for semiconductor surface characterization," IEEE 1st World

Conference on Photovoltaic Energy Conversion, vol. 2, pp. 1509-1512, 1994.

Page 182: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

169

[174] W. E. Jellett and K. J. Weber, "Accurate measurement of extremely low surface recombination

velocities on charged, oxidized silicon surfaces using a simple metal-oxide-semiconductor

structure," Applied Physics Letters, vol. 90, no. 4, p. 042104, 2007.

[175] C. Thielemann and G. Hess, "Inorganic electret membrane for a silicon microphone," Sensors

and Actuators A: Physical, vol. 61, no. 1, pp. 352-355, 1997.

[176] X. Zhang and G. M. Sessler, "Charge dynamics in silicon nitride/silicon oxide double layers,"

Applied Physics Letters, vol. 78, no. 18, pp. 2757-2759, 2001.

[177] A. Stesmans and V. V. Afanas'ev, "Corona charging damage on thermal Si/SiO2 structures with

nm-thick oxides revealed by electron spin resonance," Microelectronic Engineering, vol. 72,

no. 1, pp. 55-60, 2004.

[178] H. Jin, K. J. Weber, N. C. Dang, and W. E. Jellett, "Defect generation at the Si–SiO2 interface

following corona charging," Applied Physics Letters, vol. 90, no. 26, p. 262109, 2007.

[179] R. A. Sinton and A. Cuevas, "Contactless determination of current–voltage characteristics and

minority‐carrier lifetimes in semiconductors from quasi‐steady‐state photoconductance

data," Applied Physics Letters, vol. 69, no. 17, pp. 2510-2512, 1996.

[180] A. Cuevas and D. Macdonald, "Measuring and interpreting the lifetime of silicon wafers," Solar

Energy, vol. 76, no. 1, pp. 255-262, 2004.

[181] H. T. Nguyen, S. C. Baker-Finch, and D. Macdonald, "Temperature dependence of the radiative

recombination coefficient in crystalline silicon from spectral photoluminescence," Applied

Physics Letters, vol. 104, no. 11, p. 112105, 2014.

[182] W. Shockley and W. T. Read, "Statistics of the Recombinations of Holes and Electrons,"

Physical Review, vol. 87, no. 5, pp. 835-842, 1952.

[183] K. R. McIntosh and L. E. Black, "On effective surface recombination parameters," Journal of

Applied Physics, vol. 116, no. 1, p. 014503, 2014.

[184] D. Kane and R. Swanson, "Measurement of the emitter saturation current by a contactless

photoconductivity decay method," IEEE Photovoltaic Specialists Conference 18, pp. 578-583,

1985.

[185] A. Cuevas and M. A. Balbuena, "Review of analytical models for the study of highly doped

regions of silicon devices," IEEE Transactions on Electron Devices, vol. 36, no. 3, pp. 553-560,

1989.

[186] A. Cuevas, R. Merchan, and J. C. Ramos, "On the systematic analytical solutions for minority-

carrier transport in nonuniform doped semiconductors: application to solar cells," IEEE

Transactions on Electron Devices, vol. 40, no. 6, pp. 1181-1183, 1993.

[187] D. R. Evans. (2016). Basic Semiconductor Material Science and Solid State Physics. Available:

http://web.pdx.edu/~davide/

[188] K. C. Fong, K. R. McIntosh, and A. W. Blakers, "Accurate series resistance measurement of solar

cells," Progress in Photovoltaics: Research and Applications, vol. 21, no. 4, pp. 490-499, 2013.

[189] J.-H. Guo, P. J. Cousins, and J. E. Cotter, "Investigations of parasitic shunt resistance in n-type

buried contact solar cells," Progress in Photovoltaics: Research and Applications, vol. 14, no.

2, pp. 95-105, 2006.

[190] O. Breitenstein, J. P. Rakotoniaina, M. H. Al Rifai, and M. Werner, "Shunt types in crystalline

silicon solar cells," Progress in Photovoltaics: Research and Applications, vol. 12, no. 7, pp. 529-

538, 2004.

[191] M. A. G. J. Zhao, "Optimized antireflection coatings for high-efficiency silicon solar cells," IEEE

Transactions on Electron Devices, vol. 38, no. 8, pp. 1925 -1934, 1991.

[192] S. Mack et al., "Silicon Surface Passivation by Thin Thermal Oxide/PECVD Layer Stack Systems,"

IEEE Journal of Photovoltaics, vol. 1, no. 2, pp. 135-145, 2011.

[193] L. E. Black, B. W. H. van de Loo, B. Macco, J. Melskens, W. J. H. Berghuis, and W. M. M. Kessels,

"Explorative studies of novel silicon surface passivation materials: Considerations and lessons

learned," Solar Energy Materials and Solar Cells, vol. 188, pp. 182-189, 2018.

Page 183: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

170

[194] S. Jan, K. Mark, and C. Andrés, "Surface passivation of silicon solar cells using plasma-enhanced

chemical-vapour-deposited SiN films and thin thermal SiO2/plasma SiN stacks,"

Semiconductor Science and Technology, vol. 16, no. 3, p. 164, 2001.

[195] A. G. Aberle, Crystalline silicon solar cells : advanced surface passivation and analysis. Sydney:

Centre for Photovoltaic Engineering, University of New South Wales, 1999.

[196] S. Dauwe, L. Mittelstädt, A. Metz, and R. Hezel, "Experimental evidence of parasitic shunting

in silicon nitride rear surface passivated solar cells," Progress in Photovoltaics: Research and

Applications, vol. 10, no. 4, pp. 271-278, 2002.

[197] J. D. McBrayer, R. M. Swanson, and T. W. Sigmon, "Diffusion of Metals in Silicon Dioxide,"

Journal of The Electrochemical Society, vol. 133, no. 6, pp. 1242-1246, 1986.

[198] T. Abe and T. Takahashi, "Intrinsic point defect behavior in silicon crystals during growth from

the melt: A model derived from experimental results," Journal of Crystal Growth, vol. 334, no.

1, pp. 16-36, 2011.

[199] N. E. Grant, V. P. Markevich, J. Mullins, A. R. Peaker, F. Rougieux, and D. Macdonald, "Thermal

activation and deactivation of grown-in defects limiting the lifetime of float-zone silicon,"

physica status solidi (RRL), vol. 10, no. 6, pp. 443-447, 2016.

[200] N. E. Grant et al., "Permanent annihilation of thermally activated defects which limit the

lifetime of float-zone silicon," physica status solidi (a), vol. 213, no. 11, pp. 2844-2849, 2016.

[201] G. Dingemans, W. Beyer, M. C. M. v. d. Sanden, and W. M. M. Kessels, "Hydrogen induced

passivation of Si interfaces by Al2O3 films and SiO2/Al2O3 stacks," Applied Physics Letters, vol.

97, no. 15, p. 152106, 2010.

[202] H. Dekkers, Study and Optimization of Dry Process Technologies for Thin Crystalline Silicon

Solar Cell Manufacturing. KATHOLIEKE UNIVERSITEIT LEUVEN 2008.

[203] A. Piccirillo and A. L. Gobbi, "Physical-Electrical Properties of Silicon Nitride Deposited by

PECVD on III-V Semiconductors," Journal of The Electrochemical Society, vol. 137, no. 12, pp.

3910-3917, 1990.

[204] G. C. Schwartz, Y. S. Huang, and W. J. Patrick, "The Effective Dielectric Constant of Silicon

Dioxides Deposited in the Spaces Between Adjacent Conductors," Journal of The

Electrochemical Society, vol. 139, no. 12, pp. L118-L122, 1992.

[205] K. Fouad, V. Kaushal, T. Jie, and J. Chennupati, "Structural, compositional and optical

properties of PECVD silicon nitride layers," Journal of Physics D: Applied Physics, vol. 45, no.

44, p. 445301, 2012.

[206] P. V. Gray and D. M. Brown, "Density of SiO2-Si interface states," Applied Physics Letters, vol.

8, no. 2, pp. 31-33, 1966.

[207] S. C. Vitkavage, E. A. Irene, and H. Z. Massoud, "An investigation of Si-SiO2 interface charges

in thermally oxidized (100), (110), (111), and (511) silicon," Journal of Applied Physics, vol. 68,

no. 10, pp. 5262-5272, 1990.

[208] R. S. Bonilla, F. Woodcock, and P. R. Wilshaw, "Very low surface recombination velocity in n-

type c-Si using extrinsic field effect passivation," Journal of Applied Physics, vol. 116, no. 5, p.

054102, 2014.

[209] Z. R. Chowdhury, K. Cho, and N. P. Kherani, "High-quality surface passivation of silicon using

native oxide and silicon nitride layers," Applied Physics Letters, vol. 101, no. 2, p. 021601, 2012.

[210] Y. Kunimune et al., "Quantitative analysis of hydrogen in SiO2/SiN/SiO2 stacks using atom

probe tomography," AIP Advances, vol. 6, no. 4, p. 045121, 2016.

[211] C. Boehme and G. Lucovsky, "Dissociation reactions of hydrogen in remote plasma-enhanced

chemical-vapor-deposition silicon nitride," Journal of Vacuum Science & Technology A:

Vacuum, Surfaces, and Films, vol. 19, no. 5, pp. 2622-2628, 2001.

[212] R. Bousbih, W. Dimassi, I. Haddadi, and H. Ezzaouia, "The effect of thermal annealing on the

properties of PECVD hydrogenated silicon nitride," physica status solidi (c), vol. 9, no. 10‐11,

pp. 2189-2193, 2012.

Page 184: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

171

[213] Y. Deguchi, M. Ohnishi, Y. Takahashi, and K. Ohnishi, "Thermal annealing effect of silicon

nitride film deposited by photo‐CVD," Electronics and Communications in Japan (Part II:

Electronics), vol. 80, no. 11, pp. 30-38, 1997.

[214] V. Verlaan et al., "High-density silicon nitride deposited at low substrate temperature with

high deposition rate using hot wire chemical vapour deposition," Surface and Coatings

Technology, vol. 201, no. 22, pp. 9285-9288, 2007.

[215] V. Verlaan, C. Van der Werf, W. Arnoldbik, H. Goldbach, and R. Schropp, "Unambiguous

determination of Fourier-transform infrared spectroscopy proportionality factors: The case of

silicon nitride," Physical Review B, vol. 73, no. 19, p. 195333, 2006.

[216] A. de Calheiros Velozo, G. Lavareda, C. Nunes de Carvalho, and A. Amaral, "Thermal

dehydrogenation of amorphous silicon deposited on c-Si: Effect of the substrate temperature

during deposition," physica status solidi (c), vol. 9, no. 10-11, pp. 2198-2202, 2012.

[217] S. C. Baker-Finch and K. R. McIntosh, "The contribution of planes, vertices, and edges to

recombination at pyramidally textured surfaces," IEEE Journal of Photovoltaics, vol. 1, no. 1,

pp. 59-65, 2011.

[218] K. R. McIntosh and L. P. Johnson, "Recombination at textured silicon surfaces passivated with

silicon dioxide," Journal of Applied Physics, vol. 105, no. 12, p. 124520, 2009.

[219] F. Chen, I. G. Romijn, A. W. Weeber, J. Tan, B. Hallam, and J. Cotter, "Relationship between

PECVD silicon nitride film composition and surface and edge passivation," Proceedings of the

European Photovoltaic Solar Energy Conference, 2007.

[220] H. Dun, P. Pan, F. R. White, and R. W. Douse, "Mechanisms of Plasma‐Enhanced Silicon Nitride

Deposition Using SiH4 / N2 Mixture," Journal of The Electrochemical Society, vol. 128, no. 7,

pp. 1555-1563, 1981.

[221] C. H. Ling, C. Y. Kwok, and K. Prasad, "Silicon Nitride Films Prepared by Plasma-Enhanced

Chemical Vapour Deposition (PECVD) of SiH4/NH3/N2 Mixtures: Some Physical Properties,"

Japanese Journal of Applied Physics, vol. 25, no. 10R, p. 1490, 1986.

[222] H. Seung-Soo, C. Li, G. S. May, and A. Rohatgi, "Modeling the growth of PECVD silicon nitride

films for solar cell applications using neural networks," IEEE Transactions on Semiconductor

Manufacturing, vol. 9, no. 3, pp. 303-311, 1996.

[223] T. Ohmi, S. Sudoh, and H. Mishima, "Static charge removal with IPA solution," IEEE

Transactions on Semiconductor Manufacturing, vol. 7, no. 4, pp. 440-446, 1994.

[224] R. B. T Trupke, F Hudert, P Würfel, J Zhao, A Wang, MA Green, "Effective excess carrier

lifetimes exceeding 100 milliseconds in float zone silicon determined from

photoluminescence," Proceedings of the 19th European Photovoltaic Solar Energy Conference,

pp. 758-61, 2004.

[225] Y. Wan, K. R. McIntosh, A. F. Thomson, and A. Cuevas, "Low surface recombination velocity by

low-absorption silicon nitride on c-Si," IEEE 38th Photovoltaic Specialists Conference, vol. 2,

pp. 1-7, 2012.

[226] G. Dingemans, R. Seguin, P. Engelhart, M. C. M. v. d. Sanden, and W. M. M. Kessels, "Silicon

surface passivation by ultrathin Al2O3 films synthesized by thermal and plasma atomic layer

deposition," physica status solidi (RRL), vol. 4, no. 1-2, pp. 10-12, 2010.

[227] R. Häcker and A. Hangleiter, "Intrinsic upper limits of the carrier lifetime in silicon," Journal of

Applied Physics, vol. 75, no. 11, pp. 7570-7572, 1994.

[228] N. E. Grant et al., "Superacid-Treated Silicon Surfaces: Extending the Limit of Carrier Lifetime

for Photovoltaic Applications," IEEE Journal of Photovoltaics, vol. 7, no. 6, pp. 1574-1583, 2017.

[229] S. Bernd, P. Jana-Isabelle, F. Frank, H. Martin, and G. S. W., "Excellent Surface Passivation

Quality on Crystalline Silicon Using Industrial Scale Direct Plasma TOPCon Deposition

Technology," Solar RRL, p. 1800068, 2018.

[230] T. Niewelt; et al., "Monocrystalline Silicon to the Ultimate Lifetime Limit," Solar Energy

Materials and Solar Cells, (Submitted 2018).

Page 185: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

172

[231] L. J. V. d. Pauw, "A Method of Measuring the Resistivity and Hall Coefficient on Lamellae of

Arbitrary Shape," Philips Technical Review, vol. 20, pp. 220-224.

[232] F. Granek, M. Hermle, D. M. Huljić, O. Schultz-Wittmann, and S. W. Glunz, "Enhanced lateral

current transport via the front N+ diffused layer of n-type high-efficiency back-junction back-

contact silicon solar cells," Progress in Photovoltaics: Research and Applications, vol. 17, no. 1,

pp. 47-56, 2009.

[233] M. J. Kerr, J. Schmidt, A. Cuevas, and J. H. Bultman, "Surface recombination velocity of

phosphorus-diffused silicon solar cell emitters passivated with plasma enhanced chemical

vapor deposited silicon nitride and thermal silicon oxide," Journal of Applied Physics, vol. 89,

no. 7, pp. 3821-3826, 2001.

[234] F. W. Chen, T.-T. A. Li, and J. E. Cotter, "Passivation of boron emitters on n-type silicon by

plasma-enhanced chemical vapor deposited silicon nitride," Applied Physics Letters, vol. 88,

no. 26, p. 263514, 2006.

[235] N. M. Nursam, Y. Ren, and K. J. Weber, "PECVD Silicon Nitride Passivation on Boron Emitter:

The Analysis of Electrostatic Charge on the Interface Properties," Advances in OptoElectronics,

2010.

[236] S. G. M. Kessler, P. Altermatt, N. P. Harder, R. Brendel, "Influence of emitter profile

characteristics on thermal stability and passiviation quality of a-Si/SiNX-passivated boron

emitters," presented at the Photovoltaic Specialists Conference (PVSC), Honolulu, HI, USA,

2010.

[237] J. Schmidt et al., "Advances in the Surface Passivation of Silicon Solar Cells," Energy Procedia,

vol. 15, pp. 30-39, 2012.

[238] P. P. Altermatt et al., "Numerical modeling of highly doped Si:P emitters based on Fermi–Dirac

statistics and self-consistent material parameters," Journal of Applied Physics, vol. 92, no. 6,

pp. 3187-3197, 2002.

[239] K. Dah-Bin, J. P. McVittie, W. D. Nix, and K. C. Saraswat, "Two-dimensional thermal oxidation

of silicon - I. Experiments," IEEE Transactions on Electron Devices, vol. 34, no. 5, pp. 1008-1017,

1987.

[240] J. W. Colby and L. E. Katz, "Boron Segregation at Si  ‐  SiO2 Interface as a Function of

Temperature and Orientation," Journal of The Electrochemical Society, vol. 123, no. 3, pp. 409-

412, 1976.

[241] R. Kotipalli, R. Delamare, O. Poncelet, X. Tang, L. A. Francis, and D. Flandre, "Passivation effects

of atomic-layer-deposited aluminum oxide," EPJ Photovolt., vol. 4, p. 45107, 2013.

[242] M. A. Green, "Self-consistent optical parameters of intrinsic silicon at 300K including

temperature coefficients," Solar Energy Materials and Solar Cells, vol. 92, no. 11, pp. 1305-

1310, 2008.

[243] F. Strinitz, A. El Jaouhari, F. Schoerg, M. Fuerst, M. Plettig, and H. Kuehnlein, "Advanced

alkaline texturing and cleaning for PERC and SHJ solar cells," Energy Procedia, vol. 130, pp. 23-

30, 2017.

[244] H. Savin et al., "Black silicon solar cells with interdigitated back-contacts achieve 22.1%

efficiency," Nature Nanotechnology, Article vol. 10, p. 624, 2015.

[245] R. B. S., "Comparison of TiO2 and other dielectric coatings for buried-contact solar cells: a

review," Progress in Photovoltaics: Research and Applications, vol. 12, no. 4, pp. 253-281, 2004.

[246] J. Zhao and M. A. Green, "Optimized antireflection coatings for high-efficiency silicon solar

cells," IEEE Transactions on Electron Devices, vol. 38, no. 8, pp. 1925-1934, 1991.

[247] E. Forniés, C. Zaldo, and J. M. Albella, "Control of random texture of monocrystalline silicon

cells by angle-resolved optical reflectance," Solar Energy Materials and Solar Cells, vol. 87, no.

1, pp. 583-593, 2005.

[248] E. D. Palik, "Handbook of Optical Constants of Solids," Academic Press, pp. 759–0, 1997.

[249] K. R. McIntosh et al., "Quantifying the optical losses in back-contact solar cells," in 2014 IEEE

40th Photovoltaic Specialist Conference (PVSC), 2014, pp. 0115-0123.

Page 186: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

173

[250] E. Yablonovitch, "Statistical ray optics," Journal of the Optical Society of America, vol. 72, no.

7, pp. 899-907, 1982.

[251] T. Tiedje, E. Yablonovitch, G. D. Cody, and B. G. Brooks, "Limiting efficiency of silicon solar

cells," IEEE Transactions on Electron Devices, vol. 31, no. 5, pp. 711-716, 1984.

[252] G. M. A., "Lambertian light trapping in textured solar cells and light‐emitting diodes: analytical

solutions," Progress in Photovoltaics: Research and Applications, vol. 10, no. 4, pp. 235-241,

2002.

[253] S. C. Baker-Finch, K. R. McIntosh, D. Yan, K. C. Fong, and T. C. Kho, "Near-infrared free carrier

absorption in heavily doped silicon," Journal of Applied Physics, vol. 116, no. 6, p. 063106,

2014.

[254] K. C. Fong et al., "Contact Resistivity of Evaporated Al Contacts for Silicon Solar Cells," IEEE

Journal of Photovoltaics, vol. 5, no. 5, pp. 1304-1309, 2015.

[255] K. R. McIntosh and S. C. Baker-Finch, "A Parameterization of Light Trapping in Wafer-Based

Solar Cells," IEEE Journal of Photovoltaics, vol. 5, no. 6, pp. 1563-1570, 2015.

[256] R. De Rose, M. Zanuccoli, P. Magnone, M. Frei, E. Sangiorgi, and C. Fiegna, Understanding the

Impact of the Doping Profiles on Selective Emitter Solar Cell by Two-Dimensional Numerical

Simulation. 2013, pp. 159-167.

[257] R. Brendel, "Modeling solar cells with the dopant-diffused layers treated as conductive

boundaries," Progress in Photovoltaics: Research and Applications, vol. 20, no. 1, pp. 31-43,

2012.

[258] U. Würfel, A. Cuevas, and P. Würfel, "Charge Carrier Separation in Solar Cells," IEEE Journal of

Photovoltaics, vol. 5, no. 1, pp. 461-469, 2015.

[259] M. K. Mat Desa et al., "Silicon back contact solar cell configuration: A pathway towards higher

efficiency," Renewable and Sustainable Energy Reviews, vol. 60, pp. 1516-1532, 2016.

[260] A. A. Istratov, H. Hieslmair, and E. R. Weber, "Iron contamination in silicon technology,"

Applied Physics A, vol. 70, no. 5, pp. 489-534, 2000.

[261] G. Coletti et al., "Impact of Metal Contamination in Silicon Solar Cells," Advanced Functional

Materials, vol. 21, no. 5, pp. 879-890, 2011.

[262] Y.-J. Liu, D. M. Waugh, and H.-Z. Yu, "Impact of organic contamination on the electrical

properties of hydrogen-terminated silicon under ambient conditions," Applied Physics Letters,

vol. 81, no. 26, pp. 4967-4969, 2002.

[263] W. Kern, "The Evolution of Silicon Wafer Cleaning Technology," Journal of The Electrochemical

Society, vol. 137, no. 6, pp. 1887-1892, 1990.

[264] T. F. et al., "Cleaning after silicon oxide CMP," Microelectronic Engineering, vol. 37-38, pp. 285-

291, 1997.

[265] R. Falster and V. V. Voronkov, "Intrinsic Point Defects and Their Control in Silicon Crystal

Growth and Wafer Processing," MRS Bulletin, vol. 25, no. 6, pp. 28-32, 2000.

[266] T. Abe, "Generation and annihilation of point defects by doping impurities during FZ silicon

crystal growth," Journal of Crystal Growth, vol. 334, no. 1, pp. 4-15, 2011.

[267] W. von Ammon, R. Hölzl, J. Virbulis, E. Dornberger, R. Schmolke, and D. Gräf, "The impact of

nitrogen on the defect aggregation in silicon," Journal of Crystal Growth, vol. 226, no. 1, pp.

19-30, 2001.

[268] J. Mullins et al., "Thermally activated defects in float zone silicon: Effect of nitrogen on the

introduction of deep level states," Journal of Applied Physics, vol. 124, no. 3, p. 035701, 2018.

[269] H. R. Huff, H. Tsuya, U. Gösele, and E. S. E. Division, Silicon Materials Science and Technology:

Proceedings of the Eighth International Symposium on Silicon Materials Science and

Technology (no. 2). Electrochemical Society, 1998.

[270] A. Ourmazd and W. Schröter, "Phosphorus gettering and intrinsic gettering of nickel in silicon,"

Applied Physics Letters, vol. 45, no. 7, pp. 781-783, 1984.

Page 187: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

174

[271] M. Seibt, A. Döller, V. Kveder, A. Sattler, and A. Zozime, "Phosphorus Diffusion Gettering of

Platinum in Silicon: Formation of Near-Surface Precipitates," physica status solidi (b), vol. 222,

no. 1, pp. 327-336, 2000.

[272] W. Schröter and R. Kühnapfel, "Model describing phosphorus diffusion gettering of transition

elements in silicon," Applied Physics Letters, vol. 56, no. 22, pp. 2207-2209, 1990.

[273] M. Syre, S. Karazhanov, B. R. Olaisen, A. Holt, and B. G. Svensson, "Evaluation of possible

mechanisms behind P gettering of iron," Journal of Applied Physics, vol. 110, no. 2, p. 024912,

2011.

[274] S. P. Phang and D. Macdonald, "Direct comparison of boron, phosphorus, and aluminum

gettering of iron in crystalline silicon," Journal of Applied Physics, vol. 109, no. 7, p. 073521,

2011.

[275] S. P. Phang, W. Liang, B. Wolpensinger, M. A. Kessler, and D. Macdonald, "Tradeoffs Between

Impurity Gettering, Bulk Degradation, and Surface Passivation of Boron-Rich Layers on Silicon

Solar Cells," IEEE Journal of Photovoltaics, vol. 3, no. 1, pp. 261-266, 2013.

[276] G. W. Neudeck and R. F. Pierret, Introduction to Microelectronic Fabrication. Prentice Hall,

2002.

[277] K. Shimakura, T. Suzuki, and Y. Yadoiwa, "Boron and phosphorus diffusion through an SiO2

layer from a doped polycrystalline Si source under various drive-in ambients," Solid-State

Electronics, vol. 18, no. 11, 1975.

[278] H. R. Grinolds et al., "Nitrided-oxides for thin gate dielectrics in MOS devices," in International

Electron Devices Meeting, 1982, pp. 42-45.

[279] A. Merlos, M. Acero, M. H. Bao, J. Bausells, and J. Esteve, "TMAH/IPA anisotropic etching

characteristics," Sensors and Actuators A: Physical, vol. 37-38, no. Supplement C, pp. 737-743,

1993.

[280] P. Papet et al., "Pyramidal texturing of silicon solar cell with TMAH chemical anisotropic

etching," Solar Energy Materials and Solar Cells, vol. 90, no. 15, pp. 2319-2328, 2006.

[281] W. Menz, J. Mohr, and O. Paul, Microsystem Technology, 1 ed. WILEY-VCH, 2001.

[282] S. A. Campbell, K. Cooper, L. Dixon, R. Earwaker, S. N. Port, and D. J. Schiffrin, "Inhibition of

pyramid formation in the etching of Si p(100) in aqueous potassium hydroxide-isopropanol,"

Journal of Micromechanics and Microengineering, vol. 5, no. 3, p. 209, 1995.

[283] C.-R. Yang, P.-Y. Chen, Y.-C. Chiou, and R.-T. Lee, "Effects of mechanical agitation and

surfactant additive on silicon anisotropic etching in alkaline KOH solution," Sensors and

Actuators A: Physical, vol. 119, no. 1, pp. 263-270, 2005.

[284] E. Abdur-Rahman, I. Alghoraibi, and H. Alkurdi, "Effect of Isopropyl Alcohol Concentration and

Etching Time on Wet Chemical Anisotropic Etching of Low-Resistivity Crystalline Silicon

Wafer," International Journal of Analytical Chemistry, vol. 2017, p. 9, 2017, Art. no. 7542870.

[285] M. Amouzgar and M. Kahrizi, "A new approach for improving the silicon texturing process

using gas-lift effect," Journal of Physics D: Applied Physics, vol. 45, no. 10, p. 105102, 2012.

[286] I. Zubel, I. Barycka, K. Kotowska, and M. Kramkowska, "Silicon anisotropic etching in alkaline

solutions IV: The effect of organic and inorganic agents on silicon anisotropic etching process,"

Sensors and Actuators A: Physical, vol. 87, no. 3, pp. 163-171, 2001.

[287] J. Wei Chen et al., Preparation of Large Size Pyramidal Texture on N-Type Monocrystalline

Silicon Using TMAH Solution for Heterojunction Solar Cells. 2012, pp. 1815-1819.

[288] I. Zubel 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," Sensors and Actuators A: Physical,

vol. 93, no. 2, pp. 138-147, 2001.

[289] Y. Jiang, X. Zhang, F. Wang, and Y. Zhao, "Optimization of a silicon wafer texturing process by

modifying the texturing temperature for heterojunction solar cell applications," RSC Advances,

10.1039/C5RA09739H vol. 5, no. 85, pp. 69629-69635, 2015.

[290] K. C. F. W. Liang, P. Narangari, M. Ernst, S. Armand, J. Tong, S. Surve, D. Walter, M. Stocks, K.

McIntosh, A. Blakers, "TMAH etchant and monoTEX agent induced highly reproducible and

Page 188: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

175

non-metal contamination c-Si texturing," presented at the Asia Pacific Solar Research

Conference 2018, 2018.

[291] M. A. Kessler, T. Ohrdes, B. Wolpensinger, R. Bock, and N. P. Harder, "Characterisation and

implications of the boron rich layer resulting from open-tube liquid source BBR3 boron

diffusion processes," 34th IEEE Photovoltaic Specialists Conference (PVSC), pp. 001556-001561,

2009.

[292] S. Bandana and S. Chetan Singh, "Impact of a boron rich layer on minority carrier lifetime

degradation in boron spin-on dopant diffused n -type crystalline silicon solar cells,"

Semiconductor Science and Technology, vol. 31, no. 3, p. 035009, 2016.

[293] K. R. McIntosh, "Lumps, Humps and Bumps: Three Detrimental Effects in the Current-Voltage

Curve of Silicon Solar Cells," Centre for Photovoltaic Engineering, UNSW, 2001.

[294] H. Jin and K. J. Weber, "The Effect of LPCVD Silicon Nitride Deposition on the Si-SiO2 Interface

of Oxidized Silicon Wafers," Journal of The Electrochemical Society, vol. 154, no. 1, 2007.

[295] F. K. Chern, "Fabrication and Characterization of Elongate Solar Cells," Centre for Sustainable

Energy Systems, Australian National University, 2012.

[296] A. Edler, V. D. Mihailetchi, L. J. Koduvelikulathu, C. Comparotto, R. Kopecek, and R. Harney,

"Metallization‒induced recombination losses of bifacial silicon solar cells," Progress in

Photovoltaics: Research and Applications, vol. 23, no. 5, pp. 620-627, 2015.

[297] J. L. Murray and A. J. McAlister, "The Al-Si (Aluminum-Silicon) system," Bulletin of Alloy Phase

Diagrams, vol. 5, no. 1, p. 74, 1984.

[298] S. E. F. K. Mangersnes, "Tunneling in back-junction silicon solar cells," presented at the 28th

European Photovoltaic Solar Energy Conference and Exhibition,

[299] P. Spinelli, B. W. H. v. d. Loo, A. H. G. Vlooswijk, W. M. M. Kessels, and I. Cesar, "Quantification

of pn-Junction Recombination in Interdigitated Back-Contact Crystalline Silicon Solar Cells,"

IEEE Journal of Photovoltaics, vol. 7, no. 5, pp. 1176-1183, 2017.

[300] M. Mäenpää, M. A. Nicolet, I. Suni, and E. G. Colgan, "Contact resistivities of sputtered TiN

and Ti TiN metallizations on solar-cell-type-silicon," Solar Energy, vol. 27, no. 4, pp. 283-

287, 1981.

[301] C. Félix, G. Vandoni, W. Harbich, J. Buttet, and R. Monot, "Surface mobility of Ag on Pd(100)

measured by specular helium scattering," Physical Review B, vol. 54, no. 23, pp. 17039-17050,

1996.

[302] M. Harald, I. Andreas, K. Robert, Z. Willem, A. O. Heinz, and D. D. Ewan, "Spectral mismatch in

calibration of photovoltaic reference devices by global sunlight method," Measurement

Science and Technology, vol. 16, no. 6, p. 1250, 2005.

[303] B. F. S. C. and M. K. R., "Reflection of normally incident light from silicon solar cells with

pyramidal texture," Progress in Photovoltaics: Research and Applications, vol. 19, no. 4, pp.

406-416, 2011.

[304] J. Bardeen, "Surface States and Rectification at a Metal Semi-Conductor Contact," Physical

Review, vol. 71, no. 10, pp. 717-727, 1947.

[305] D. K. Schroder, Semiconductor Material and Device Characterization. Wiley-IEEE Press, 1998.

[306] M. J. Turner and E. H. Rhoderick, "Metal-silicon Schottky barriers," Solid-State Electronics, vol.

11, no. 3, pp. 291-300, 1968.

[307] H. B. Michaelson, "The work function of the elements and its periodicity," Journal of Applied

Physics, vol. 48, no. 11, pp. 4729-4733, 1977.

[308] S. A. Meng Tao, Darshak Udeshi, Nasir Basit, Eduardo Maldonado, Wiley P. Kirk, "Low Schottky

barriers on n-type silicon (001)," Applied Physics Letters, vol. 83, no. 13, pp. 2593-2595, 2003.

[309] A. Y. C. Yu, "Electron tunneling and contact resistance of metal-silicon contact barriers," Solid-

State Electronics, vol. 13, no. 2, pp. 239-247, 1970.

[310] H. Yu et al., "Contact resistivities of metal-insulator-semiconductor contacts and metal-

semiconductor contacts," Applied Physics Letters, vol. 108, no. 17, p. 171602, 2016.

Page 189: High efficiency IBC solar cell with Oxide-Nitride-Oxide passivation · 2019. 12. 20. · Teng Choon Kho August 2019 A thesis submitted for the degree of Doctor of Philosophy of the

176

[311] J. Melskens, B. W. H. v. d. Loo, B. Macco, L. E. Black, S. Smit, and W. M. M. Kessels, "Passivating

Contacts for Crystalline Silicon Solar Cells: From Concepts and Materials to Prospects," IEEE

Journal of Photovoltaics, vol. 8, no. 2, pp. 373-388, 2018.