+ All Categories
Home > Documents > Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master...

Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master...

Date post: 09-Aug-2020
Category:
Upload: others
View: 1 times
Download: 0 times
Share this document with a friend
65
Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas-Phase Heterogeneous Carbon Dioxide Reduction by Annabelle Po Yin Wong A thesis submitted in conformity with the requirements for the degree of Master of Science Department of Chemistry University of Toronto © Copyright by Annabelle Po Yin Wong 2017
Transcript
Page 1: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas-Phase Heterogeneous Carbon Dioxide Reduction

by

Annabelle Po Yin Wong

A thesis submitted in conformity with the requirements for the degree of Master of Science

Department of Chemistry University of Toronto

© Copyright by Annabelle Po Yin Wong 2017

Page 2: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

ii

Enhancement of Hydride-Terminated Silicon Nanocrystals for

Gas-Phase Heterogeneous Carbon Dioxide Reduction

Annabelle Po Yin Wong

Master of Science

Department of Chemistry

University of Toronto

2017

Abstract

Utilization of solar energy for the conversion of carbon dioxide into fuels and chemical

feedstocks serves as a promising solution to mitigate anthropogenic greenhouse gas emissions

while providing global energy security and environmental protection. Silicon nanocrystals (ncSi)

are attractive materials for the photoreduction of CO2 due to their favorable earth-abundance,

non-toxicity, optical properties, and high surface area. Herein, through rational modification of

ncSi by introducing boron and phosphorus dopants utilizing a novel, facile sol-gel synthesis, it

was demonstrated that dopants successfully enhanced the gas-phase CO2 adsorption capacity and

solar fuel production rate. Remarkably, it was found that phosphorus-doped ncSi was the best

performer among the various ncSi samples due to the combination of the number of surface

hydrides and the addition of electronegative surface atoms. Preliminary results of the

optimization of reaction conditions in a batch photoreactor enabled an enhancement of the solar

fuel production rate.

Page 3: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

iii

Acknowledgments

I would like to sincerely thank my supervisor, Professor Geoffrey Ozin, for his support, guidance

and encouragement throughout the course of my project at the Department of Chemistry. His

passion has shaped my appreciation for nanochemistry, while his wisdom, mentorship and “crazy

ideas” have helped me develop as a critical thinker, scientist, and innovator.

I owe particular thanks to Dr. Wei Sun and Chenxi Qian for their patience, chemical insights,

and valuable advice on the silicon nanocrystal projects. I also sincerely thank the past members

of the Silicon Team, Dr. Naoto Shirahata and Dr. Dongzhi Chen, for their in-depth discussion on

silicon. This project would also not have been possible without the invaluable discussions and

assistance from my colleagues in the Ozin group. In particular, I would like to thank Sue

Mamiche for always making sure that our lab runs smoothly, Yuchan Dong for the multi-reactor

tests and surface area measurements, Leo Diehl for the multi-reactor tests, Abdinoor Jelle for the

HRTEM and XPS analysis, Mireille Ghoussoub for the TGA, and Jia Jia and Ziqi Zhang for the

high intensity reactor tests. I would also like to acknowledge Thomas Wood, Dr. Paul O’Brien

and Amit Sandhel for their guidance on the photoreactors and answering my endless questions

about the reactor system. And to the 10 am coffee crew for helping me stay awake throughout

the day.

I would like to extend my gratitude to my collaborators in the Solar Fuels Cluster at the

University of Toronto. It was an honor to have worked with such brilliant and passionate

scientists and engineers.

Lastly, I would like to thanks my parents, my sister, and my friends for their continuous

encouragement.

Page 4: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

iv

Table of Contents

Acknowledgments.......................................................................................................................... iii

Table of Contents ........................................................................................................................... iv

List of Tables ................................................................................................................................ vii

List of Schemes ............................................................................................................................ viii

List of Figures ................................................................................................................................ ix

List of Abbreviations ..................................................................................................................... xi

List of Appendices ....................................................................................................................... xiii

Chapter 1 Introduction and Background ..........................................................................................1

Introduction and background ......................................................................................................1

1.1 Solar fuels: towards a sustainable carbon neutral economy ................................................1

1.2 Hydride-terminated silicon nanocrystals for heterogeneous reduction of carbon

dioxide..................................................................................................................................2

1.3 Boron and phosphorus doped silicon nanocrystals ..............................................................3

1.3.1 Boron and phosphorus co-doped silicon nanocrystals and the surface

frustrated Lewis pairs ...............................................................................................4

1.3.2 Other attractive properties of doped silicon nanocrystals ........................................5

1.3.3 Location of dopants..................................................................................................5

1.3.4 Boron and phosphorus doped silicon nanocrystals as a candidate for

heterogeneous reduction of carbon dioxide .............................................................7

1.4 Summary of chapters ...........................................................................................................9

Chapter 2 Synthesis and Characterization of Doped and Pristine Free-Standing Hydride-

Terminated Silicon Nanocrystals ..............................................................................................10

Synthesis and characterization of doped and pristine free-standing hydride-terminated

silicon nanocrystals ...................................................................................................................10

2.1 Experimental ......................................................................................................................10

2.1.1 Preparation of boron and phosphorus doped free-standing hydride-terminated

silicon nanocrystals ................................................................................................10

Page 5: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

v

2.1.2 Characterization .....................................................................................................12

2.2 Results and discussion .......................................................................................................13

2.2.1 Crystal phases, crystallite size, and surface area ...................................................13

2.2.2 ATR-FTIR spectroscopy ........................................................................................15

2.2.3 X-ray photoelectron spectroscopy .........................................................................17

2.2.4 Surface hydrophobicity ..........................................................................................21

2.3 Summary ............................................................................................................................22

Chapter 3 Investigation of Boron and Phosphorus Doped Silicon Nanocrystals in Gas-Phase

Heterogeneous Reduction of Carbon Dioxide ..........................................................................23

Investigation of boron and phosphorus doped silicon nanocrystals in gas-phase

heterogeneous reduction of carbon dioxide ..............................................................................23

3.1 Experimental ......................................................................................................................23

3.1.1 Sample preparation and storage .............................................................................23

3.1.2 Preparation of hydride-terminated silicon nanocrystals and molecular sieves

for water removal experiments ..............................................................................23

3.1.3 Gas-phase CO2 reduction measurements ...............................................................24

3.1.4 Data analysis ..........................................................................................................24

3.1.5 CO2 adsorption capacity ........................................................................................25

3.2 Results and discussion .......................................................................................................25

3.2.1 13CO production via reverse water-gas shift reaction ............................................25

3.2.2 CO2 adsorption capacity ........................................................................................29

3.2.3 Effect of sample storage in air ...............................................................................31

3.2.4 Utilization of only solar energy for RWGS reaction .............................................32

3.2.5 Unwanted side reaction: increase of 12CO production ...........................................34

3.2.6 Removal of water for 13CO production rate enhancement .....................................35

3.3 Summary ............................................................................................................................37

Chapter 4 Conclusion and Outlook ................................................................................................38

Page 6: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

vi

Conclusion and outlook ............................................................................................................38

4.1 Conclusion .........................................................................................................................38

4.2 Outlook ..............................................................................................................................39

References ......................................................................................................................................42

Appendix A : Powder X-ray Diffraction Pattern ...........................................................................47

Appendix B : Irradiance Spectra ....................................................................................................50

Appendix C: Thermogravimetric Analysis ....................................................................................51

Page 7: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

vii

List of Tables

Table 1. Estimated crystallite sizes of doped and pristine ncSi from the PXRD patterns using the

Scherrer equation and from the HRTEM images, and BET surface areas. .................................. 14

Table 2. Dispersibility of hydride-terminated doped and pristine ncSi in polar and non-polar

solvents. ........................................................................................................................................ 22

Page 8: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

viii

List of Schemes

Scheme 1. The preparation of free-standing hydride-terminated B,P-ncSi. The blue and red

circles represent boron and phosphorus dopants, respectively. The boron dopant source, the

phosphorus dopant source, and both dopant sources were omitted for the preparation of P-ncSi,

B-ncSi and ncSi, respectively. The B,P-ncSi were drop-casted onto a borosilicate glass

microfibre filter, which were used for CO2 reduction experiments. ............................................. 11

Page 9: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

ix

List of Figures

Figure 1. PXRD patterns of a) B-ncSi, P-ncSi, B,P-ncSi and pristine ncSi and b) a close-up

region of B,P-ncSi. A small peak at 34.4 ° is observed, indicated by the red arrow. ................... 13

Figure 2. High resolution TEM images of a) ncSi, b) B-ncSi, c) P-ncSi, d-f) B,P-ncSi. Red

circles correspond to the lattice planes of Si (111); yellow, ncBP (111); and green, ncBP (100);

scale bar, 10 nm. ........................................................................................................................... 15

Figure 3. Fourier transform infrared spectra of B-ncSi, P-ncSi, B,P-ncSi and ncSi. ................... 16

Figure 4. Si 2p X-ray photoelectron spectra of a) B-ncSi, b) P-ncSi, c) B,P-ncSi and d) ncSi. ... 18

Figure 5. X-ray photoelectron spectra of the B 1s region of a) B-ncSi and b) B,P-ncSi and P 2p

region of c) P-ncSi and d) B,P-ncSi. ............................................................................................. 19

Figure 6. X-ray photoelectron spectra of the B 1s region of a) B-ncSi embedded in SiO2 and b)

B,P-ncSi embedded in SiO2. ......................................................................................................... 20

Figure 7. Digital photographs of a) ncSi b) P-ncSi c) B-ncSi d) B,P-ncSi in EtOH. ................... 21

Figure 8. Surface area normalized 13CO production rates of B-ncSi, P-ncSi, B,P-ncSi and ncSi

for four to five consecutive runs under the irradiation of 0.20 Sun at 150 °C for 6 h with 1:1 CO2

to H2. ............................................................................................................................................. 26

Figure 9. Surface area normalized 13CO production rates of B-ncSi, P-ncSi, B,P-ncSi, and ncSi,

tested at 150 °C for 6 h with 1:1 CO2 to H2 in the dark (D) and light with 0.2 Sun (L). .............. 28

Figure 10. Surface area normalized carbon dioxide adsorption capacities of B-ncSi, P-ncSi and

ncSi at 35 °C. ................................................................................................................................ 29

Figure 11. Explanation for the increase in CO2 adsorption capacity due to the introduction of

electronegative atoms in b) B-ncSi and c) P-ncSi compared to a) ncSi. The green arrows indicate

the location of potential CO2 adsorption sites. Some bonds are omitted for simplicity. .............. 30

Page 10: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

x

Figure 12. Surface area normalized 13CO production rates of fresh B-ncSi, P-ncSi, B,P-ncSi and

ncSi samples and samples stored in air for 7 and 14 days tested a) in the light and b) in the dark

at 150 °C. ...................................................................................................................................... 31

Figure 13. CO production rates of B-ncSi, P-ncSi, B,P-ncSi and ncSi under irradiation of 50

Suns without external heating for 4 h. .......................................................................................... 33

Figure 14. The 12CO (black) and 13CO (red) GCMS peak areas obtained from the reaction of B,P-

ncSi with CO2 and H2 in the a) light and b) dark at 150 °C over five to six consecutive runs. .... 34

Figure 15. CO production rates of ncSi only and ncSi with 3A molecular sieves for three

consecutive runs under the irradiation of 0.20 Sun at 150 °C with 1:1 ratio of 13CO2 to H2 for 3 h.

Tested molecular sieve beads were replaced by newly regenerated beads before each run. ........ 36

Page 11: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

xi

List of Abbreviations

ATR attenuated total reflectance

BET Brunauer-Emmett-Teller

B-ncSi boron doped silicon nanocrystals

B-ncSi-SiO2 boron doped silicon nanocrystals embedded in SiO2

B,P-ncSi boron and phosphorus co-doped silicon nanocrystals

B,P-ncSi-SiO2 boron and phosphorus co-doped silicon nanocrystals embedded in SiO2

DFT density functional theory

FLP frustrated Lewis pair

FTIR Fourier transform infrared

GC gas chromatography

GCMS gas chromatography mass spectrometry

HRTEM high resolution transmission electron microscopy

LSPR localized surface plasmon resonance

ncSi silicon nanocrystals

ncSi:H hydride-terminated silicon nanocrystals

P-ncSi phosphorus doped silicon nanocrystals

P-ncSi-SiO2 phosphorus doped silicon nanocrystals embedded in SiO2

PXRD powder X-ray diffraction

RWGS reverse water-gas shift

SiNW silicon nanowire

Page 12: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

xii

TGA thermogravimetric analysis

XPS X-ray photoelectron spectroscopy

Page 13: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

xiii

List of Appendices

Appendix A: Powder X-ray Diffraction Pattern……………………………………….………...59

Appendix B: Irradiance Spectra……………………………………………………….…………62

Appendix C: Thermogravimetric Analysis………………………………………………………63

Page 14: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

1

Chapter 1 Introduction and Background

Introduction and background

1.1 Solar fuels: towards a sustainable carbon neutral economy

Anthropogenic greenhouse gas emissions, namely carbon dioxide emissions from combustion of

fossil fuels, are increasing at an alarming rate and are the highest in recorded history.1 These

emissions stem from our dependency on fossil fuel usage, leading to concerning environmental

issues such as climate change. With an increasing demand for global energy consumption and a

diminishing supply of non-renewable fossil fuels, renewable energy such as solar, wind, tidal and

geothermal have been exploited for energy generation. However, the intermittency of these

renewable sources presents a major challenge for delivering energy and electricity on demand, so

the energy generated must be stored.2 In order to mitigate anthropogenic CO2 emissions and to

supply sufficient energy for global usage, the utilization and conversion of CO2 to value-added

chemicals and fuels are key strategies to tackle both of these issues. Some common examples of

value-added chemicals and fuels are carbon monoxide, methane, other short-chain hydrocarbons,

methanol, dimethyl ether, and formic acid. However, CO2 is an extremely thermodynamically

stable molecule and requires an immense amount of energy for the bonds to dissociate and the

conversion to take place.3,4 The ultimate solution is to take advantage of the indefinitely

renewable solar energy, both light and heat generated from the sun, to catalytically convert CO2

into chemical feedstocks and fuels – known as “solar fuels”. The CO2 produced by these

renewable fuels would then be re-captured and recycled into fuels, creating a “neutral carbon

cycle”. In this way, solar energy is captured and stored in the form of chemical bonds of fuels,

which can be simply stored, released on-demand, and safely transported using existing

infrastructure.

To have a meaningful effect on climate change, renewable energy and the environment, solar

fuel production requires a global research effort. To date, the majority of the research focused on

solar-driven hydrogen production by water splitting. Hydrogen is a sustainable fuel when

produced by photocatalytic or photoelectrochemical water splitting and can be used to power, for

example, hydrogen fuel cells. However, it is not a practical choice of fuel due to safety concerns,

Page 15: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

2

storage and transportation issues, and the lack of existing infrastructure for hydrogen.2 In order to

realize more efficient, scalable and economically viable solar fuel production methods,

development of materials for effective CO2 reduction is critical for advancement in this field to

overcome the energy crisis.

As mentioned above, CO is one of the value-added chemical feedstocks which can be produced

from CO2. One may not realize that CO is a useful chemical to produce, but in fact, CO is a

common reagent in the chemical industry. For instance, when combined with H2, this mixture

called syngas can be processed to produce liquid hydrocarbon fuels via the well-known Fischer-

Tropsch synthesis using existing technologies.5–8 The production of CO from CO2 occurs via the

reverse water-gas shift (RWGS) reaction, where CO2 is mixed with hydrogen, and carbon

monoxide and water are produced (Reaction 1).

CO2 + H2 → CO + H2O (1)

Ideally, the CO2 would be captured from the atmosphere or flue gases from industrial plants, and

the hydrogen would be produced by solar-driven water splitting to create a sustainable system.

The RWGS reaction has a high activation energy barrier, and therefore, is an energy intensive

process.5,9 Currently the reaction can be processed at temperatures above 800 °C at 1 atmosphere

with the use of solid metal catalysts in order to selectively convert CO2 to CO and to prevent the

unwanted formation of coke.5 If only this reaction could be driven by solar energy, then a

sustainable carbon-neutral carbon-cycle would be created.

1.2 Hydride-terminated silicon nanocrystals for heterogeneous reduction of carbon dioxide

Many of the catalysts being studied for CO2 reduction consists of “endangered” metal

elements4,10,11 and therefore, it is important to exploit a more sustainable option for solar fuel

production. Silicon is an attractive candidate for sustainable technology applications and for

scaling-up to realize practical solar fuel production due to its earth-abundance, low-cost and low-

toxicity. Nanostructured silicon is of particular interest in solar-to-chemical conversion

applications owing to its high surface area and narrow band gap with a near-infrared to visible

absorption spectrum.12–14 Since the first demonstration of using silicon in photoelectrolysis of

water by Candea et al. in 1976,15 silicon has been investigated for solar fuel generation.

Page 16: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

3

However, research efforts of silicon for solar fuel production is mainly focused on producing H2

from water.12,16 Nanocrystalline silicon has also been studied as a hydrogen storage material due

to its high surface area and its ability to adsorb and desorb hydrogen.17,18

Recently, our group has demonstrated the ability of hydride-terminated silicon nanocrystals,

denoted ncSi:H, for the conversion of carbon dioxide to carbon monoxide via the RWGS

reaction in a hydrogen environment (Reaction 2). This gas-phase reaction was performed in a

batch reactor at 170 °C with a pressure of 27 psi under solar simulated irradiation, and product

gases were tested using 13C isotope tracing experiments. We have found that the surface hydrides

(Si–H) of ncSi:H are responsible for the reduction of CO2 and selectively produce CO under

these conditions.

(2)

Furthermore, we have demonstrated that this is a light-assisted reaction, which can produce CO

up to a rate of 250 μmol g-1 h-1 under 15 Suns at 150 °C. We attributed the light enhancement to

the well-known photothermal effect of ncSi. However, the CO production rate continues to

decrease over subsequent cycles. It was found that this is because the number of Si–H

diminishes, while Si–O–Si and Si–O–H groups are produced as the reaction proceeds. It was also

demonstrated that the reaction between these groups and H2 does not regenerate Si–H bonds. In

other words, the active site, Si–H, is deactivated and the reaction can only proceed

stoichiometrically. With these exciting findings, we set out to further explore the possibilities of

improving the CO2 reducing ability of ncSi, by either enhancing the reaction rate or even better,

rendering the reaction catalytic.

1.3 Boron and phosphorus doped silicon nanocrystals

Boron and phosphorus are the most commonly used dopants for modifying the electronic and

optical properties of silicon by introducing free charge carriers (holes and electrons,

respectively). They are of great importance in optoelectronic applications since the conductivity

of silicon can be tuned by carefully controlling the concentration of substitutional dopants.19–21

Some applications include but are not limited to solar cells20,22,23, light emitting diodes

(LED)14,24, bioimaging25,26, and photocatalysis.12,15,27,28

Page 17: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

4

Motivation for doping ncSi with boron and phosphorus is the possibility of creating a surface

analogue of a molecular frustrated Lewis pair (FLP) to render the CO2 photoreduction of ncSi

catalytic, as well as the opportunity to explore other attractive properties of doped ncSi which

could enhance the reactivity of CO2 towards reduction.

1.3.1 Boron and phosphorus co-doped silicon nanocrystals and the surface frustrated Lewis pairs

The motivation of utilizing boron and phosphorus doped ncSi for CO2 reduction was inspired by

the molecular frustrated Lewis pair (FLP) developed by Stephan and coworkers. An FLP is

generated when the formation of a classical Lewis acid-base adduct is inhibited due to steric

hindrance of bulky ligands. The first FLPs developed were phosphine-borane molecules (P···B)

which are able to reversibly activate hydrogen to form the hydrogenated zwitterion

phosphonium-borate (PH+···BH-) molecules.29–31 After hydrogenation, FLPs are able to reduce

small molecules such as imines and enamines.30,31 Our motivation is to create an analogue of the

molecular FLP on the surface of ncSi with boron and phosphorus, in hopes of heterolytically

splitting H2 on P and B surface atoms which subsequently reduces CO2 to CO, rendering the

photoreduction catalytic. It is also interesting that boron phosphide nanotubes have been shown

to heterolytically dissociate hydrogen in a theoretical study.32

Recently, boron and phosphorus co-doped ncSi, from here on denoted B,P-ncSi, were developed

by Fujii and coworkers and were synthesized by co-sputtering or from hydrogen silsesquioxane

(HSQ).33 Their motivation for synthesizing B,P-ncSi was to develop ligand-free ncSi since

organic capping ligands, which are employed to prevent aggregation of nanoparticles, have

detrimental effects on charge transport. B,P-ncSi were found to be well dispersible in polar

solvents such as methanol and water and therefore could be used for large-scale solution-

processed optoelectronic fabrication and biomedical applications.33–35 Results from theoretical

studies of B,P-ncSi suggest that the polar B–P pairs are preferentially formed on the surface of

the ncSi, which induces electrostatic repulsion between individual B,P-ncSi, enabling colloidal

stability.34,36

Inspired by the above research endeavors, our curiosity was spiked to explore metal-free, earth-

abundant hydride-terminated boron and phosphorus doped ncSi for CO2 reduction, utilizing the

light and heat from the sun.

Page 18: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

5

1.3.2 Other attractive properties of doped silicon nanocrystals

The addition of boron and phosphorus can drastically change the properties of ncSi, which could

assist in the enhancement of CO production rates by hydrogenation of CO2. Our group has

previously demonstrated the increase of CO production rates in the RWGS reaction by

increasing the local heating of nanostructures.13,37 Furthermore, from our study on ncSi:H, we

have demonstrated that the photoreduction of CO2 can be enabled thermally (in the dark) and the

light-enhanced CO production rate was attributed to the photothermal effect. In this way, the

addition of dopants could potentially increase the CO production rates by increasing the local

temperatures of ncSi by two processes: 1) introducing defect states in the band gap of ncSi and

2) introducing localized surface plasmon resonance (LSPR).

By incorporating dopants into ncSi, sub-band gap dopant states are introduced38–40 and non-

radiative recombination of charge carriers could occur via the Shockley-Read-Hall process,

inducing local heating in the nanostructures.41 Shockley-Read-Hall recombination occurs when

excited electrons and holes recombine through these dopant trap states and the energy loss is

released as heat or light.40,42 At sufficiently high dopant concentrations, boron and phosphorus

singly-doped ncSi (B-ncSi and P-ncSi) are also known to demonstrate LSPR,43–45 which occurs

when light interacts with a nanoparticle with dimensions smaller than the incident wavelength,

resulting in a conduction electron plasmon localized at the nanoparticle.46 LSPR is an attractive

property for applications such as photothermal therapy, and hot electron generation for

photocatalysts, as it provides local heating at the nanoscale.41 In this way, if local temperatures

can be increased by non-radiative recombination and plasmonic effects induced by dopants, then

the rate of CO2 to CO conversion rate could be enhanced.

1.3.3 Location of dopants

Within the past two decades, much theoretical and experimental research have been carried out

to examine the location of boron and phosphorus dopants with respect to ncSi, whether they are

preferentially located on the surface, in the sub-surface, in the core of ncSi or in a surrounding

SiO2 matrix. Understanding the doping efficiencies and the preferential locations of dopants are

essential to control the dopant concentrations and to better understand the electronic and optical

properties of ncSi.20,36,47,48 With regards to utilization of B and P doped ncSi in the application of

CO2 reduction, it is worthwhile to examine the location of dopants to elucidate the influence of

Page 19: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

6

the electronic and surface properties of doped ncSi on its ability to reduce CO2 into fuels. This is

especially true since new interfaces and interactions are introduced between the surface of doped

ncSi and gaseous molecules in the system.

The location and doping effects have been studied for both free-standing and silicon oxide

embedded boron and phosphorus doped ncSi. The solubility limits of boron and phosphorus in

silicon are 1% and 0.3%, respectively, which means that there should be dopant segregation on

the surface of ncSi if dopant concentrations are beyond the solubility limit.20,33

Multiple research groups have reported that B atoms preferentially reside on the surface of ncSi,

whereas P atoms are preferentially located in the core of ncSi for single- and co-doped ncSi and

silicon nanowires (SiNWs).36,47,48 Xie et al. reported that boron doped ncSi embedded in a SiO2

matrix synthesized by co-sputtering resulted in the presence of boron on both the surface and in

the core of ncSi, based on the B(0) and oxidized B peaks observed in the B 1s X-ray

photoelectron spectra.48 Fukata et al. synthesized boron and phosphorus singly- and co-doped

SiNWs by catalytic laser ablation. The change in concentrations of B and P from the surface to

the core of the SiNWs was probed by Raman spectroscopy and electron spin resonance,

respectively, using a layer-by-layer oxidation and HF etching technique. It was also found that P

tends to reside in the crystalline core of SiNWs and B near the surface of the SiNWs.49 Guerra et

al. also investigated the preferential locations of p-type (B and Al) and n-type dopants (N and P)

with respect to ncSi, by examining the lowest theoretical formation energies at various locations

of ncSi embedded in SiO2 and hydroxyl-terminated free-standing ncSi. Similar to the

experimental results discussed above, it was reported that B and Al are the most energetically

stable at the surface, whereas N and P are most stable in the core of ncSi. This is because the

diffusion barrier of B and Al atoms within the SiO2 matrix is much lower than N and P atoms,

and that B atoms are the most energetically stable when bound to 2 Si atoms and 2 oxygen atoms

(at the ncSi/SiO2 interface).36 Theoretical calculations demonstrated that in both B and P co-

doped SiNWs and ncSi, the formation of B–P bonds is energetically favorable and stabilizes B

atoms in the crystalline region of silicon near the surface of SiNW and ncSi.36,49 Fujii and

coworkers demonstrated that boron and phosphorus co-doped ncSi are able to disperse very well

in polar solvents such as methanol and water. They attributed their hydrophilicity to a negatively

charged boron-rich surface and the dispersibility in polar solvents to electrostatic repulsion of the

negatively charged co-doped ncSi.33–35

Page 20: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

7

In this way, it is thought that the location of dopants must be independent of synthetic methods,

shape, or size of nanostructured silicon. However, Kortshagen, Pi and colleagues observed that

the opposite occurs for boron and phosphorus singly-doped ncSi synthesized by nonthermal

plasma – P atoms were experimentally found to locate near the surface of ncSi, and more B

atoms exist in the core of ncSi.47,50 They attributed this experimental result to a weaker Si–B

bond compared to Si–P on the surface of ncSi.50

The location of dopants is also helpful in determining whether the dopant element is

electronically active or not. When dopant atoms are substitutionally doped into the core of ncSi,

they are most likely electronically active. However, if dopant atoms reside on the surface of ncSi,

they may be passivated by surface oxides. Then, free charge carriers are compensated and no

longer present, rendering the ncSi electronically inactive. Also, if the B and P dopant

concentrations are equal in B,P-ncSi, then excess holes and electrons from B and P, respectively,

would be compensated.20

Pertaining to the application of reduction of CO2, it is critical to understand the distribution of

dopants with respect to the surface and volume of ncSi. For instance, it would be expected that

dopants that reside on the surface of ncSi would reduce the number of Si–H active sites, which

may decrease the CO production rates. If free charge carriers play a critical role, then having

dopants only on the surface of ncSi may have detrimental effects on the CO production rate as

free charge carriers are compensated by surface passivating atoms. On the other hand, if dopants

in the core do not enhance the CO production rate, then the synthesis method should be tailored

such that most of the dopants are on the surface of the ncSi. Understanding the role of dopants

and charge carriers could also help elucidate the mechanism of the CO2 reduction reaction. For

example, if free charge carriers are absent and the ncSi is still active, then the implication would

be that charge carriers are not directly involved in the reaction mechanism.

1.3.4 Boron and phosphorus doped silicon nanocrystals as a candidate for heterogeneous reduction of carbon dioxide

P-type silicon (p-Si) has been demonstrated to reduce CO2 in the aqueous phase.12,27,28,51,52 The

use of p-Si for CO2 photoreduction to CO can be dated back to the 1980s28,52 and research on the

modification of p-Si for CO2 reduction has been ongoing since.28 However, there are a number of

issues associated with CO2 photoreduction in the aqueous phase, including the competing

Page 21: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

8

hydrogen evolution reaction from water, low solubility of CO2, and the scalability, energy

efficiency and economics of the process.4,12,13,28

To date, almost all boron and phosphorus doped ncSi are synthesized by chemical vapor

deposition, ion implantation, plasma synthesis, laser ablation and co-sputtering.20 The Veinot

group has developed a facile synthesis for high-purity, near monodisperse ncSi utilizing a

(HSiO1.5)n sol-gel precursor. (HSiO1.5)n can be synthesized by hydrolysis and condensation of

HSiCl3, followed by the thermal treatment of (HSiO1.5)n under a slightly reducing atmosphere.

This leads to a disproportionation reaction and nucleation and growth of silicon clusters and ncSi

are formed within a SiO2 matrix. The ncSi can then be liberated from the matrix by etching away

the SiO2 by an aqueous HF solution. Moreover, the size of ncSi can be easily tuned by the

temperature and time of the thermal treatment.53,54 A modified method is employed in the work

presented within this thesis to incorporate boron and phosphorus dopants into ncSi. The synthesis

and characterization methods are detailed in Chapter 2.

To summarize the aforementioned properties of boron and phosphorus doped ncSi, the rate of

CO production from CO2 hydrogenation would, therefore, depend on a combination of the

following:

Number of surface hydrides (the active sites)

Surface area

Electronic properties

Surface species

Photothermal properties

Presence of impurities

Adsorption capacity of CO2

Reactor test conditions

Page 22: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

9

To the best of our knowledge, boron and phosphorus doped ncSi have not been previously

explored for gas-phase heterogeneous reduction of CO2 for solar fuel production. In this thesis,

the effect of introducing boron and phosphorus into ncSi on the light-assisted conversion of CO2

to CO via the reverse water-gas shift reaction is investigated.

1.4 Summary of chapters

This thesis is organized into 4 chapters. In Chapter 2, the experimental procedures for the

synthesis and characterization of B-ncSi, P-ncSi, B,P-ncSi and ncSi are detailed. The

determination of surface species and the location of dopants are also discussed. Chapter 3

outlines the methodologies of the gas-phase CO2 reduction experiments performed in a

photoreactor. The CO production results of B-ncSi, P-ncSi, B,P-ncSi and ncSi in the light, in the

dark, and after storage in air are compared and discussed. This chapter also presents the

preliminary results of the water removal experiment on pristine ncSi. Chapter 4 summarizes the

findings presented in this thesis and provides an outlook on future development on this work.

Page 23: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

10

Chapter 2 Synthesis and Characterization of Doped and Pristine Free-

Standing Hydride-Terminated Silicon Nanocrystals

Synthesis and characterization of doped and pristine free-standing hydride-terminated silicon nanocrystals

2.1 Experimental

2.1.1 Preparation of boron and phosphorus doped free-standing hydride-terminated silicon nanocrystals

The synthesis of boron and phosphorus co-doped silicon nanocrystals (B,P-ncSi) is depicted in

Scheme 1. In a typical synthesis of boron and phosphorus containing cross-linked (HSiO1.5)n sol-

gel polymer, denoted B,P-(HSiO1.5)n, 15 mL of HSiCl3 (148 mmol, Sigma Aldrich) was added to

a beaker equipped with a magnetic stir bar and cooled in a dry ice/isopropanol bath (-78 °C). In a

separate vessel, 1.2 g of boric acid (19 mmol, Sigma Aldrich) and 0.8 mL of phosphoric acid (12

mmol, 85 wt%, Caledon Laboratories) were dissolved in 60 mL of distilled water. This solution

was added to the cooled HSiCl3, while stirring, in one aliquot to immediately yield a white

precipitate. The beaker was then removed from the isopropanol/dry ice bath and left at room

temperature for 30 min to ensure removal of HCl vapour and complete hydrolysis and

condensation. The white precipitate was washed by distilled water and dried by vacuum

filtration. It was then transferred to a round-bottom flask and further dried under vacuum at 100

°C overnight in an oil bath to afford B,P-(HSiO)n which is a white powder.

B-ncSi and P-ncSi were prepared by the same method, but without the addition of H3PO4 and

H3BO3, respectively. Pristine ncSi was prepared without the addition of H3PO4 and H3BO3.

Page 24: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

11

Scheme 1. The preparation of free-standing hydride-terminated B,P-ncSi. The blue and red

circles represent boron and phosphorus dopants, respectively. The boron dopant source, the

phosphorus dopant source, and both dopant sources were omitted for the preparation of P-ncSi,

B-ncSi and ncSi, respectively. The B,P-ncSi were drop-casted onto a borosilicate glass

microfibre filter, which were used for CO2 reduction experiments.

Boron containing (HSiO1.5)n, B-(HSiO1.5)n, was prepared in the same manner by adding 1.2 g of

boric acid dissolved in 60 mL distilled water into 15 mL HSiCl3 (148 mmol, Sigma Aldrich).

Phosphorus containing (HSiO1.5)n, P-(HSiO1.5)n, was prepared by adding of a mixture of 0.8 mL

of phosphoric acid and 60 mL distilled water into 15 mL HSiCl3. Pristine (HSiO1.5)n was

prepared by adding 60 mL distilled water into 15 mL HSiCl3, without the addition of boric acid

or phosphoric acid.

B-(HSiO1.5)n, P-(HSiO1.5)n, B,P-(HSiO1.5)n or (HSiO1.5)n was transferred to a quartz reaction boat

and heated first at 400 °C for 30 min and then at 1100 °C for 2 h (18 °C min-1) under a constant

gas flow of 95% Ar/5% H2 in a quartz tube furnace. Thermal treatment of (HSiO1.5)n in a

reducing environment causes the disproportionation reaction to occur and allows nucleation and

Page 25: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

12

growth of ncSi. The resulting brown powder is B-ncSi, P-ncSi, B,P-ncSi or ncSi, respectively,

embedded in a SiO2 matrix.

In order to liberate doped or pristine ncSi from the SiO2 matrix, 1.2 g of the brown powder was

ground in a mortar and pestle with 20 mL of EtOH to ensure uniformity of grain size and was

transferred to a Teflon beaker equipped with a magnetic stir bar. 40 mL of HF (48%, aq.,

Caledon Laboratories) were subsequently added to the beaker. Personnel should be well trained

in handling HF. This mixture was stirred for 2 h to etch away the SiO2 matrix. For B-ncSi and

B,P-ncSi, the mixture was then centrifuged at 3800 rpm for 10 min, and the HF/EtOH solution

was decanted to isolate the ncSi. The free-standing B-ncSi or B,P-ncSi were then washed with

acetone twice and isolated by centrifugation to ensure complete removal of HF. For P-ncSi and

pristine ncSi, after stirring in the HF/EtOH mixture for 2 h, free-standing P-ncSi or ncSi were

extracted into pentane and isolated by centrifugation. The final products were dried in a vacuum

oven overnight. Different extraction methods are used due to the difference in surface

hydrophilicity of ncSi.

2.1.2 Characterization

Power X-ray diffraction (PXRD) was performed on a Bruker D2-Phaser X-ray diffractometer

using Cu Kα radiation at 30 kV. Fourier-transform infrared (FTIR) spectroscopy was performed

using a Perkin Elmer Spectrum-One FT-IR fitted with a universal attenuated total reflectance

(ATR) sampling accessory with a diamond coated zinc selenide window. X-ray photoelectron

spectroscopy (XPS) was performed using an ultrahigh vacuum chamber with base pressure of

10-9 torr. A Thermo Scientific K-Alpha XPS spectrometer, with an Al Kα X-ray source

operating at 12 kV, 6 A and X-ray wavelengths of 1486.7 eV, was used. The spectra were

obtained with analyzer pass energy of 50 eV with energy spacing of 0.1 eV. Data analysis was

performed using the Thermo Scientific Avantage software. High resolution transmission electron

microscopy (HRTEM) was performed using a Hitachi HF3300 Environmental-CFE-TEM

operated at a voltage of 300 kV. HRTEM images were analyzed with the ImageJ software.

Page 26: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

13

2.2 Results and discussion

B-ncSi, P-ncSi, B,P-ncSi and ncSi were characterized by a variety of methods to determine the

composition, size, and surface properties, which are essential to elucidate the difference in CO2

reduction activity of the various ncSi samples.

2.2.1 Crystal phases, crystallite size, and surface area

The PXRD patterns of B-ncSi, P-ncSi, B,P-ncSi and ncSi are shown in Figure 1. The diffraction

peaks at 28.7 °, 47.7 °, and 56.5 ° correspond to the (111), (220), and (311) lattice planes of

diamond cubic silicon (JCPDS #27-1402), respectively. No other phases were observed in the

diffraction patterns of B-ncSi, P-ncSi, and pristine ncSi, but in the B,P-ncSi sample a small peak

appears at about 34.4 °, corresponding to the (111) lattice plane of cubic boron phosphide

(Figure 1b, JCPDS #11-0119). Sugimoto et al. reported that during the thermal treatment of

silicon-rich borophosphosilicate films at 1200 °C under N2, B-O and P-O bonds were reduced to

cubic BP nanocrystals (ncBP), while Si(III) becomes fully oxidized to Si(IV).55 In this way,

ncBP is most likely formed in the same manner from BPO4, which can be formed by a mixture

of boric and phosphoric acid, in (HSiO1.5)n.56 The presence of ncBP is also confirmed by

HRTEM as discussed below. ncBP being very resistant to HF left residual ncBP in the sample

after etching in HF/EtOH for 6 h.

Figure 1. PXRD patterns of a) B-ncSi, P-ncSi, B,P-ncSi and pristine ncSi and b) a close-up

region of B,P-ncSi. A small peak at 34.4 ° is observed, indicated by the red arrow.

Page 27: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

14

The crystallite sizes of the doped and pristine ncSi are estimated by the Scherrer equation and are

shown in Table 1. It is not surprising that the size of the doped ncSi is larger than pristine ncSi,

for it is known that dopants tend to accelerate the nucleation and growth of ncSi.19,57 It is also

important to note that the peaks of the doped ncSi samples appear to be asymmetrical, which

indicates a bimodal distribution of sizes.

Table 1. Estimated crystallite sizes of doped and pristine ncSi from the PXRD patterns using the

Scherrer equation and from the HRTEM images, and BET surface areas.

Crystallite Size

from PXRD

pattern (nm)

Crystallize Size

from HRTEM

images (nm)

BET Surface Area

(m2 g-1)

B-ncSi 15 5 - 15 290

P-ncSi 30 5 - 10 257

B,P-ncSi 30 5 - 20 296

ncSi 5 5 - 8 327

The PXRD patterns were also obtained prior to the liberation of doped and pristine ncSi

embedded in SiO2 (Appendix A). Sharp peaks, indicative of large crystals, were only observed in

the doped ncSi samples. This demonstrates that boron and phosphorus dopants have an effect on

the growth of ncSi during thermal treatment.

Figure 2 shows typical high resolution TEM images of ncSi, B-ncSi, P-ncSi, and B,P-ncSi

samples. Lattice spacing of 0.31 nm (circled in red), corresponding to Si (111) planes, are

observed in the images for all doped and pristine ncSi. The measured crystallite sizes of ncSi, B-

ncSi, P-ncSi, and B,P-ncSi are shown in Table 1. The large size distribution observed are

supported by the bimodal size distribution shown in the PXRD pattern. In B,P-ncSi, besides Si

(111), lattice planes with d-spacing of 0.25 nm and 0.40 nm, corresponding to the (111) and

(100) planes of cubic BP, were also observed (Figure 2f). The BP lattice planes were not

observed in the HRTEM images of ncSi, B-ncSi or P-ncSi, confirming that the crystals are only

formed in the presence of both B and P dopants and that the lattices correspond to BP. These

results support the results from PXRD.

Page 28: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

15

Figure 2. High resolution TEM images of a) ncSi, b) B-ncSi, c) P-ncSi, d-f) B,P-ncSi. Red

circles correspond to the lattice planes of Si (111); yellow, ncBP (111); and green, ncBP (100);

scale bar, 10 nm.

2.2.2 ATR-FTIR spectroscopy

In order to probe the surface species of ncSi, ATR-FTIR spectroscopy was used to analyze the

samples. The FTIR spectra of freshly etched samples of B-ncSi, P-ncSi, B,P-ncSi and ncSi

(Figure 3) all showed two dominant peaks at ~ 2100 cm-1 and 900 cm-1, attributed to the

characteristic stretching and bending modes of Si–Hx (x = 1, 2, 3), respectively, indicating that

silicon hydrides are the dominant surface species on all samples. A small amount of C-Hx from

residual organics (i.e., solvents) is also observed at ~ 2900 cm-1 and no Si–C at 680 cm-1 is

observed for all samples. No B–H (2500 cm-1 and 1800 cm-1) or B–O (1400 cm-1) peaks are

Page 29: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

16

observed in the IR spectra of B-ncSi and B,P-nSi , and no P–H (2276 cm-1) or P–O–P (920 cm-1)

peaks are observed in the P-ncSi and B,P-ncSi samples.

Figure 3. Fourier transform infrared spectra of B-ncSi, P-ncSi, B,P-ncSi and ncSi.

The surfaces of all samples are slightly oxidized, as evident from the OSi–O, OSi–H, and O–H

stretching modes at ~1060, 2050, and 3300 cm-1, respectively. The relative amount of surface

oxidation was approximated by taking the sum of the transmittance of the OSi–O and O–H peaks

and dividing it by the Si–H peak for each sample. The results showed that for pristine ncSi, the

oxidized Si peaks are 22% of the Si–H peak, the lowest among the 4 samples, while they are

30%, 49%, and 56% for B-ncSi, P-ncSi and B,P-ncSi, respectively. This suggests that ncSi has

the highest amount of Si–H with respect to Si–O, followed by B-ncSi, P-ncSi, and then B,P-ncSi.

The adventitious surface oxidation stemmed from oxidation by solvents or inadvertent exposure

to air and should be minimized in order to maintain the maximum number of Si–H sites to obtain

the highest CO production rates. The amount of oxidation is also examined by other

characterization methods discussed later. Since Si–H is responsible for the reduction of CO2, this

observation is crucial for understanding the difference in CO production rates of doped and

pristine ncSi.

Page 30: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

17

2.2.3 X-ray photoelectron spectroscopy

Doped and pristine ncSi were further characterized by X-ray photoelectron spectroscopy (XPS).

XPS does not only determine the elemental composition and oxidation states of each element,

but can also give us some insights into the location of dopants with respect to ncSi (e.g.

substitutional lattice doping or surface doping). The Si 2p spectra of B-ncSi, P-ncSi, B,P-ncSi,

and ncSi are displayed in Figure 4. The peak at a lower binding energy of ~ 100 eV is assigned

to Si(0), a Si atom that makes up the core of ncSi, whereas the peak at the higher binding energy

103-104 eV originates from Si(IV) (oxidized surface Si atoms). The fitted peaks at 100 eV and

100.7 eV in ncSi are assigned to Si 2p3/2 and Si 2p1/2, respectively (Figure 4d). A third peak at

101 eV is fitted in the Si 2p spectra, which is assigned to a silicon sub-oxide. However, in the B-

ncSi sample (Figure 4a), it appears that the ratio of the intensity of the peak at ~ 101 eV to Si(0)

is much greater in B-ncSi compared to other samples. This is because this peak at ~ 101 eV is

attributed to Si–B, which provides evidence that the incorporation of B in the Si crystal lattice

has been achieved. Therefore, there is a greater percentage of B atoms that are incorporated on

the surface or in the core of silicon in B-ncSi than in B,P-ncSi. This is supported by the B 1s

spectrum discussed below.

The ratio of Si(IV) to Si(0) also varies in different samples, which stems from adventitious

surface oxidation of ncSi. As shown in the Si 2p spectra of B-ncSi and B,P-ncSi in Figure 4a and

4c, the ratio of Si(IV) to Si(0) is the highest, approximately half of the Si is oxidized to Si(IV).

This surface oxidation most likely stemmed from the washing step with acetone.

Page 31: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

18

Figure 4. Si 2p X-ray photoelectron spectra of a) B-ncSi, b) P-ncSi, c) B,P-ncSi and d) ncSi.

A broad peak observed in the B 1s spectra of both B-ncSi and B,P-ncSi is due to the mixed

valence states of boron – trivalent B(0), tetravalent B(0), and B(III) – which have binding

energies of 185 eV, 188 eV, and 193 eV, respectively (Figure 5). The B 1s and P 2p spectra were

not able to be fitted quantitatively due to the low signal-to-noise ratio. Trivalent and tetravalent B

are defined as B atoms that are connected to three and four B or Si atoms, respectively, and are

substitutional B atoms with zero valence charge.44 The oxidized B species at 193 eV are

attributed to surface B atoms which are oxidized in B-ncSi and B,P-ncSi. In this way, the

presence of both B(0) and oxidized B suggests that B atoms exist in the core and on the surface

of the ncSi.

Page 32: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

19

Figure 5. X-ray photoelectron spectra of the B 1s region of a) B-ncSi and b) B,P-ncSi and P 2p

region of c) P-ncSi and d) B,P-ncSi.

The B 1s spectra of B-ncSi and B,P-ncSi embedded in SiO2, denoted B-ncSi-SiO2 and B,P-ncSi-

SiO2, respectively, were also obtained to confirm that substitutional B is present in the ncSi after

thermal treatment (Figure 6). Two peaks at ~ 188.5 eV (B(0)) and ~ 194 eV (B(III)) are observed

in B-ncSi-SiO2 and B,P-ncSi-SiO2, which indicate that B doping in the Si lattice was achieved

after thermal treatment. B atoms related to the B(III) peak likely originate from B–O residing in

the interface between ncSi and SiO2 or within the SiO2 matrix. In B-ncSi-SiO2 (Figure 6a), the

approximate ratio of the area under the B(0) and B(III) peaks is 3:7, meaning that ~ 30% of B is

doped within the ncSi lattice, whereas 70% of the B is either doped on the surface of ncSi or

reside in the SiO2 matrix. On the other hand, the B(0) peak is very weak in B,P-ncSi-SiO2

(Figure 6b). Therefore, there may only be minimal amounts of B dopant existing in the core of

ncSi, supported by the XPS of free-standing B,P-ncSi, even though the same amount of doping

reagent, H3BO3, was used in the synthesis. This is partially due to the formation of some ncBP.

Page 33: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

20

Figure 6. X-ray photoelectron spectra of the B 1s region of a) B-ncSi embedded in SiO2 and b)

B,P-ncSi embedded in SiO2.

In the P 2p spectra of P-ncSi and B,P-ncSi, a broad peak, which stemmed from the mixed

valence states of phosphorus, is observed. The peaks of both P-ncSi and B,P-ncSi are centred at

135 eV, which is attributed to P(V) (oxidized surface P atoms). The lower binding energy of 130

eV is assigned to P(0) (substitutional P in the core of ncSi). The spectrum of P-ncSi is cut off at

the P(0) and the broad peak stretches to a high binding energy of 140 eV, which is assigned to

some P-F bonds on the sample surface. This suggests that all or almost all of the P atoms reside

on the surface of ncSi, existing as P–O and P–F, and minimal or no P atoms reside in the core of

the ncSi. For B,P-ncSi, there appears to be a very small amount of P(0). It is not surprising that

most of the P atoms reside on the surface of ncSi since the solubility limit of P (0.3%) is lower

than B (1%).20,50

From the XPS analysis, it can be concluded that 1) B-ncSi consists of both substitutional and

oxidized surface B atoms, 2) P-ncSi consists of P–O and P–F on the surface of ncSi, and 3) B,P-

ncSi consists of a very small number of substitutional B and P atoms and some surface B–O and

P–O bonds. By examining the surface species and the location of dopants of ncSi, it will allow us

to elucidate the difference in reducing power of B-ncSi, P-ncSi, B,P-ncSi, and ncSi for the

generation of solar fuels by the hydrogenation of CO2 using the light and/or heat of the sun.

Page 34: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

21

2.2.4 Surface hydrophobicity

The surface hydrophobicity of ncSi can be observed by their ability to disperse in various

solvents. It was observed that ncSi and P-ncSi were dispersible in pentane, a non-polar solvent,

and were not dispersible in polar solvents such as ethanol, where the opposite is observed for B-

ncSi and B,P-ncSi (Table 2). As shown in Figure 7a and 7b, ncSi and P-ncSi immediately

precipitated to the bottom of the vial when EtOH was added, which is indicative that these two

samples have hydrophobic surfaces. On the other hand, a brown dispersion is observed for B-

ncSi and B,P-ncSi in EtOH with some ncSi settled to the bottom of the vial (Figure 7c and 7d). It

has been reported that B-doped silicon and B,P-ncSi composed of a boron-rich surface are

hydrophilic25,35,58,59 and that B,P-ncSi can be well dispersed in polar solvents such as MeOH and

water.35 This observation of B-ncSi and B,P-ncSi further supports the XPS results that B atoms

exist on the surface of the B-containing ncSi. The precipitate in B-ncSi and B,P-ncSi in EtOH are

most likely large crystals or ncSi that consist of less surface B atoms. Furthermore, hydride-

terminated ncSi is known to be hydrophobic, suggesting that ncSi and P-ncSi may contain more

surface hydrides than B-ncSi and B,P-ncSi. Therefore, the observation of hydrophobicity of the

various ncSi samples is not only essential to the isolation of the ncSi during synthesis, but also

provides a clue to the surface properties of the ncSi. The ability to modify the surface

hydrophobicity of ncSi without the addition of organic capping ligands is also advantageous

since ncSi can be used for various applications which uses solvents of different polarities, and in

the context of this thesis work, to avoid carbon contamination in the hydrogenation of CO2.

Figure 7. Digital photographs of a) ncSi b) P-ncSi c) B-ncSi d) B,P-ncSi in EtOH.

Page 35: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

22

Table 2. Dispersibility of hydride-terminated doped and pristine ncSi in polar and non-polar

solvents.

Non-polar

(e.g. pentane)

Polar

(e.g. EtOH)

ncSi

P-ncSi

B-ncSi

B,P-ncSi

2.3 Summary

For the first time, B-ncSi, P-ncSi, and B,P-ncSi have been prepared from the (HSiO1.5)n sol-gel

polymer using trichlorosilane, boric acid and phosphoric acid. It was found that there is a

variation of crystallite sizes among the ncSi samples and that the ncSi crystal lattice remained

unchanged when boron and phosphorus were incorporated. No other phases except ncSi were

observed in B-ncSi and P-ncSi. However, some ncBP were observed in B,P-ncSi. It was found

that pristine ncSi consists of the highest ratio of Si–H to Si–O and B,P-ncSi consists of the least

Si–H. From the XPS analysis, it was concluded that B-ncSi consists of some substitutional B

atoms with a majority of B–O on the surface, while P-ncSi only consists of P atoms on the

surface existing as P–O and P–F, and B,P-ncSi only contains minimal amount of substitutional

and oxidized surface B and P atoms. With most of the dopant atoms residing on the surface of

ncSi, it is not surprising that we found no infrared spectroscopy evidence of LSPR due to the

small number of free charge carriers. The surface of pristine ncSi was found to be the least

oxidized and B,P-ncSi is the most oxidized, with B-ncSi and P-ncSi falling in between. Despite

the observation that P-ncSi seems to more oxidized than B-ncSi from the FTIR analysis, P-ncSi

should consist of more Si–H because its surface is hydrophobic, while B-ncSi is hydrophilic.

Characterization from FTIR spectroscopy also confirmed that all of the samples are hydride-

terminated ncSi.

Page 36: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

23

Chapter 3 Investigation of Boron and Phosphorus Doped Silicon

Nanocrystals in Gas-Phase Heterogeneous Reduction of Carbon Dioxide

Investigation of boron and phosphorus doped silicon nanocrystals in gas-phase heterogeneous reduction of carbon dioxide

3.1 Experimental

3.1.1 Sample preparation and storage

Silicon nanocrystal samples were prepared by drop-casting sonicated dispersions of B-ncSi, P-

ncSi, B,P-ncSi or ncSi in acetone onto borosilicate glass microfibre filters, which provide

increased surface area and mechanical stability. The samples were dried under a flow of N2 gas

and subsequently under vacuum for at least 30 min to yield approximately 2 mg of ncSi on a

glass filter. Samples were kept in an inert atmosphere in the dark prior to CO2 reduction

experiments. For the oxidation experiments, samples were stored in air in the dark.

3.1.2 Preparation of hydride-terminated silicon nanocrystals and molecular sieves for water removal experiments

For the water removal experiments, pristine hydride-terminated ncSi were prepared from SiO.

SiO (Sigma Aldrich, –325 mesh) was transferred to a quartz reaction boat and heated at a rate of

18 °C min-1 to 900 °C and then was held at 900 °C for 1 h under a constant gas flow of 95%

Ar/5% H2 in a quartz tube furnace. In order to liberate the ncSi from the SiO2 matrix, 0.3 g of

ncSi embedded in SiO2, a brown powder, was ground in a mortar and pestle with 10 mL of EtOH

to ensure uniformity of grain size and was transferred to a Teflon beaker equipped with a

magnetic stir bar. 20 mL of HF (48%, aq., Caledon Laboratories) were subsequently added to the

beaker and the mixture was stirred for 2 h 30 min. The ncSi were then extracted from the HF

solution by pentane. The sample preparation for CO2 reduction experiments is identical to the

procedure described in Section 3.1.1, except that the ncSi were drop-casted directly from the

pentane dispersion from the extraction step onto the glass filter. 3A molecular sieves (Sigma

Aldrich, 8-12 mesh), which are crystalline microporous aluminosilicates utilized for water

adsorption, were regenerated at 290 °C overnight to remove any adsorbed water molecules.

Page 37: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

24

3.1.3 Gas-phase CO2 reduction measurements

Gas-phase CO2 reduction experiments were conducted in a custom-fabricated 3.8 mL stainless

steel batch reactor with a fused silica view port sealed with Viton O-ring. The reactor was

evacuated by an Alcatel dry pump after placing the sample in the reactor and then was purged

with H2 gas (99.9995%) for 20 min at a flow rate of 15 mL min-1. For the water removal

experiments, three 3A molecular sieve beads were placed inside the batch reactor on the

borosilicate glass microfibre filter containing ncSi, and were replaced with three newly

regenerated beads for each run. After filling the reactor with 15 psi H2 and 15 psi 13CO2 (99.9

atomic %, Sigma Aldrich) – a 1:1 stoichiometric ratio for the RWGS reaction – the reactor was

sealed, heated to, and held at 150 °C. The pressure of the reactor was monitored by an OMEGA

OX309 pressure transducer. The temperature of the reactor was measured by a thermocouple

placed in contact with the sample and was controlled by an OMEGA CN616 6-Zone temperature

controller. The measurements under light were irradiated under a 1000 W Hortilux Blue metal

halide bulb for a period of 3 or 6 h at an intensity of 200 W m-2 (equivalent to 0.20 Sun). For

experiments performed under high intensity light, samples were exposed to a 300 W Xe lamp

(Newport) with an intensity of ~ 50 Suns focused onto the sample. The irradiance spectra of the

lamps can be found in Appendix A. Product gases were separated by a 3’ MoleSieve 13X and a

6’ HayeSep D column and analyzed by a flame ionization detector (FID) and thermal

conductivity detector (TCD) installed in a SRI-8610 gas chromatograph. Isotope tracing

experiments were conducted with an Agilent 7890A gas chromatograph mass spectrometer with

a 60 m GS-CarbonPLOT column fed into the mass spectrometer, and product gases were

detected by a TCD.

3.1.4 Data analysis

A blank borosilicate glass microfibre filter was used as a control, placed in the reactor, and

treated under the identical conditions as the ncSi samples. The 12CO detected from the blank

glass filter was subtracted from the results of the samples. Then the ratio of 12CO to 13CO

produced was estimated by the ratio of the corresponding GCMS peak areas (mass-to-charge

ratios of 28 and 29 amu, respectively). The peaks were fitted assuming that they are Gaussian

and the areas under the curves were determined by the Peak Analyzer tool of OriginPro software.

The ratio of 12CO to 13CO was then applied to the total CO production rates, calculated from the

GC peak area, to obtain the final 13CO production rates.

Page 38: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

25

3.1.5 CO2 adsorption capacity

The CO2 adsorption capacities of doped and pristine samples were investigated using

thermogravimetric analysis (TGA) with a TA Instruments Q500 thermogravimetric analyzer. All

of the following steps were performed at 35 °C. A desorption step was first performed by

flowing N2 at a rate of 100 mL min-1 for 90 min. Then the sample was introduced to 100% dry

CO2 at a flow rate of 100 mL min-1 for 2 h and then followed by a final desorption step by

flowing N2 at a rate of 100 mL min-1 for 90 min. The weight gain resulting from the introduction

of CO2 is used for the calculation of CO2 adsorption capacity.

3.2 Results and discussion

3.2.1 13CO production via reverse water-gas shift reaction

The ability of B-ncSi, P-ncSi, B,P-ncSi, and pristine ncSi to reduce CO2 was examined in a batch

photoreactor with a 1:1 ratio of 13CO2:H2 – the stoichiometric ratio for the RWGS reaction. It is

important to note that 13C isotope tracing experiments were performed to verify that any carbon-

containing products formed did not stem from adventitious carbon residues. The samples in the

reactor were exposed to simulated solar light using a 1000 W metal halide lamp with an intensity

of 0.20 Sun and were heated at 150 °C for 6 h for multiple cycles. Other than unreacted 13CO2,

no other 13C-containing species were detected under these conditions. The results demonstrate

that hydride-terminated B-ncSi, P-ncSi, B,P-ncSi, and pristine ncSi are all able to

heterogeneously reduce CO2 to CO via the RWGS reaction of rates of μmol g-1 h-1. Since the

crystallite size distributions are quite different among the ncSi samples, the 13CO production

rates are normalized to their BET surface areas (Table 1) to ensure that the difference in rates did

not arise from the difference in surface areas. The surface area normalized rates are presented in

Figure 8. Comparing the 13CO production rates of the first run, P-ncSi exhibited the highest rate

– achieving a rate of 214 nmol m-2 h-1 – followed by ncSi, B-ncSi and B,P-ncSi with rates of 114,

71, and 51 nmol m-2 h-1, respectively. In fact, P-ncSi is able to reach a maximum rate of 275

nmol m-2 h-1 for the first run, which corresponds to 71 μmol g-1 h-1. As shown in Figure 8, the

13CO production rate decreased over consecutive runs for all ncSi samples, which is likely

because CO2 underwent a stoichiometric reaction, and not catalytic one, similar to the results

discussed in Section 1.2. This also suggests that B,P-ncSi with surface B and P atoms do not

generate a surface FLP that is able to heterolytically split H2 and subsequently reduce CO2 under

Page 39: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

26

these conditions to render the reaction catalytic. It can be seen that the 13CO production rate

drastically decreased in the second run for all ncSi samples as most of the surface hydrides have

been consumed after the first cycle. Furthermore, the higher the rate of the first run, the greater

the decrease in rate from run 1 to 2. Then, the rate slowly decreases from run 2 to 5, as the

number of available surface hydrides decreases.

Figure 8. Surface area normalized 13CO production rates of B-ncSi, P-ncSi, B,P-ncSi and ncSi

for four to five consecutive runs under the irradiation of 0.20 Sun at 150 °C for 6 h with 1:1 CO2

to H2.

It is interesting to note that the 13CO production rates of B-containing ncSi are lower than the

ncSi, whereas the rate of P-ncSi is greater than ncSi. The lower 13CO production rates of B-ncSi

and B,P-ncSi compared to ncSi may be explained by several reasons. Large ncBP were observed

in the B,P-ncSi samples from the PXRD pattern and HRTEM images and therefore, a control

experiment was carried out to test whether the ncBP are able to reduce CO2. In attempt to afford

ncBP by etching away all of the ncSi from B,P-ncSi-SiO2, B,P-ncSi-SiO2 was stirred in an

HF/EtOH solution for 6 h. However, after 6 h, a small amount of ncSi still remained in the

product. This sample containing mostly ncBP was tested for CO2 reduction activity, but showed

Page 40: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

27

only a small amount of 13CO produced. The ncBP are most likely not able to reduce CO2. In this

case, the inactive ncBP would only add to the mass of the B,P-ncSi sample and subsequently

decrease the “apparent” CO production rate, which if corrected for inactive ncBP, would be even

higher.

Furthermore, after liberation of B-ncSi and B,P-ncSi from the SiO2 matrix, B-ncSi and B,P-ncSi

were extracted by acetone from the HF/EtOH solution. During the extraction procedure,

formation of bubbles was observed in acetone. When the acetone was decanted and distilled

water was added to the B-ncSi, bubbles were again observed. The formation of bubbles in

acetone was most likely formed by the oxidation of surface hydrides by water in the acetone,

producing H2 gas via the following reaction,16

Si–H + H2O → Si–OH + H2 (3)

Here, Si–H (surface silicon hydride), the reducing agent of the RWGS reaction, is oxidized to

Si–OH, which is incapable of reducing CO2.13 Therefore, the CO production rates of B-ncSi and

B,P-ncSi may be lower than the pristine ncSi and P-ncSi due to the surface oxidation during the

extraction procedure. The surface oxidation is also evident from XPS and FTIR spectra. As

shown in the Si 2p XPS (Figure 4), the peak area ratio of Si(IV) to Si(0) in B-ncSi and B,P-ncSi

is much greater than ncSi and P-ncSi. From the analysis from the FTIR spectra (Figure 3), it was

found that B,P-ncSi has the highest ratio of Si–OH to Si–H bonds among all the samples and that

B-ncSi has a higher ratio of Si–OH to Si–H than ncSi. This oxidation problem may be prevented

by using anhydrous polar solvents to wash B-ncSi and B,P-ncSi during this step.

As discussed in Section 1.3, the substitutional doping of B and P into the core of ncSi introduces

excess holes and electrons, respectively. If some P atoms are doped into the core of P-ncSi, then

the presence of excess electrons from P may be a possible contributor to its high CO production

rate. P-ncSi having an excess of electrons compared to B-ncSi could render the surface Si–H

more hydridic by accepting extra electrons from P, thereby giving P-ncSi the highest reducing

power. With regards to B,P-ncSi, any excess electrons from P would be compensated by the

holes from B. As a result, B,P-ncSi most likely does not contain any excess electrons and

therefore does not exhibit an enhanced CO production rate observed for P-ncSi.

Page 41: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

28

Figure 9. Surface area normalized 13CO production rates of B-ncSi, P-ncSi, B,P-ncSi, and ncSi,

tested at 150 °C for 6 h with 1:1 CO2 to H2 in the dark (D) and light with 0.2 Sun (L).

CO2 reduction experiments were also performed in the dark to elucidate the effect of light on the

ncSi samples. In the dark at 150 °C, the first run of P-ncSi also exhibited the highest rate at 92

nmol m-2 h-1 and is followed by B,P-ncSi, ncSi and B-ncSi at 40, 33, and 26 nmol m-2 h-1,

respectively. The corresponding surface area corrected rates of the first run in the light and dark

are plotted in Figure 9. The results demonstrate that the RWGS reaction by hydride-terminated

doped and pristine ncSi can be enabled thermally, as previously reported by our group for ncSi.13

The 13CO production rates of all samples are enhanced in the light at 150 °C compared to the

dark at 150 °C, which clearly illustrates that this is a light-assisted reaction. With only irradiation

of 0.2 Sun, the 13CO production rates of B-ncSi, P-ncSi and ncSi in the light are approximately

2.5 times greater in the light than in the dark. This difference is much smaller for B,P-ncSi

sample most likely due to the suppressed rate from the inactive ncBP. If the experiments are

performed under an irradiation from a higher intensity lamp (e.g. 1 Sun) and with temperature

controlled at 150 °C, then a greater difference in rate between light and dark may be observed

and the photothermal effects can be compared.

Page 42: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

29

3.2.2 CO2 adsorption capacity

The ability of doped and pristine ncSi to adsorb CO2 is measured by thermogravimetric analysis

(TGA) at 35 °C. The TGA curves can be found in Appendix C. The weight gain observed in the

plot after the introduction of CO2 is used for calculating the CO2 adsorption capacity of the

sample and the surface area corrected results are presented in Figure 10. TGA experiments were

conducted at 35 °C since it was difficult to deduce whether the weight increase arose from CO2

adsorption or the reaction of ncSi with CO2 when performed at elevated temperatures. It is

therefore assumed that this trend is similar at 150 °C, the temperature at which the CO2 reduction

experiments were performed. Interestingly, no CO2 adsorption is observed for pristine ncSi and

an enhancement of CO2 adsorption capacity compared to ncSi is observed for B-ncSi (170 nmol

CO2 m-2) and P-ncSi (90 nmol CO2 m

-2).

Figure 10. Surface area normalized carbon dioxide adsorption capacities of B-ncSi, P-ncSi and

ncSi at 35 °C.

The enhancement of CO2 adsorption capability of doped ncSi may be explained by the increase

of electronegativity of boron and phosphorus compared to silicon. The Pauling

electronegativities of Si, B, P, O, and F are 1.90, 2.04, 2.19, 3.5, and 4.0 respectively.36,60,61 From

Page 43: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

30

the density functional theory (DFT) calculations of pristine ncSi:H, we have shown that the

oxygen atom of CO2 approaches the Si bonded to one hydride and undergoes subsequent

dissociation to yield CO and surface Si–OH.13 Following this model, the CO2 adsorption

capacity would be expected to increase as the electronegativity of the dopant increases, as

depicted in Figure 11.

Figure 11. Explanation for the increase in CO2 adsorption capacity due to the introduction of

electronegative atoms in b) B-ncSi and c) P-ncSi compared to a) ncSi. The green arrows indicate

the location of potential CO2 adsorption sites. Some bonds are omitted for simplicity.

When a dopant with higher electronegativity is introduced, electron density is drawn from

surface Si atoms to the dopant atoms (B or P), rendering the Si atom more positively charged

(δ++; Figure 11b and 11c). In this way, the electrostatic attraction between the O on CO2 and Si

should increase, thereby increasing the ncSi’s CO2 adsorption capacity. Although P is more

electronegative than B, the results demonstrated that the adsorption capacity is higher for B-ncSi

than P-ncSi. This could be attributed to the B–O and Si–O bonds on the surface of B-ncSi. In this

case, electron density is drawn towards the oxygen from B or Si, resulting in a partially positive

B or Si atom, attracting nearby CO2 molecules (Figure 11b). Surface P–O and P–F bonds would

also play a role in enhancing the CO2 adsorption since O and F are more electronegative than P

(Figure 11c). Yet, B-ncSi has a greater CO2 adsorption capacity most likely because there are

more B–O and Si–O bonds on the surface than there are P–O, P–F, and Si–O bonds on the

surface of P-ncSi, as evident from the FTIR (Figure 3) and XPS spectra (Figure 5), and the

observations on the surface hydrophobicity (Section 2.2.4). It was demonstrated that the dopant

elements, both B and P, enhance the CO2 adsorption capacity of ncSi. It would be expected that

Page 44: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

31

the CO2 adsorption capacity of B,P-ncSi is similar to B-ncSi since these two samples have

similar surface properties as discussed earlier. Despite that both B-ncSi and P-ncSi have an

enhanced CO2 adsorption capacity, only P-ncSi yielded a higher CO production rate than ncSi. In

this way, the increased CO2 reduction activity of P-ncSi is not solely due to the increase of CO2

adsorption capacity.

3.2.3 Effect of sample storage in air

It is important to examine the stability of silicon hydrides with storage in air if they are to

become technologically viable for solar fuel generation, especially since hydride-terminated ncSi

are prone to surface oxidation. Surface oxidation of ncSi would decrease the number the surface

hydrides which are responsible for CO2 reduction and it would be interesting to examine the

effects of dopants on the stability of CO production. In this way, all ncSi samples were stored in

the air and in the dark for 7 and 14 days and their CO production rates were compared (Figure

12).

Figure 12. Surface area normalized 13CO production rates of fresh B-ncSi, P-ncSi, B,P-ncSi and

ncSi samples and samples stored in air for 7 and 14 days tested a) in the light and b) in the dark

at 150 °C.

Page 45: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

32

Even after 14 days of storage in air, P-ncSi still produces 13CO at a rate of 48 nmol m-2 h-1 or 12

μmol g-1 h-1, approximately 5.5 times greater than ncSi (8.8 nmol m-2 h-1), under light at 150 °C.

Furthermore, the CO production rate of pristine ncSi is the lowest of all samples after 14 days

both in the light and in the dark at 150 °C. Therefore, this suggests that dopants are providing

some stability to the amount of CO produced. The rates of B-ncSi and B,P-ncSi also appear to be

more stable than ncSi and P-ncSi over time of storage. This is most likely because B-ncSi and

B,P-nSi are more oxidized and have less Si–H than ncSi and P-ncSi, leading to a slower rate of

oxidation of surface hydrides. It also appears that the effect of light decreases over storage time,

which is likely because the amount of CO produced after oxidation is limited by the number of

available Si–H.

3.2.4 Utilization of only solar energy for RWGS reaction

Up to this point, doped and pristine ncSi are shown to convert CO2 to CO using solar irradiation

with external heating. However, in order to maximize the usage of solar energy for the RWGS

reaction, experiments were performed to investigate if solar energy alone could effectively

convert 13CO2 to 13CO. In this way, ncSi samples were exposed to a high intensity simulated

solar lamp with an intensity of 50 Suns for 4 h without any additional heating from an external

source. In a realistic process, such high solar intensities can be easily achieved by employing

existing solar concentrators such as parabolic trough solar collectors.62

Page 46: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

33

Figure 13. CO production rates of B-ncSi, P-ncSi, B,P-ncSi and ncSi under irradiation of 50

Suns without external heating for 4 h.

The CO production rate of doped and pristine ncSi are plotted in Figure 13. A fresh sample of P-

ncSi again yielded the highest CO production rate, reaching an impressive rate of 1.86 μmol m-2

h-1, which corresponds to 0.48 mmol g-1 h-1. It is then followed by ncSi, B,P-ncSi, and B-ncSi

with rates of 1.17, 0.64, and 0.55 μmol m-2 h-1. This trend is similar to the results observed under

0.20 Sun at 150 °C.

Page 47: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

34

3.2.5 Unwanted side reaction: increase of 12CO production

Figure 14. The 12CO (black) and 13CO (red) GCMS peak areas obtained from the reaction of

B,P-ncSi with CO2 and H2 in the a) light and b) dark at 150 °C over five to six consecutive runs.

As discussed in Section 3.2.1, the 13C isotope tracing results have confirmed that the 13CO

production rate decreases over consecutive runs. However, the total CO production rate in the

light appears to be stable. It was then observed that the stability of the total CO production rate is

due to an increase of 12CO production over consecutive runs (Figure 14a). Figure 14 shows the

integrated peak areas of 12CO and 13CO obtained from the GCMS over consecutive runs; the

peak area is dependent on the mass of the sample and therefore, the CO production rate of the

sample tested in the light is still higher than one tested in the dark, in spite of the larger 13CO

peak area obtained in the dark. The increase in 12CO production appears to only occur with ncSi

and BP-ncSi samples. This result further places an emphasis on the importance of 13C isotope

tracing experiments. Also, interestingly, the percentage increase of 12CO over consecutive runs

only occurs under light at 150 °C. In the dark at 150 °C, the 12CO peak area appears to be

relatively constant over 5 cycles (Figure 14b). Therefore, the 12C side reaction seems to be a

light-assisted reaction that is competing with the 13CO2 reduction reaction. In other words, as

more surface hydrides are consumed over consecutive runs, more 12C-containing species begin to

react to yield 12CO. Interestingly, it was reported by Kang et al. that hydride-terminated ncSi are

able to react with organic molecules in the presence of light.63 However, more thorough

experiments are required to further elucidate what reactions are occurring and how they are

Page 48: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

35

occurring. If the side reaction can be identified and suppressed, then perhaps the 13CO production

rate may be further increased.

3.2.6 Removal of water for 13CO production rate enhancement

Molecular sieves are aluminosilicates commonly used for removal of water, alcohols, and some

common gasses from chemical processes and 3A molecular sieves are able to adsorb water

molecules.64 Besides CO, water is the other product yielded from the RWGS reaction (Reaction

1). As discussed previously, water can have detrimental effects to the photoreduction rates as it

can oxidize surface Si–H to Si–OH (Reaction 3), which are no longer able to reduce CO2.13

Therefore, according to Le Châtelier's principle, by reducing or eliminating adsorbed or product

water molecules, not only can the RWGS reaction be shifted to the product side, but more

importantly, the number of surface Si–H sites may be maintained by shifting Reaction 3 to the

left hand side.

With this hypothesis, three regenerated molecular sieve beads were placed inside the batch

reactor with a ncSi sample to adsorb any water present in the reactor. They were then tested

under a 1:1 ratio of 13CO2:H2 at 150 °C in the light for 3 h. The results are presented in Figure

15.

Page 49: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

36

Figure 15. CO production rates of ncSi only and ncSi with 3A molecular sieves for three

consecutive runs under the irradiation of 0.20 Sun at 150 °C with 1:1 ratio of 13CO2 to H2 for 3 h.

Tested molecular sieve beads were replaced by newly regenerated beads before each run.

As shown in Figure 15, without the addition of molecular sieves, pristine ncSi is able to produce

CO at a rate of 122 μmol g-1 h-1 for the first run. With the addition of molecular sieves in the

batch reactor, the CO production rate is increased to an impressive rate of 149 μmol g-1 h-1. The

decrease in CO production rates over consecutive runs is also reduced when molecular sieves

were added and even after the second run, the ncSi sample is able to produce 31 μmol g-1 h-1.

Furthermore, a control experiment under identical test conditions was performed by placing three

molecular sieve beads and a blank glass filter in the reactor. No 13C-containing species other than

unreacted 13CO2 were detected, confirming that the molecular sieves are not responsible for

13CO2 reduction. It is also noteworthy that the amount of 12CO produced also diminished with the

addition of molecular sieves, which suggests that the oxidation of carbon contamination species

by water may be suppressed. Molecules which are adsorbed by the beads could be verified by

obtaining a thermogravimetric analysis-mass spectrum (TGA-MS) of the used beads in the

future. Also, it is hypothesized that with higher light intensities, more water would be generated

and adsorbed by the molecular sieves, and the CO production rate per Sun could be further

Page 50: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

37

enhanced. The 13CO production rates of ncSi, B-ncSi, and P-ncSi prepared from (HSiO1.5)n may

also be improved by applying this water removal technique. If water and carbon contamination

are the cause of the side reaction discussed in Section 3.2.5, then the 12CO production rate may

also be suppressed. Other water removal methods could also be explored. For instance, testing

ncSi in a flow reactor may assist in removing water from the system.

3.3 Summary

We have successfully converted gaseous 13CO2 to 13CO via the reverse water-gas shift reaction

utilizing hydride-terminated B-ncSi, P-ncSi, B,P-ncSi, and ncSi as reducing agents in a batch

reactor. The results have shown that this is a stoichiometric light-assisted reaction and can be

enabled by solar energy alone without any auxiliary heating sources. P-ncSi, which consists of

surface P–O and P–F species, demonstrated the highest 13CO production rate among the doped

and pristine ncSi, achieving a maximum rate of 275 nmol m-2 h-1 (equivalent to 71 μmol g-1 h-1)

under the irradiation of 0.2 Sun. Pristine ncSi demonstrated the next highest 13CO production

rate, followed by B-ncSi and B,P-ncSi. It was found that B-ncSi and P-ncSi have greater CO2

adsorption capacities than ncSi. This is likely because they contain more surface polar bonds

(i.e., Si–B, Si–P, B–O, P–O and P–F), increasing the adsorption of CO2 molecules to the more

positively charged surface Si, B, or P. Although B-ncSi showed a greater CO2 adsorption

capacity, its CO production rate is lower than ncSi. This is due to the reduced number of surface

Si–H on B-ncSi compared to ncSi and P-ncSi. A similar trend was also observed when the

samples were tested in the dark at 150 °C. P-ncSi also exhibited the highest rates after storing in

the air for 7 and 14 days and under an intense irradiation of 50 Suns. The remarkably high 13CO

production rate of P-ncSi is therefore attributed to the combination of enhanced CO2 adsorption

capacity, higher number of surface Si–H sites, and the possible enhancement of the hydridic

character and reactivity of Si–H.

Further improvement in CO production rate of pristine ncSi was also investigated. By removing

adsorbed or product water from the batch reactor using molecular sieve beads and shortening the

reaction time to 3 h, an astounding CO production rate of 149 μmol g-1 h-1 could be achieved for

the first run and a more stable rate for 3 consecutive runs was obtained.

Page 51: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

38

Chapter 4 Conclusion and Outlook

Conclusion and outlook

4.1 Conclusion

The work presented in this thesis aims to develop nanocrystalline silicon as a superior

photoreduction agent for solar fuel production. The effects of boron and phosphorus dopants in

hydride-terminated silicon nanocrystals were investigated. Hydride-terminated B-ncSi, P-ncSi,

B,P-ncSi and ncSi were successfully prepared by sol-gel synthesis. Through various

characterization techniques, it was concluded that substitutional doping in the crystal lattice or

surface doping was successful for B-ncSi, P-ncSi and B,P-ncSi, which consist of a bimodal

distribution of crystallite sizes. B-ncSi and B,P-ncSi were found to be more oxidized and consist

of less surface hydrides than P-ncSi and ncSi.

B-ncSi, P-ncSi, B,P-ncSi and ncSi were all capable of reducing 13CO2 to 13CO via the reverse

water-gas shift reaction. All of the samples were shown to undergo a stoichiometric reaction,

likely due to the oxidation of Si–H as reported in our previous study.13 In order to probe the

effect of solar simulated light on the RWGS reaction, all of the samples were subjected to tests in

the light and in the dark. It was demonstrated that CO2 reduction can be enabled thermally and is

enhanced with light. The reducing ability of doped and pristine ncSi using an intense light

without any external heating demonstrates that the reaction can be enabled using solar energy

alone. The dopants also appeared to provide some stability to the 13CO production rate over

storage time in air.

P-ncSi exhibited the highest 13CO production rate among all of the samples, achieving a

remarkable maximum rate of 276 nmol m-2 h-1 under irradiation of only 0.2 Sun within a period

of 6 h. Adventitious oxidation of ncSi has been shown to have a negative effect on the CO

production rate as it reduces the number of Si–H available for CO2 reduction. The lower CO

production rate of B-ncSi compared to ncSi and P-ncSi was attributed to its surface oxidation

and for B,P-ncSi, it was attributed to the combination of surface oxidation and the presence of

inactive ncBP. It was found that the CO2 adsorption capacity of ncSi is enhanced when boron or

phosphorus dopants were added, and it is hypothesized that the electronegativity of surface

Page 52: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

39

species was responsible for the enhancement. However, CO2 adsorption is not the only factor

that yielded a boost in CO production of P-ncSi. The high CO production rate of P-ncSi was

ascribed to the combination of the number of Si–H and the increased CO2 adsorption. The

addition of some substitutional P may also render the surface Si–H more hydridic due to the

excess electrons present in P-ncSi. Removal of water from the photoreactor by molecular sieves

and optimization of test conditions were also shown to be important to the enhancement of CO

production rates.

4.2 Outlook

This thesis presents significant findings to enhance the CO2 reduction capability of inexpensive,

earth-abundant and non-toxic silicon as a reducing agent using the light and/or heat from the sun.

In spite of these findings, much work in the future must be carried out to realize practical fuel

production in a CO2 refinery.

With regards to synthesis and characterization, several parameters could be optimized to control

the concentration dopants of ncSi and the location of these dopants. For instance, the choice of

silane precursor can be explored to allow slower sol-gel network formation and condensation to

better incorporate dopant elements. The rate and time of heating of doped ncSi in the SiO2 matrix

may also have an effect on the incorporation of dopants. Furthermore, the effect of dopant

concentrations has not been investigated for CO2 reduction. The balance between dopant

concentrations and the amount of surface hydride is also important as Si-H are compensated with

increase of surface dopants. Modifications of the synthesis and improvement of handling

procedures could be made to further minimize surface oxidation of ncSi in order to achieve

maximum CO production rates. Other characterization methods such as electron energy loss

spectroscopy (EELS) elemental mapping and ultraviolet photoelectron spectroscopy (UPS)

would be able to confirm the location and electronic properties of dopants. The number of Si-H

responsible for the reduction of CO2 can be quantified using TGA.

In order to optimize the surface properties of ncSi for CO2 reduction, the reaction mechanism of

the doped and pristine ncSi must also be understood. Characterization by PXRD, HRTEM, XPS,

and FTIR could be performed after the CO2 reduction experiments to examine what changes

have taken place. Other characterization techniques such as in situ FTIR and in situ TGA-MS

could assist in determining the surface species formed as the reaction proceeds in the presence of

Page 53: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

40

light compared to in the dark at various temperatures. The electronegativity effects on CO2

adsorption can be examined with the help of DFT calculations by looking at the changes in

Bader charges when dopant atoms are introduced. The introduction of other electronegative

elements could also clarify if the increase of CO2 adsorption stems from polar surface bonds.

Theoretical models could also help identify differences in the reaction mechanism when CO2 and

H2 are introduced between doped and pristine ncSi.

With regards to CO2 reduction testing, test parameters, including but not limited to the

temperature and pressure of the reactor, the type of photoreactor (batch vs flow), intensity of the

light, mass and thickness of the sample, and time of reaction, have not been optimized yet. For

instance, if the all of the possible CO being produced are afforded within 3 h, running the test for

6 h will decrease the CO production rate by half. Therefore, it is imperative to optimize these

conditions in order to boost solar fuel production rates.

Experiments could be performed to probe the photothermal effects of doped and pristine ncSi. In

the current work, irradiation of only 0.2 Sun was utilized for CO2 reduction experiments. In order

to compare the photothermal effects of doped and pristine ncSi, the samples could be tested

under the irradiation of a lamp with a higher intensity (e.g., 1 Sun) with the temperature

controlled at 150 °C. An increase of local temperatures of the doped ncSi may be induced by

non-radiative recombination of free charge carriers and the introduction of dopant trap states.

The local temperature can be probed by Raman spectroscopy.13,37

Many challenges still need to be overcome to realize large-scale solar fuel production despite

recent advances in carbon dioxide conversion and solar fuel production. One of the bottlenecks

of utilizing hydride-terminated ncSi for solar fuel production is its inability to reduce CO2 to CO

catalytically, and one solution is to incorporate a small amount of a co-catalyst to render the

RWGS reaction catalytic. This thesis provided some techniques and insights into the

enhancement of gas-phase light-assisted heterogeneous reduction of CO2 using hydride-

terminated silicon nanocrystals – a material that is inexpensive, earth-abundant, and non-toxic.

We have demonstrated that the incorporation of dopants in ncSi successfully enhanced the CO2

adsorption capacity and the ability of ncSi to convert CO2 to CO in the gas-phase. With these

remarkable findings along with silicon’s attractive properties, silicon nanocrystals, without

Page 54: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

41

doubt, offer the potential of realizing feasible solar fuel production to drive our society towards

carbon neutrality.

Page 55: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

42

References

(1) Dlugokencky, E.; Tans, P. Trends in Atmospheric Carbon Dioxide

www.esrl.noaa.gov/gmd/ccgg/trends/ (accessed Jan 2, 2017).

(2) Olah, G. A.; Prakash, G. K. S.; Goeppert, A. J. Am. Chem. Soc. 2011, 133 (33), 12881–

12898.

(3) Tu, W.; Zhou, Y.; Zou, Z. Adv. Mater. 2014, 26 (27), 4607–4626.

(4) Chang, X.; Wang, T.; Gong, J. Energy Environ. Sci. 2016, 9 (7), 2177–2196.

(5) Wolf, A.; Jess, A.; Kern, C. Chem. Eng. Technol. 2016, 39 (6), 1040–1048.

(6) Smestad, G. P.; Steinfeld, A. Ind. Eng. Chem. Res. 2012, 51, 11828–11840.

(7) Sun, W.; Qian, C.; Chen, K. K.; Ozin, G. A. ChemNanoMat 2016, 2 (9), 847–855.

(8) Herron, J. A.; Kim, J.; Upadhye, A. A.; Huber, G. W.; Maravelias, C. T. Energy Environ.

Sci. 2015, 8 (1), 126–157.

(9) Ghuman, K. K.; Hoch, L. B.; Wood, T. E.; Mims, C.; Singh, C. V.; Ozin, G. A. ACS

Catal. 2016, 6 (9), 5764–5770.

(10) Davies, E.; Renner, R. Chemistry World. 2011, pp 51–54.

(11) North, M.; Styring, P. Faraday Discuss. 2015, 183, 489–502.

(12) Sun, K.; Shen, S.; Liang, Y.; Burrows, P. E.; Mao, S. S.; Wang, D. Chem. Rev. 2014, 114,

8662–8719.

(13) Sun, W.; Qian, C.; He, L.; Ghuman, K. K.; Wong, A. P. Y.; Jia, J.; Jelle, A. A.; O’Brien,

P. G.; Reyes, L. M.; Wood, T. E.; Helmy, A. S.; Mims, C. A.; Singh, C. V.; Ozin, G. A.

Nat. Commun. 2016, 7, 12553.

(14) Priolo, F.; Gregorkiewicz, T.; Galli, M.; Krauss, T. F. Nat. Nanotechnol. 2014, 9 (1), 19–

32.

Page 56: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

43

(15) Candea, R. M.; Kastner, M.; Goodman, R.; Hickok, N. J. Appl. Phys. 1976, 47 (6), 2724–

2726.

(16) Liu, D.; Li, L.; Gao, Y.; Wang, C.; Jiang, J.; Xiong, Y. Angew. Chemie Int. Ed. 2015, 54

(10), 2980–2985.

(17) Kale, P.; Gangal, A. C.; Edla, R.; Sharma, P. Int. J. Hydrogen Energy 2012, 37 (4), 3741–

3747.

(18) Neiner, D.; Kauzlarich, S. M. Chem. Mater. 2010, 22 (2), 487–493.

(19) Xie, M.; Li, D.; Chen, L.; Wang, F.; Zhu, X.; Yang, D. Appl. Phys. Lett. 2013, 102 (12),

123108.

(20) Oliva-Chatelain, B. L.; Ticich, T. M.; Barron, A. R. Nanoscale 2016, 8 (4), 1733–1745.

(21) Lechner, R.; Stegner, A. R.; Pereira, R. N.; Dietmueller, R.; Brandt, M. S.; Ebbers, A.;

Wiggers, H.; Stutzmann, M. J. Appl. Phys. 2008, 104 (5), 53701.

(22) Tian, B.; Zheng, X.; Kempa, T. J.; Fang, Y.; Yu, N.; Yu, G.; Huang, J.; Lieber, C. M.

Nature 2007, 449, 885–889.

(23) Cho, E.-C.; Park, S.; Hao, X.; Song, D.; Conibeer, G.; Park, S.-C.; Green, M. A.

Nanotechnology 2008, 19, 245201.

(24) Ng, W. L.; Lourenc, M. A.; Gwilliam, R. M.; Ledain, S.; Shao, G.; Homewood, K. P.

Nature 2001, 410, 1036–1039.

(25) Fujii, M.; Sugimoto, H.; Hasegawa, M.; Imakita, K. J. Appl. Phys. 2014, 115 (8), 84301.

(26) Sugimoto, H.; Imakita, K.; Fujii, M. RSC Adv. 2015, 5 (119), 98248–98253.

(27) Hinogami, R.; Nakamura, Y.; Yae, S.; Nakato, Y. J. Phys. Chem. B 1998, 102 (6), 974–

980.

(28) White, J. L.; Baruch, M. F.; Pander, J. E.; Hu, Y.; Fortmeyer, I. C.; Park, J. E.; Zhang, T.;

Liao, K.; Gu, J.; Yan, Y.; Shaw, T. W.; Abelev, E.; Bocarsly, A. B. Chem. Rev. 2015, 115

Page 57: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

44

(23), 12888–12935.

(29) Welch, G. C.; Juan, R. R. S.; Masuda, J. D.; Stephan, D. W. Science 2006, 314 (5802),

1124–1126.

(30) Stephan, D. W.; Erker, G. Angew. Chemie Int. Ed. 2010, 49 (1), 46–76.

(31) Stephan, D. W. Org. Biomol. Chem. 2012, 10 (30), 5740–5746.

(32) Beheshtian, J.; Soleymanabadi, H.; Kamfiroozi, M.; Ahmadi, A. J. Mol. Model. 2012, 18

(6), 2343–2348.

(33) Sugimoto, H.; Fujii, M.; Imakita, K. Nanoscale 2014, 6 (21), 12354–12359.

(34) Sugimoto, H.; Fujii, M.; Imakita, K.; Hayashi, S.; Akamatsu, K. J. Phys. Chem. C 2012,

116 (33), 17969–17974.

(35) Fujii, M.; Sugimoto, H.; Imakita, K. Nanotechnology 2016, 27 (26), 262001.

(36) Guerra, R.; Ossicini, S. J. Am. Chem. Soc. 2014, 136 (11), 4404–4409.

(37) Jia, J.; O’Brien, P. G.; He, L.; Qiao, Q.; Fei, T.; Reyes, L. M.; Burrow, T. E.; Dong, Y.;

Liao, K.; Varela, M.; Pennycook, S. J.; Hmadeh, M.; Helmy, A. S.; Kherani, N. P.;

Perovic, D. D.; Ozin, G. A. Adv. Sci. 2016, 3 (10), 1600189.

(38) Fujii, M.; Toshikiyo, K.; Takase, Y.; Yamaguchi, Y.; Hayashi, S.; Fujii, M.; Toshikiyo,

K.; Takase, Y.; Yamaguchi, Y.; Hayashi, S. J. Appl. Phys. 2003, 94 (3), 1990–1995.

(39) Oliveira, E. L. De; Albuquerque, E. L.; Sousa, J. S. De; Farias, G. A. Appl. Phys. Lett.

2009, 94 (10), 103114.

(40) Shockley, W.; Read, W. T. Phys. Rev. 1952, 87 (5), 835–842.

(41) Brongersma, M. L.; Halas, N. J.; Nordlander, P. Nat. Nanotechnol. 2015, 10 (1), 25–34.

(42) Hall, R. N. Phys. Rev. 1952, 87, 387.

(43) Zhou, S.; Pi, X.; Ni, Z.; Ding, Y.; Jiang, Y.; Jin, C.; Delerue, C.; Yang, D.; Nozaki, T.

Page 58: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

45

ACS Nano 2015, 9 (1), 378–386.

(44) Kramer, N. J.; Schramke, K. S.; Kortshagen, U. R. Nano Lett. 2015, 15 (8), 5597–5603.

(45) Zhou, S.; Ni, Z.; Ding, Y.; Sugaya, M.; Pi, X.; Nozaki, T. ACS Photonics 2016, 3 (3),

415–422.

(46) Willets, K. A.; Van Duyne, R. P. Annu. Rev. Phys. Chem. 2007, 58, 267–297.

(47) Pi, X. D.; Gresback, R.; Liptak, R. W.; Campbell, S. A.; Kortshagen, U. Appl. Phys. Lett.

2008, 92 (12), 123102.

(48) Xie, M.; Li, D.; Chen, L.; Wang, F.; Zhu, X.; Yang, D. Appl. Phys. Lett. 2013, 102 (12).

(49) Fukata, N.; Kaminaga, J.; Takiguchi, R.; Rurali, R.; Dutta, M.; Murakami, K. J. Phys.

Chem. C 2013, 117 (39), 20300–20307.

(50) Zhou, S.; Pi, X.; Ni, Z.; Luan, Q.; Jiang, Y.; Jin, C.; Nozaki, T.; Yang, D. Part. Part. Syst.

Charact. 2015, 32 (2), 213–221.

(51) Alenezi, K.; Ibrahim, S. K.; Li, P.; Pickett, C. J. Chem. A Eur. J. 2013, 19 (40), 13522–

13527.

(52) Taniguchi, I.; Aurian-Blajeni, B.; Bockris, J. O. Electrochim. Acta 1984, 29 (7), 923–932.

(53) Kelly, J. A.; Henderson, E. J.; Veinot, J. G. C.; Kelly, J. A.; Henderson, E. J. Chem.

Commun. 2010, 46, 8704–8718.

(54) Clark, R. J.; Aghajamali, M.; Gonzalez, C. M.; Hadidi, L.; Islam, M. A.; Javadi, M.;

Mobarok, M. H.; Purkait, T. K.; Robidillo, C. J. T.; Sinelnikov, R.; Thiessen, A. N.;

Washington, J.; Yu, H.; Veinot, J. G. C. Chem. Mater. [Online early access] 2016.

(55) Sugimoto, H.; Fujii, M.; Imakita, K. RSC Adv. 2014, 5 (11), 8427–8431.

(56) Kmecl, P.; Bukovec, P. Acta Chim. Slov. 1999, 46 (2), 161–171.

(57) Khelifi, R.; Mathiot, D.; Gupta, R.; Muller, D.; Roussel, M.; Duguay, S. Appl. Phys. Lett.

2013, 102 (1), 13116.

Page 59: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

46

(58) Bock, R.; Altermatt, P. P.; Schmidt, J.; Kessler, M. A.; Ohrdes, T.; Wolpensinger, B.;

Harder, N. Semicond. Sci. Technol. 2010, 25 (5), 55001.

(59) Arai, E.; Nakamura, H.; Terunuma, Y. J. Electrochem. Soc. 1973, 120 (7), 980–987.

(60) Cao, X.; Hamers, R. J. J. Phys. Chem. B 2002, 106 (8), 1840–1842.

(61) Reichenbächer, K.; Süss, H. I.; Hulliger, J. Chem. Soc. Rev. 2005, 34, 22–30.

(62) Fernández-García, A.; Zarza, E.; Valenzuela, L.; Pérez, M. Renewable Sustainable Energy

Rev. 2010, 14 (7), 1695–1721.

(63) Kang, Z.; Tsang, C. H. A.; Wong, N.-B.; Zhang, Z.; Lee, S.-T. J. Am. Chem. Soc. 2007,

129 (40), 12090–12091.

(64) Molecular Sieves - Technical Information Bulletin

http://www.sigmaaldrich.com/chemistry/chemical-synthesis/learning-center/technical-

bulletins/al-1430/molecular-sieves.html (accessed Nov 8, 2016).

Page 60: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

47

Appendix A: Powder X-ray Diffraction Pattern

Powder X-ray diffraction patterns of B-ncSi, P-ncSi, B,P-ncSi, and pristine ncSi heat treated at

400 °C for 30 min and then at 1100 °C for 2 h at a heating rate of 18 °C min-1 under 95% Ar/5%

H2 in a quartz tube furnace.

B-ncSi

Page 61: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

48

P-ncSi

B,P-ncSi

Page 62: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

49

ncSi

Page 63: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

50

Appendix B: Irradiance Spectra

Irradiance spectra of the 1000 W Hortilux Blue metal halide lamp, 300 W xenon lamp, and the

AM1.5 solar irradiance. The light from the 300 W xenon lamp was focused to reach an intensity

of approximately 50 Suns.

Page 64: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

51

Appendix C: Thermogravimetric Analysis

The weight gain from the introduction of CO2 is used for the calculation of CO2 adsorption

capacity (the weight percent difference between the plateaus of the red and pink curves). All of

the TGA were performed at 35 °C.

B-ncSi

Page 65: Enhancement of Hydride-Terminated Silicon Nanocrystals for Gas … · Annabelle Po Yin Wong Master of Science Department of Chemistry University of Toronto 2017 Abstract Utilization

52

P-ncSi

ncSi


Recommended