Laser Selective Photoactivation of Amorphous TiO2 Films to Anatase

Jul 23, 2018 - (1) Since the past decade, it also serves a wide variety of applications ... whereas conversion to rutile appears between 800 and 1100 ...
1 downloads 0 Views 2MB Size
Subscriber access provided by Kaohsiung Medical University

Article

Laser selective photo-activation of amorphous TiO2 films to anatase and/or rutile crystalline phases Jaime A Benavides, Charles Trudeau, Luis Felipe Gerlein, and Sylvain G. Cloutier ACS Appl. Energy Mater., Just Accepted Manuscript • DOI: 10.1021/acsaem.8b00171 • Publication Date (Web): 23 Jul 2018 Downloaded from http://pubs.acs.org on July 25, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 22 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

1

Laser selective photo-activation of amorphous TiO2

2

films to anatase and/or rutile crystalline phases

3

4

Jaime A. Benavides, Charles P. Trudeau, Luis F. Gerlein, Sylvain G. Cloutier*.

5

Département de génie électrique, École de Technologie Supérieure,

6

1100 Notre-Dame Ouest, Montréal, Québec H3C 1K3, Canada

7

E-mail: [email protected]

8

Keywords: Nanoparticles, TiO2, sol-gel, anatase, rutile, oxygen vacancies, laser-assisted

9

conversion, film patterning

10 11

Abstract

12

Titanium dioxide (TiO2) is a remarkable metal-oxide semiconductor with unique optoelectronic

13

properties ideal for photovoltaics and photocatalytic conversion. The principal crystalline phases

14

for TiO2 are anatase, rutile and brookite. The combination of both anatase and rutile crystalline

15

structures can positively impact the optoelectronic properties of TiO2 films. With standard sol-

ACS Paragon Plus Environment

1

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 2 of 22

1

gel processing, high-temperature conversion generally yields one dominant phase and limits the

2

combined use of anatase and rutile TiO2 for optoelectronic devices. We report on a singular route

3

to controllably-engineer hybrid nanocrystalline films of TiO2 at room temperature to

4

synergistically exploit both anatase and rutile TiO2 phases. Relying on sol-gel chemistry, this

5

approach starts from an amorphous film and uses photo-induced activation using a low-power

6

laser to achieve specific spatially-controlled pattern consisting of different TiO2 crystalline

7

phases within the same film. While achieving remarkable precision, reproducibility and control,

8

we also avoid costly high-temperature, ion metal-assisted or specific atmospheric processing that

9

currently prevents the integration of TiO2 in several optoelectronic platforms. In the future, we

10

believe this unprecedented level of control and the ability to engineer the TiO2 crystalline

11

structure at the microscopic scale will allow the design and fabrication of novel high-

12

performance TiO2 hybrids for energy conversion and environmental applications.

13 14 15

Introduction

16

oxidizing power, high photochemical corrosive resistance and low cost1. Since the last decade, it

17

also serves a wide variety of applications including sensing devices2, organic dye- and quantum

18

dot-sensitized solar cells3, photo-electrochromics4, water-splitting5 and photocatalysis6 for

19

extensive environmental applications, among others. All these applications of TiO2 stem from its

20

unique optoelectronic properties, which strongly depend on the crystalline structure. Moreover, it

21

is already established that synergistic effects from combining multiple crystalline structures can

22

have positive consequences on these properties7–9. It has proven challenging to synthesize a

23

particular crystalline phase while coexisting next to the other, considering the conditions

24

necessary to achieve each one of them and take advantage of these synergistic effects.

Titanium dioxide is widely used for its non-toxicity, chemical and biological stability, strong

ACS Paragon Plus Environment

2

Page 3 of 22 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

1

Thermal annealing between 450°C and 1100°C is generally required to achieve the desired

2

crystalline structure starting from amorphous TiO210. The anatase structure generally appears

3

around 450ºC, while conversion to rutile appears between 800°C to 1100°C11. Anatase TiO2

4

displays a tetragonal atomic structure with six atoms in each primitive cell. This phase is

5

energetically metastable in bulk TiO212. Its relatively wide 3.2 eV electronic bandgap makes it a

6

superb blocking layer for optoelectronic or photovoltaic devices13,14. In contrast, bulk rutile TiO2

7

being also tetragonal, possesses 12 atoms in its primitive cell but this phase displays higher

8

energetic equilibrium and it is the most common form of TiO2 found naturally. Rutile TiO2 has a

9

lower electronic bandgap, 3.0 eV, and is better suited for photoelectrochemical applications.

10

15

The Brookite phase is the least studied one because is difficult to synthesize pure16 and it is out

11

of the scope of this paper. Although rutile is acknowledged as the most stable phase for bulk

12

TiO2, anatase is more stable for TiO2 nanoparticles smaller than 50 nm17.

13

High-temperature treatment is also generally used to sinter nanocrystalline TiO2 particles and

14

form interconnected nanoparticle networks. This sintering process is generally performed around

15

500ºC18. Given the wide range of possibilities that TiO2 has to offer for the development of new

16

flexible, lightweight and wearable technologies requiring low-temperature substrates, researchers

17

have generally tried to crystallize or sinter the TiO2 before its integration in the devices. As such,

18

the high temperatures used to reach the anatase and rutile crystalline phases can also be most

19

detrimental to temperature-sensitive devices. Different approaches have been explored to produce TiO2-based devices at low temperatures18–

20 21

22

22

process requires a thin adhesion layer between the pre-sintered TiO2 film and the substrate,

23

followed by an application of high pressures and moderate temperatures18.

. For example, pre-sintered TiO2 porous layers can be used, but the relatively complex transfer

Alternative

ACS Paragon Plus Environment

3

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 22

1

approaches towards low-temperature device assembly at low temperature include compression

2

methods and the use of conductive inks or electrolytes between the TiO2 particles19–22. While

3

these approaches partially succeeded in avoiding high-temperature sintering, it remains that the

4

degree of crystallinity and the optoelectronic properties of the resulting TiO2 films are far from

5

ideal.

6

Nevertheless, achieving high-quality crystalline TiO2 films at room-temperature remains one

7

of today’s most important technical challenges towards low-cost optoelectronic devices

8

including photovoltaic cells and photocatalytic device architectures. With the advent of laser-

9

based particle sintering used for additive manufacturing, researchers have demonstrated

10

optically-induced phase conversion from anatase-to-amorphous TiO2 using low intensity visible

11

light under vacuum23. More recently, the same photo-activation process was reported under

12

oxygen-poor conditions24. Finally, it was also shown that Fe and Al-doped TiO2 nanoparticles

13

can respectively promote (Fe) or inhibit (Al) the anatase-to-rutile phase transition under UV-

14

laser irradiation, taking the process one step further by achieving micropatterning of rutile

15

phase25.

16

In this paper, we present a much simpler method for triggering the crystallization of a TiO2

17

sol-gel precursor at low-energies and achieve complete control of the crystalline phase at room-

18

temperature and in ambient air using low-power laser-induced photo-activation of an amorphous

19

TiO2 nanoparticle films. As we show, this well-controlled photo-activation can allow conversion

20

from amorphous-to-anatase, amorphous-to-rutile, amorphous-to-mixture of anatase/rutile or from

21

amorphous-to-anatase-to-rutile simply by adjusting the power density of the laser. This simple

22

process is easily scalable to industrial-like environments and it is perfectly compatible with

23

emerging laser-based 3D printing technologies.

ACS Paragon Plus Environment

4

Page 5 of 22 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

Figure 1. Characterization methods. a) TEM image of the pristine amorphous TiO2 nanoparticles as synthetized using a sol-gel route b) AFM image of a thin film of amorphous TiO2 nanoparticles produced by sol-gel chemistry. The average nanoparticle diameter is estimated around 10 nm. c) SEM image of the amorphous TiO2 films after focusing under SEM. Black arrows point out the

charging effect areas (where the electron beam was focused). White areas in the film correspond to crystallized areas where the focused electron beam triggered nucleation and transformation from amorphous to crystalline TiO2. 1 2 3

Fabrication of the patternable amorphous TiO2 nanoparticle films using sol-gel chemistry

4

First, a glass substrate is placed at the bottom of a beaker filled with the TiO2 nanoparticles

5

dispersion produced by sol-gel chemistry following the protocol described in the Methods

6

section. The beaker is then covered with a perforated parafilm membrane to allow slow

7

evaporation of the hexane solvent at room-temperature. When the solvent is completely

8

evaporated (approx. 24 hours), we remove the glass substrate with a 23µm thickness film of

9

amorphous TiO2 nanoparticles created over its surface. The amorphous nature of the TiO2

ACS Paragon Plus Environment

5

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 22

1

nanoparticle film shown in Figure 1 was confirmed by XRD and Raman spectroscopy26 (see

2

Figure S-1 in supplementary information).

3

During the SEM observation of this amorphous nanoparticle film, we observed that the

4

focused electron beam from the SEM was sufficient to initiate a nucleation and crystallization of

5

the amorphous TiO2 film (Fig. 1c, see also movie S-1). The whiter areas indicate where the

6

nucleation process occurred. From there, it spreads along the film following fractal patterns. This

7

was the first indicator to confirm that this protocol required very low activation energy compared

8

with previous results from the literature. Additionally, it was also noticed that thicker areas of the

9

TiO2 film, require an increase of the irradiation voltage from 5kV to 15 kV to trigger nucleation

10

under SEM. Unfortunately, it is not possible to obtain Raman signatures of those crystallized

11

features since they are too small for the spatial resolution available in our confocal Raman

12

microscope. To explore in more detail this phenomenon, we decided to test the behavior of the

13

amorphous TiO2 films under direct laser-induced photo-activation.

14 15

Patterning on-demand crystalline structure over amorphous TiO2 films

16 17

As the SEM results successfully demonstrated the very low activation energy required to

18

initiate the phase transformation, we used low-power laser-induced photoactivation to generate

19

mixed structures of anatase and/or rutile simply by controlling the power density of the laser. To

20

do so, we used a WITEC Alpha300 confocal Raman microscope equipped with a 60 mW fiber-

21

coupled continuous-wave laser at 532 nm and a mechanical attenuator. The source beam is then

22

coupled through a 10X objective for excitation of the sample, which is mounted on a motorized

23

high-precision XYZ stage.

ACS Paragon Plus Environment

6

Page 7 of 22 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

1

Figure 2. Spatially-resolved laser-induced photoactivation process. a) Schematics of the experimental setup used for laser-induced photo-activation of the TiO2 structure.

b) Laser scanning microscope image (LEXT OLS4100; Olympus) of the amorphous TiO2 films after crystalline phase alteration. The patterned letters ÉTS consist of anatase-phase crystalline TiO2, while the underline consists of rutile-phase crystalline

TiO2. The darker background consists of amorphous TiO2. The different crystalline phases are confirmed by Raman micro-spectroscopy measurements as shown in c) and d), respectively.

2 3

Our results shown in Figure 2 confirm we can controllably photo-activate the phase transition

4

from the amorphous TiO2 to the anatase and/or rutile phase at room-temperature and in ambient

5

environment simply by adjusting the excitation power density. A power density above 75W mm-

6

2

7

transition to rutile TiO2 can be completed using an incident power density above 445 W mm-2

8

(the horizontal line under “ÉTS” in Fig.2). Both crystalline phases are confirmed by the Raman

is required to complete the transition to anatase TiO2(the “ÉTS” letters in Fig.2), whereas the

ACS Paragon Plus Environment

7

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 8 of 22

1

micro-spectroscopy results in Figure 2(c,d). Indeed, Figure 3 shows high-resolution Raman

2

spectra of each crystalline structure obtained at these different laser power densities. When using

3

an intermediate power density (i.e. 275 W mm-2) we observe a mixture of both anatase and rutile

4

crystalline TiO2.

5

6

Figure 3. Raman spectra of the various TiO2 crystalline structures obtained using different

7

excitation laser power densities.

8

To understand the influence of the incident laser power density over the crystallization process,

9

we recorded the transient evolution of the Raman signature over a period of 1 minute after

10

opening the laser shutter at the 5 second mark. Figure 4 shows the laser-assisted crystallization

11

kinetics to reach anatase (Fig.4a,d), mixed anatase and rutile (Fig.4b,e) and rutile (Fig.4c,f)

12

phases. Here, it is possible to observe that the crystallization process for all three TiO2

13

polymorphs starts as soon as the amorphous film is exposed to the laser irradiation.

14

ACS Paragon Plus Environment

8

Page 9 of 22 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

1 2

Figure 4. Evolution of crystalline structure as function of time and Raman intensity. Figures

3

a), b) and c) correspond to the Raman spectra evolution for the transition from amorphous to (a)

4

pure anatase, (b) mixed anatase/rutile and (c) pure rutile observed at 75, 275 and 445 W mm-2

5

power densities respectively taken over 1 minute. The laser shutter opens at the 5-second mark.

6

The Raman peak intensity evolution for the dominant Raman peak is recorded over 1 hour for

7

the conversion to (d) anatase, (e) mixed anatase/rutile and (f) rutile. The dashed line marks the

8

time required to reach 90% of maximum Raman intensity.

9

For the pure anatase crystalline phase obtained using a 75 W mm-2 power density, the

10

identifying peak at 152cm-1 starts to appear during the first 6 seconds of laser exposure. The

11

other characteristic peaks of the anatase phase at 395cm-1, 513cm-1 and 634cm-1

12

defined after only 30 seconds of exposure. The other TiO2 crystalline phase transitions observed

13

exhibit similar behavior shown in Figure 4(d).

(27)

are well

14

In the mixed crystalline structure generated under 275 W mm-2 exposure, we observe well-

15

defined rutile characteristic peaks such as 152cm-1, 261cm-1, 420cm-1 and 610cm-1 instantly after

16

laser exposure. As expected, the intensity of all these peaks increases over time but, contrary to

ACS Paragon Plus Environment

9

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 22

1

the characteristic peak ratios present in rutile, the peak at 152cm-1 still dominates, suggesting

2

incomplete transition from anatase to rutile. Indeed, the transient evolution of the 152cm-1 peak

3

behaves similarly to what can be observed for anatase, indicating the presence of a mixed,

4

anatase and rutile, crystalline phase achieved at this power density. Figure 4(e), shows the

5

evolution of the 152cm-1 rutile peak, which rapidly increases upon laser exposure before

6

reaching saturation. On the other hand, the mix anatase/rutile phase behaves more as a

7

transitional state, where contributions of both phases can be perceived simultaneously. As such,

8

the mixture phase displays both strong 152 cm-1 (from anatase) and 420 cm-1 (from rutile) peaks.

9

With higher portion of anatase phase dominating the Raman spectra, it is possible that the peak

10

at 420 cm-1 displays different kinetics due to the contribution of the adjacent anatase peaks and

11

displays its distorted and irregular behavior.

12

For the rutile crystalline structure generated at higher power densities, it is possible to observe

13

in Figure 4(f) the well-defined rutile peaks at 152cm-1, 261cm-1, 420cm-1 and 610cm-1

(28)

14

immediately after laser exposure. The Figure 4(f) depicts the evolution of the dominant 420cm-1

15

rutile peak.

16

In all cases, we notice little changes in the intensity of the Raman signal after the first 600

17

seconds of exposure, suggesting that conversion is complete after a short exposure time and that

18

the process is not cumulative over time.

19

To study in detail the patterning process and its effect on the amorphous TiO2 film structure,

20

we decided to pattern both anatase and rutile crystalline structures in a checkered pattern. The

21

dominant crystalline phase is confirmed by performing a Raman mapping of the photo-activated

22

area (Figure 5). Again, each transition, from amorphous to anatase or rutile is selectively

23

achieved by controlling the laser power density. It is important to note that patterning velocity is

ACS Paragon Plus Environment

10

Page 11 of 22 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

1

not uniform during the process since all the patterns that are shown in this work are produced by

2

moving manually the XY positioner using the joystick. Nevertheless, we are certain that

3

automation of this process will yield excellent outcomes. Since laser exposure induces crystalline

4

conversion in the amorphous TiO2 films, the excitation power density within the confocal Raman

5

system is kept below 75 W mm-2 for the acquisition of the Raman data.

6

Anatase areas show a well-defined shape. Trace A (Fig. 5a) corresponds to the path left by the

7

laser during the anatase patterning process. This path is narrow with approximately 9.3 µm of

8

width. Rutile areas are also well defined but the crystalized area is bigger than intended. This

9

comes as a result of the wider trace left by the laser during the rutile pattering process. The laser

10

spot does not change while increasing the power density but the affected area in the film results

11

larger. Indeed, wider trace R (Fig. 5a) has a width of approximately 17.2 µm.

12

13 14

Figure 5. Patterning on-demand crystalline structure. a) Laser scanning microscope image

15

(LEXT OLS4100; Olympus) of checkered pattern over amorphous TiO2 films. Anatase TiO2 is

16

generated in squares 1 and 2, while rutile TiO2 is generated in the squares 3 and 4. b) Micro-

17

Raman image of the pattern. The blue and red colors represent anatase and rutile TiO2

18

respectively. c) Topographic 3D surface reconstructions for the checkered pattern. 3D image

ACS Paragon Plus Environment

11

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 22

1

generated using software ImageJ based on the data given by the laser scanning microscope

2

analysis.

3 4

The visible difference between traces A and R in Figure 5a. arises from the different values of

5

power density required to crystalize each phase; i.e. an increment in power density during the

6

patterning process translated in an increment in the width of the trace left by the laser.

7

It is possible to appreciate a clear contrast difference between the amorphous area and the

8

crystallized regions in the patterned sample (Figure 5a). In the crystallized regions, the color

9

changes from dark to light gray and cracks can be clearly observed in all crystallized regions,

10

especially in the rutile TiO2 regions. These cracks appear as a direct consequence of the natural

11

densification of the TiO2 during the crystallization process. Indeed, the atoms reorganize during

12

the laser-induced photo-activation process. This reorganization at the atomic level results in an

13

important reduction in volume due to a difference in packing density for each phase. Indeed, the

14

density values reported in the literature are 3830 kg m-3 and 4240 kg m-3 for anatase and rutile

15

respectively29. Such reduction generates cracks in the photo-converted regions. Cracks observed

16

in the rutile TiO2 regions are more pronounced than cracks formed in anatase regions. This

17

contrast is explained by the smaller unit-cell volume for rutile TiO2 and higher density due to the

18

increased number of atoms. As shown in the Table S-1 of supplementary information, the

19

volume of the anatase unit-cell is more than twice that of rutile. In other words, during rutile

20

patterning, atoms get closer and the reorganization of its lattice generates more shrinkage,

21

resulting in larger and deeper cracks when the material crystallizes. This loss of volume on the

22

material becomes more obvious when we analyze the structure using a 3D topographic surface

23

reconstruction of the checkered pattern. In Figure 5c, it is noticeable that the anatase tiles are

ACS Paragon Plus Environment

12

Page 13 of 22 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

1

shallower, smoother and more regular than rutile tiles. As expected, rutile regions also exhibit

2

deeper cracks as stronger densification of the TiO2 occurs in rutile compared with anatase.

3

Phase stability within nanocrystalline TiO2 is notorious for its great dependence on the particle

4

size30,31. Under 14 nm, it is possible to achieve conversion to anatase at low-temperatures,

5

starting from amorphous TiO2 nanoparticles32,33. This explains the results seen here, since the

6

size measured for the amorphous TiO2 nanoparticles synthetized are around 10 nm. We believe

7

that localized thermal effects caused by the focused laser are responsible for the crystalline phase

8

transition.

9

In this work the irradiation source is a 532nm laser, whilst smaller than the 3.2 eV bandgap of

10

TiO2, the absorption of visible light is attributed to the presence of oxygen vacancies34 generated

11

during the synthesis of these TiO2 nanoparticles. Additionally, phase transition from anatase to

12

rutile can be induced by light irradiation in rich oxygen atmospheres23,35. The phase transition

13

phenomenon relies on the ability of the material to absorb or desorb molecular oxygen29. In other

14

words, the high concentration of molecular oxygen facilitates the phase transition due to the

15

ionic mobility created in oxygen vacancies. Indeed, high concentration of oxygen vacancies

16

creates band states within the TiO2 bandgap which effectively allows the absorption of the 532

17

nm laser source while raising the energy of the system36. Also, it promotes the lattice relaxation,

18

helping with the rearrangement process of the ions that participate in the phase transition.25,37

ACS Paragon Plus Environment

13

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 22

1 2

Figure 6. Proposed Jablonski diagram for the laser-induced phase transition from anatase to

3

rutile TiO2. The band states make possible the phase transition under visible light.

4

In our particular case, the oxygen molecules act as very efficient photoexcited electron

5

scavengers, trapping the excited electrons from the conduction band into the surface states of the

6

TiO238. The oxygen molecules are then adsorbed at the surface of the TiO2 nanoparticles to

7

partially compensate these oxygen vacancies39. The presence of oxygen vacancies promotes the

8

formation of Ti3+ sites in the crystal structure as the electrons left behind by the vacancy are

9

distributed on neighboring Ti sites, reducing them from Ti4+ to Ti3+.38,40,41 At this point, assisted

10

by the continuous laser irradiation, the adsorbed oxygen molecule passivates the TiO2 by

11

bridging the metallic ions42. As result of the phase transition and the relaxation of the electrons, a

12

vibrational phonon is emitted as heat (Figure 6).

13 14

Conclusions

ACS Paragon Plus Environment

14

Page 15 of 22 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

1

In this paper, we report low-power laser-assisted photoactivation of TiO2 films performed at

2

room-temperature in ambient environment. This simple approach allows to selectively and

3

controllably generate conversion anatase and/or rutile crystalline phases simply by controlling

4

the laser power density incident on a film of amorphous TiO2 nanoparticles synthesized through

5

an incomplete nonhydrolytic sol-gel ester elimination reaction of titanium isopropoxide and oleic

6

acid.

7

Raman micro-spectroscopy is used to confirm the transition to pure anatase, mixed

8

anatase/rutile and pure rutile crystalline phases generated using different laser power densities.

9

Transient evolutions are also measured using time-resolved Raman micro-spectroscopy. The

10

results show that crystallization occurs in the very first seconds of irradiation, that this effect is

11

permanent and non-accumulative. Laser scanning microscope images and 3D topographic

12

surface reconstruction of the photo-activated areas evidence the loss in volume after

13

crystallization due to denser atomistic reorganization after phase transition. The method

14

presented in this work can be used to spatially-organize different crystalline phases with high

15

level of precision to study new synergistic effects7,8. Our results are consistent with the literature

16

suggesting that amorphous TiO2 nanoparticles under 14 nm allow laser-assisted phase transition

17

to anatase. Regarding the subsequent transition from anatase to rutile, we confirm it relies on the

18

high capacity of the synthetized TiO2 to absorb molecular oxygen23. These results are consistent

19

with the state-of-the-art. Most importantly, all the phase transitions presented on this work are

20

achieved without requiring any kind of dopant in the TiO2 film prior photoactivation.

21

Our confocal Raman system allows for scan speeds up-to 4000µm2 s-1 while allowing phase

22

transitions. This potentially allows larger scale fabrication confined within the size constrains of

23

our system, 25 x25 mm2. Nonetheless, both phases can be patterned using this area for a usable

ACS Paragon Plus Environment

15

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 16 of 22

1

device that can be readily used in optoelectronic devices. Other industrial systems that allow for

2

larger scale and faster fabrication are possible as long as the radiation power density condition is

3

met.

4

Throughout this paper, we have discussed the importance of the easy integration of TiO2 in

5

many photo-electro-catalytic applications. We believe this accurate, reproducible and scalable

6

method of producing crystalline TiO2 at the microscopic level opens the door to novel pathways

7

of TiO2-based platforms, especially for energy conversion and environmental applications.

8 9 10 11

Methods Chemicals. Oleic acid (90%, Aldrich) (OA), titanium (IV) isopropoxide (TTIP, Aldrich, 97%), ethanol (ACS reagent, ≥99.5% (200 proof, absolute) and hexane.

12

Synthesis of TiO2 nanoparticles. TiO2 nanoparticles are synthesized through an incomplete

13

nonhydrolytic sol-gel ester elimination reaction of titanium isopropoxide (TTIP) and oleic acid

14

(OA)43. TTIP (3.36g) is added to OA (10 g) at room temperature under nitrogen atmosphere. The

15

resulting mixture is heated to 280˚C in a period of 20 min (14˚C min−1) and it is kept at this

16

temperature for 2h. The light-yellow solution gradually turns dark brown. At this point, the

17

solution is cooled down to room temperature and ethanol in excess is added to yield a beige

18

precipitate. The solution is centrifuged for 30 min to recuperate the nanoparticles. Finally, the

19

solution is re-dispersed in hexane and ready to use. The TiO2 nanoparticles obtained from this

20

synthesis are shown in Figure 1a), from the AFM image Figure 1b) an average particle diameter

21

of 10.5 nm was calculated.

22

ACS Paragon Plus Environment

16

Page 17 of 22 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1 2

ACS Applied Energy Materials

Corresponding Author *E-mail: [email protected] (S.G.C.)

3 4

Authors contribution

5

J.B. performed both, the concept and the synthesis of the TiO2 patternable films. C.T.

6

performed the Raman measurements, analysis and interpretation of Raman results. L.F.G.

7

performed the 3D reconstruction of the sample and the time evolution analysis. Interpretations of

8

the results and data analysis were discussed between all the authors. The first draft of the

9

manuscript was written by J.B. and revised by C.T., L.F.G. and S.G.C. All authors have

10

approved the final version of the manuscript before its submission.

11 12

Acknowledgements

13

Sylvain G. Cloutier thanks the Canada Research Chair and the NSERC Discovery programs

14

for their support.

15

Funding Sources

16

S.G.C. is most thankful for the financial support from the Canada Research Chairs (Award

17

231814) and NSERC (Award 491592).

18

Abbreviations

19

TiO2, titanium dioxide; OA, oleic acid; TTIP, titanium (IV) isopropoxide.

ACS Paragon Plus Environment

17

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 22

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43

References (1) Gupta, S. M.; Tripathi, M. A Review of TiO2 Nanoparticles. Chin. Sci. Bull. 2011, 56 (16), 1639. (2) Simić, M.; Manjakkal, L.; Zaraska, K.; Stojanović, G. M.; Dahiya, R. TiO2-Based Thick Film PH Sensor. IEEE Sens. J. 2017, 17 (2), 248–255. (3) McDonald, S. A.; Konstantatos, G.; Zhang, S.; Cyr, P. W.; Klem, E. J. D.; Levina, L.; Sargent, E. H. Solution-Processed PbS Quantum Dot Infrared Photodetectors and Photovoltaics. Nat. Mater. 2005, 4 (2), 138–142. (4) Wei, N.; Cui, H.; Wang, C.; Zhang, G.; Song, Q.; Sun, W.; Song, X.; Sun, M.; Tian, J. Bi2O3 Nanoparticles Incorporated Porous TiO2 Films as an Effective P-n Junction with Enhanced Photocatalytic Activity. J. Am. Ceram. Soc. 2017, 100 (4), 1339–1349. (5) Guo, L.; Yang, Z.; Marcus, K.; Li, Z.; Luo, B.; Zhou, L.; Wang, X.; Du, Y.; Yang, Y. MoS2/TiO2 Heterostructures as Nonmetal Plasmonic Photocatalysts for Highly Efficient Hydrogen Evolution. Energy Environ. Sci. 2017. (6) Mueses, M. A.; Machuca-Martinez, F.; Li Puma, G. Effective Quantum Yield and Reaction Rate Model for Evaluation of Photocatalytic Degradation of Water Contaminants in Heterogeneous Pilot-Scale Solar Photoreactors. Chem. Eng. J. 2013, 215–216, 937–947. (7) Kafizas, A.; Carmalt, C. J.; Parkin, I. P. Does a Photocatalytic Synergy in an Anatase–Rutile TiO2 Composite Thin-Film Exist? Chem. – Eur. J. 2012, 18 (41), 13048–13058. (8) Li, G.; Richter, C. P.; Milot, R. L.; Cai, L.; Schmuttenmaer, C. A.; Crabtree, R. H.; Brudvig, G. W.; Batista, V. S. Synergistic Effect between Anatase and Rutile TiO2 Nanoparticles in Dye-Sensitized Solar Cells. Dalton Trans. 2009, No. 45, 10078–10085. (9) Ohno, T.; Tokieda, K.; Higashida, S.; Matsumura, M. Synergism between Rutile and Anatase TiO2 Particles in Photocatalytic Oxidation of Naphthalene. Appl. Catal. Gen. 2003, 244 (2), 383–391. (10) Krylova, G.; Na, C. Photoinduced Crystallization and Activation of Amorphous Titanium Dioxide. J. Phys. Chem. C 2015, 119 (22), 12400–12407. (11) Sun, Y.; Egawa, T.; Zhang, L.; Yao, X. High Anatase-Rutile Transformation Temperature of Anatase Titania Nanoparticles Prepared by Metalorganic Chemical Vapor Deposition. Jpn. J. Appl. Phys. 2002, 41 (8B), L945. (12) Satoh, N.; Nakashima, T.; Yamamoto, K. Metastability of Anatase: Size Dependent and Irreversible Anatase-Rutile Phase Transition in Atomic-Level Precise Titania. Sci. Rep. 2013, 3. (13) Hanaor, D. A. H.; Sorrell, C. C. Review of the Anatase to Rutile Phase Transformation. J. Mater. Sci. 2010, 46 (4), 855–874. (14) Kavan, L.; Zukalova, M.; Vik, O.; Havlicek, D. Sol-Gel Titanium Dioxide Blocking Layers for Dye-Sensitized Solar Cells: Electrochemical Characterization. Chemphyschem Eur. J. Chem. Phys. Phys. Chem. 2014, 15 (6), 1056–1061. (15) Mo, S.-D.; Ching, W. Y. Electronic and Optical Properties of Three Phases of Titanium Dioxide: Rutile, Anatase, and Brookite. Phys. Rev. B 1995, 51 (19), 13023–13032. (16) Di Paola, A.; Bellardita, M.; Palmisano, L. Brookite, the Least Known TiO2 Photocatalyst. Catalysts 2013, 3 (1), 36–73.

ACS Paragon Plus Environment

18

Page 19 of 22 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45

ACS Applied Energy Materials

(17) Hwu, Y.; Yao, Y. D.; Cheng, N. F.; Tung, C. Y.; Lin, H. M. X-Ray Absorption of Nanocrystal TiO2. Nanostructured Mater. 1997, 9 (1), 355–358. (18) Dürr, M.; Schmid, A.; Obermaier, M.; Rosselli, S.; Yasuda, A.; Nelles, G. LowTemperature Fabrication of Dye-Sensitized Solar Cells by Transfer of Composite Porous Layers. Nat. Mater. 2005, 4 (8), 607–611. (19) Haque, S. A.; Palomares, E.; Upadhyaya, H. M.; Otley, L.; Potter, R. J.; Holmes, A. B.; Durrant, J. R. Flexible Dye Sensitised Nanocrystalline Semiconductor Solar Cells. Chem. Commun. 2003, No. 24, 3008–3009. (20) Lindström, H.; Magnusson, E.; Holmberg, A.; Södergren, S.; Lindquist, S.-E.; Hagfeldt, A. A New Method for Manufacturing Nanostructured Electrodes on Glass Substrates. Sol. Energy Mater. Sol. Cells 2002, 73 (1), 91–101. (21) Boschloo, G.; Lindström, H.; Magnusson, E.; Holmberg, A.; Hagfeldt, A. Optimization of Dye-Sensitized Solar Cells Prepared by Compression Method. J. Photochem. Photobiol. Chem. 2002, 148 (1–3), 11–15. (22) Kado, T.; Yamaguchi, M.; Yamada, Y.; Hayase, S. Low Temperature Preparation of Nano-Porous TiO2 Layers for Plastic Dye Sensitized Solar Cells. Chem. Lett. 2003, 32 (11), 1056–1057. (23) Ricci, P. C.; Casu, A.; Salis, M.; Corpino, R.; Anedda, A. Optically Controlled Phase Variation of TiO2 Nanoparticles. J. Phys. Chem. C 2010, 114 (34), 14441–14445. (24) Ricci, P. C.; Carbonaro, C. M.; Stagi, L.; Salis, M.; Casu, A.; Enzo, S.; Delogu, F. Anatase-to-Rutile Phase Transition in TiO2 Nanoparticles Irradiated by Visible Light. J. Phys. Chem. C 2013, 117 (15), 7850–7857. (25) Vásquez, G. C.; Peche-Herrero, M. A.; Maestre, D.; Gianoncelli, A.; Ramírez-Castellanos, J.; Cremades, A.; González-Calbet, J. M.; Piqueras, J. Laser-Induced Anatase-to-Rutile Transition in TiO2 Nanoparticles: Promotion and Inhibition Effects by Fe and Al Doping and Achievement of Micropatterning. J. Phys. Chem. C 2015, 119 (21), 11965–11974. (26) Huang, A. P.; Di, Z. F.; Chu, P. K. Microstructure and Visible-Photoluminescence of Titanium Dioxide Thin Films Fabricated by Dual Cathodic Arc and Nitrogen Plasma Deposition. Surf. Coat. Technol. 2007, 201 (9–11), 4897–4900. (27) Zeng, G.; Li, K.-K.; Yang, H.-G.; Zhang, Y.-H. Micro-Raman Mapping on an Anatase TiO2 Single Crystal with a Large Percentage of Reactive (001) Facets. Vib. Spectrosc. 2013, 68, 279–284. (28) Li, L.; Yan, J.; Wang, T.; Zhao, Z.-J.; Zhang, J.; Gong, J.; Guan, N. Sub-10 Nm Rutile Titanium Dioxide Nanoparticles for Efficient Visible-Light-Driven Photocatalytic Hydrogen Production. Nat. Commun. 2015, 6, 5881. (29) Diebold, U. The Surface Science of Titanium Dioxide. Surf. Sci. Rep. 2003, 48 (5), 53– 229. (30) Zhang, H.; Banfield, J. F. Kinetics of Crystallization and Crystal Growth of Nanocrystalline Anatase in Nanometer-Sized Amorphous Titania. Chem. Mater. 2002, 14 (10), 4145–4154. (31) Madras, G.; McCoy, B. J. Kinetic Model for Transformation from Nanosized Amorphous TiO2 to Anatase. Cryst. Growth Des. 2007, 7 (2), 250–253. (32) Ma, H. L.; Yang, J. Y.; Dai, Y.; Zhang, Y. B.; Lu, B.; Ma, G. H. Raman Study of Phase Transformation of TiO2 Rutile Single Crystal Irradiated by Infrared Femtosecond Laser. Appl. Surf. Sci. 2007, 253 (18), 7497–7500.

ACS Paragon Plus Environment

19

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35

Page 20 of 22

(33) Reyes-Coronado, D.; Rodríguez-Gattorno, G.; Espinosa-Pesqueira, M. E.; Cab, C.; Coss, R. de; Oskam, G. Phase-Pure TiO 2 Nanoparticles: Anatase, Brookite and Rutile. Nanotechnology 2008, 19 (14), 145605. (34) Emeline, A. V.; Kuznetsov, V. N.; Rybchuk, V. K.; Serpone, N. Visible-Light-Active Titania Photocatalysts: The Case of N-Doped s—Properties and Some Fundamental Issues. Int. J. Photoenergy 2008, 2008, 19. (35) Marinescu, C.; Sofronia, A.; Rusti, C.; Piticescu, R.; Badilita, V.; Vasile, E.; Baies, R.; Tanasescu, S. DSC Investigation of Nanocrystalline TiO2 Powder. J. Therm. Anal. Calorim. 2011, 103 (1), 49–57. (36) Choudhury, B.; Choudhury, A. Oxygen Defect Dependent Variation of Band Gap, Urbach Energy and Luminescence Property of Anatase, Anatase–Rutile Mixed Phase and of Rutile Phases of TiO2 Nanoparticles. Phys. E Low-Dimens. Syst. Nanostructures 2014, 56, 364–371. (37) Koparde, V. N.; Cummings, P. T. Phase Transformations during Sintering of Titania Nanoparticles. ACS Nano 2008, 2 (8), 1620–1624. (38) Komaguchi, K.; Maruoka, T.; Nakano, H.; Imae, I.; Ooyama, Y.; Harima, Y. ElectronTransfer Reaction of Oxygen Species on TiO2 Nanoparticles Induced by Sub-Band-Gap Illumination. J. Phys. Chem. C 2010, 114 (2), 1240–1245. (39) Komaguchi, K.; Nakano, H.; Araki, A.; Harima, Y. Photoinduced Electron Transfer from Anatase to Rutile in Partially Reduced TiO2 (P-25) Nanoparticles: An ESR Study. Chem. Phys. Lett. 2006, 428 (4), 338–342. (40) Nakaoka, Y.; Nosaka, Y. ESR Investigation into the Effects of Heat Treatment and Crystal Structure on Radicals Produced over Irradiated TiO2 Powder. J. Photochem. Photobiol. Chem. 1997, 110 (3), 299–305. (41) DeSario, P. A.; Chen, L.; Graham, M. E.; Gray, K. A. Effect of Oxygen Deficiency on the Photoresponse and Reactivity of Mixed Phase Titania Thin Films. J. Vac. Sci. Technol. A 2011, 29 (3), 031508. (42) Stagi, L.; Carbonaro, C. M.; Corpino, R.; Chiriu, D.; Ricci, P. C. Light Induced TiO2 Phase Transformation: Correlation with Luminescent Surface Defects. Phys. Status Solidi B 2015, 252 (1), 124–129. (43) Joo, J.; Kwon, S. G.; Yu, T.; Cho, M.; Lee, J.; Yoon, J.; Hyeon, T. Large-Scale Synthesis of TiO2 Nanorods via Nonhydrolytic Sol−Gel Ester Elimination Reaction and Their Application to Photocatalytic Inactivation of E. Coli. J. Phys. Chem. B 2005, 109 (32), 15297–15302.

36

Competing interests

37

A patent application will be filed in the coming weeks and the content of this article draft shall

38

be reviewed under confidentiality.

39

ACS Paragon Plus Environment

20

Page 21 of 22 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

ACS Applied Energy Materials

1

Competing financial interests

2

The authors declare that they have no competing financial interests.

3

ACS Paragon Plus Environment

21

ACS Applied Energy Materials 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

1

Page 22 of 22

GRAPHIC FOR TABLE OF CONTENTS ONLY

2 3

ACS Paragon Plus Environment

22