• 沒有找到結果。

PEC hydrogen generation based on nanomaterial photoelectrodes

Chapter 2 Literature Review

2.3 Photoelectrochemical water splitting

2.3.1 PEC hydrogen generation based on nanomaterial photoelectrodes

The use of nanomaterial-based PEC for water splitting dates back to 1997.

Fitzmaurice and coworkers reported the studies of charge separation in a

nanostructured TiO2 membrane sensitized with Ru complexes.[44] The long-lived

charge separation observed was an important finding that suggests that water

splitting is practically feasible on nanostructured TiO2 material. In the following few

years, Khan and Akikusa reported the photoresponse of nanocrystalline n-TiO2,

photoanodes are stable for water splitting. Significantly, the nanocrystalline n-Fe2O3

films showed higher photoresponse compared to those prepared by compression of

n-Fe2O3 powder or by thermal oxidation of metallic Fe sheets, indicating that

high-quality nanostructured materials could improve the overall efficiency of the

water splitting reaction. Since then, the PEC performance of different

nanocrystalline metal oxide films has been studied.[26]

Besides 0D nanostructures, 1D nanostructures, such as nanotubes, nanowires, and

nanorods, are expected to exhibit much improved transport properties than

nanoparticles. The first demonstration of PEC water splitting using 1D nanostructure

as photoelectrodes was reported by Lindquist and coworkers in 2000.[47,48] They

reported the photoelectrochemistry of hematite nanorod arrays for the photoanode. It

is generally accepted that recombination of electrons and holes, trapping of electrons

by oxygen deficiency sites, and low mobility of the holes cause the low

photoresponse for hematite films. In comparison to nanoparticles, nanorods improve

the transportation of carriers and thus reduce the recombination losses at grain

boundaries, as illustrated in Figure 2.11a. Moreover, the small diameter of the

nanorods minimizes the distance for holes to diffuse to the SC/electrolyte interface,

as shown in Figure 2.11b. Ideally, when the nanorod radius is smaller than the hole

well beyond the efficiencies reported for polycrystalline and nanocrystalline films of

hematite. This work demonstrated that nanorod arrays could potentially address

some of the fundamental PEC issues and increase the photon-to-current yield of

hematite.

Figure 2.1 Schematic operating principles of DMFC.[27]

Figure 2.2 Breakdown of anode, cathode, and electrolyte-related performance

losses in a DMFC.[29]

Figure 2.3 Schematic illustration of the synthesis procedure of the composite.[30]

Figure 2.4 TEM images of Pt/CNT (inset: enlarged image).[30]

Figure 2.5 TEM images of 20 wt% PtRu/C nanocatalyst prepared with different

molar ratios of Pt:Ru (a) LRTEM of 1:1, (b) HRTEM of 1:1.[31]

Figure 2.6 Typical TEM image of the CNT/PyPBI/Pt. Pt nanoparticles are loaded

Table 2.1 Reactor temperature ranges for initial processing of different fuels and

CO content of different process streams after the initial reaction.[11]

Figure 2.7 Schematic diagram of the integrated methanol steam reformer

system.[11,38]

Figure 2.8 Schematic of a typical PEC device and its basic operation mechanism

for hydrogen generation from water splitting.[40]

Figure 2.9 Energy diagram of a PEC cell consisting of an n-type SC photoanode

and a metal cathode for water splitting.[40]

Figure 2.10 Type of photoelectrode for PEC water splitting (SC-semiconductor;

M-metal).[21]

Figure 2.11 (a) Schematic representation of the electron transport through

spherical particles and nanorods. (b) A schematic of a Fe2O3 photoanode for water

splitting. The small dismeter of the nanowires ensures a short hole diffusion

length.[40,48]

1D & 3D nanoarchitectures

Carbon nanotube ZnO nanorod ZnO-CuO

inverse opal

Electrochemicalmeasurement

1D & 3D nanoarchitectures

Carbon nanotube ZnO nanorod ZnO-CuO

inverse opal

Electrochemicalmeasurement

Chapter 3

Experiment Methods

3.1 Flowchart of experiment process

3.2 Preparation of nanomaterials

3.2.1 Synthesis of CN

x

NTs and Pt NPs deposition

For the synthesis of the CNx NTs, an iron catalyst layer was deposited on Ti/Si

substrates by ion beam sputtering prior to the NT growth step. Then CNx NTs were

grown on the precoated substrates by MPECVD method (Figure 3.1), which has

been reported in our previous paper.[7] For Pt NPs deposition, a potential of -0.1 V

vs Ag/AgCl in 0.5M H2SO4 & 0.0025M H2PtCl6 mixture solutions was performed

for 15 sec as shown in Figure 3.2. Other details for the electrochemical deposition

process are reported by Quinn and co-workers.[65]

3.2.2 Fabrication of ZnO NRs and metallic Cu NPs deposition

For ZnO array preparation, a thin film of ZnO was first deposited on the inner

surface of microchannels prior to the NR growth step by solution method, which

acted as a seed layer. This was followed by growth of aligned ZnO NRs on the

precoated substrates by chemical bath deposition (CBD) method. The CBD growth

was performed using equimolecular mixtures of zinc nitrate hexahydrate (99.5 %,

Aldrich) and hexamethylenetetramine (99 %, Aldrich) as source precursors and a

reaction temperature of 90 oC for 7 h (Figure 3.3). Other details of the NR array

growth process were same as those described in the procedure proposed by

precursor salts (copper nitrate trihydrate, 99.5 %, Aldrich) that were subsequently

reduced (Figure 3.4). This reducing step is the boiling under reflux, which leads to

Cu/Cu+ colloids. The synthesis of size-selected Cu NPs immobilized on ZnO NRs

was carried out in ethylene glycol (EG; as reducing agent) solutions containing

different amounts of sodium hydroxide (0.5 M NaOH in EG) with pH value between

9 and 11.[85] The decoration concentrations were controlled between 0.5 and 3 mM.

The solutions were stirred for 30 min at room temperature, subsequently heated

under reflux to 190 ℃ for 2 h, and then cooled in air. Dark brown solutions

containing Cu NPs were formed in this manner, referred to as colloidal solutions in

this work.

3.2.3 Synthesis of microwave-activated CuO NT/ZnO NR nanocomposites

For synthesis of ZnO nanorods inside the microchannels, chemical bath deposition

growth was performed using the equimolecular mixtures of zinc nitrate hexahydrate

(99.5 %, Aldrich) and hexamethylenetetramine (99 %, Aldrich) as source precursors,

at a reaction temperature of 90 oC for 7 h. Other details for the NR array growth

process were similar to the procedure proposed by Vayssieres.[84] Typically, the

one-step direct impregnation method involved impregnation of ZnO

mixing Cu(NO3)2‧3H2O (99.5 %, Aldrich) in deionized water (18.2 MΩ).

Cu(NH3)42+ complex cations were prepared by adding a concentrated ammonia

solution (28-30 wt %) dropwise into the above aqueous solution until the pH value

reached 11 under vigorous stirring. The resulting homogeneous solutions were then

gently heated in an oven at 90 oC for over 2 h. Finally, CuO nanotip/ZnO nanorod

nanoarchitectures on the inner surface of microchannels were formed.

The as-derived CuO nanotip/ZnO nanorod catalyst precursors were processed in a

microwave oven (2.45 GHz; TMO-17MA; TATUNG CO., Ltd.; Taiwan) for 10 min

to obtain the final microwave-irradiated CuO nanotip/ZnO nanorod catalyst

precursors. The microwave chamber was 452 mm in length, 262 mm in width, and

325 mm in height. During the microwave irradiation, the substrate temperature,

environmental atmosphere and microwave power were maintained at room

temperature, air and 600 W, respectively. For comparison, the as-derived CuO

nanotip/ZnO nanorod catalyst precursors were also treated in a conventional

annealing furnace at 450 oC for 1 h to obtain the conventional thermal-treated CuO

nanotip/ZnO nanorod catalyst precursors

3.2.4 Preparation of O

2

plasma-activated CuO-ZnO inverse opals

For synthesis of CuO-ZnO inverse opals, 10 wt % of polystyrene (PS) colloidal

well-defined structures were formed inside the microchannels. These PS

particle-packed microchannels were heated at 90 ℃ for 3-6 h in order to improve

the connectivity between the neighboring particles. The copper-zinc precursor

solution containing 0.5 M Cu(NO3)2‧3H2O (99.5%, Aldrich) and 0.5 M Zn(NO3)2

6H2O (98%, Aldrich) in ethanol was then applied dropwise over the surface of the

PS layer. These infiltrated samples were then placed in air at room temperature for

2-4 h. Finally, the resulting CuZnO/PS composites were heated in air to 300 ℃ to

obtain the CuO-ZnO inverse opals (Figure 3.6).

These pristine CuO-ZnO inverse opals were then exposed to O2 plasma for a short

duration of 3-15 min to obtain O2-plasma treated CuO-ZnO inverse opals. The

process of plasma-chemical surface modification was performed in a parallel-plate

reactor with a DC flowing discharge (PCD-150; ALL REAL TECHNOLOGY CO.,

Ltd.; Taiwan). The distance between ground electrode and powered electrode was

about 6 cm. The plasma chamber was 250 mm in diameter and 140 mm in height.

The electrode was in the form of disk with around 6 inch in size. For the O2 plasma

treatment, the substrate temperature, total gas pressure and DC power were

maintained at room temperature, 200 mTorr and 50 W, respectively.

3.2.5 Preparation of carbon-modified ZnO inverse opals

After room temperature evaporation of water from the suspension, PS opals with

well-defined structures were formed on the ITO substrate. These PS particle-packed

templates were heated at 90 ℃ for 3-6 h in order to improve the connectivity

between the neighboring particles. The zinc precursor solution containing 0.5 M

Zn(NO3)2‧6H2O (98%, Aldrich) in ethanol was then applied dropwise over the

surface of the PS layer. These infiltrated samples were then placed in air at room

temperature for 2-4 h. Finally, the resulting ZnO/PS composites were heated in air to

300 ℃ to obtain the carbon-modified ZnO inverse opals. For comparison, the ZnO

noninverse opals could be prepared by the same fabrication procedure only without

using PS particle-packed templates.

3.3 Experimental measurements and characteristics analysis

For methanol reforming reaction, the Al-alloy (6061) chip of the microreactor was

made by ourselves through a laser machining method. Ten microchannels per

Al-alloy chip were separated by 800 μm fins. The width, depth and length of the

microchannels were 500μm, 200μm and 4.3 cm, respectively. For the calculation of

the catalyst weight for the steam reforming reactions, we measure the total weight

by precision electronic balance (± 0.1 mg; Sartorius). After reduction of catalysts in

a H2/N2 (5/95) at a flow rate of 50 mL min-1 at 200 ℃ for 1 hr, premixed water,

gas streams were analyzed online with a thermal-conductivity-detector gas

chromatograph (TCD-GC; China Chromatography CO., LTD.) and CO detector

(Gastech CO., Ltd.; GTF200).

For photoelectrochemical measurement, a water-splitting photoelectrode was used

as the working electrode with surface area of 0.5-1 cm2, a platinum plate as counter

electrode, and Ag/AgCl as reference electrode. All PEC studies were operated in a

1M Na2SO4 (pH7.0) solution as supporting electrolyte medium by using

Electrochemical Multichannel Solartron Analytical System. The water-splitting

photoelectrode was illuminated with a xenon lamp equipped with filters to simulate

the AM1.5 spectrum.

For material characterization, XRD analyses were performed on a Bruker D8

Advance diffractometer with Cu (40 kV, 40 mA) radiation. SEM measurements were

made on a JEOL 6700 filed-emission SEM. XPS spectra were obtained using a

Microlab 350 system. For obtaining TEM images, the products on the substrate were

scratched and dispersed on a carbon-coated Cu grid, and analyzed using a JEOL

JEM-2100 TEM system. Micro-Raman analyses were performed on a Jobin Yivon

Labram HR800 spectrometer. XAS analyses were performed on a beamline BL17C1

and BL20A1 at the National Synchrotron Radiation Research Center (NSRRC),

Cooling lines

Dummy Load

A-5000 Microwave power source

Mechanical Pump

Figure 3.1 Schematic diagram of microwave-plasma enhanced chemical-vapor

deposition (MPECVD) facility.

Figure 3.2 Schematic diagram of electrochemical deposition facility.

Zn2+

HMT Heating by oven

Arrayed

Arrayed ZnOZnONRsNRs

9Concentration

Arrayed ZnOZnONRsNRs

9Concentration

Cu NPs/ZnO ZnO NRs NRs

9pH Value

Cu NPs/ZnO ZnO NRs NRs

9pH Value

9Temp. Ramp Rate Figure 3.3 ZnO array preparation.

Figure 3.4 Cu NP/ZnO NR nanocomposites preparation.

Cu2+

Figure 3.5 Cu NT/ZnO NR nanocomposites preparation.

Figure 3.6 Procedure for the preparation of CuO-ZnO inverse opals using

polystyrene colloidal crystal templates.

Chapter 4

Effects of Nitrogen-Doping on the Microstructure, Bonding and Electrochemical Activity of Carbon Nanotubes

4.1 Introduction

Carbon nanoscience and nanotechnology have been developed very rapidly over

the past decade since the discovery of carbon nanotubes (CNTs). Miniaturization of

electronic and electrochemical (EC) devices using the single-walled CNTs has been

demonstrated.[49-51] Meanwhile, the remarkable structure of CNTs offers attractive

scaffolds for further anchoring of nanoparticles (NPs) and biomolecules, which is

highly desirable for energy conversion/storage and molecular sensing

applications.[6,52-62] For these applications, surface modification of the CNTs or

attaching functional groups on the sidewall become a key issue. In particular,

surface modification offers an opportunity to improve the EC reactivity of CNTs

through facilitating an efficient route for their electron-transfer (ET) kinetics with

ambient species or specific biomolecules. Therefore, understanding the ET behavior

between the CNT surface structures and the active entities are essential.

effectively functionalize CNTs surface.[63,64] Despite that many solutions to

modify CNTs are available, simple and reliable process to achieve such goal is still

lacking. In the past we have introduced heteroatom dopant such as nitrogen (N)

in-situ during the CNT growth and found it effective not only to change the atomic

structure of the CNTs into bamboo like,[7] but also to improve their electrochemical

(EC) performance down the road.[6] Although many reports on nitrogen doped

carbon nanotubes (CNx NTs) are available in the literature,[7-10] the role of

N-doping in carbon nanotube and its resultant functionality is still not clearly

understood.

In this paper, systematic studies on the effect of N incorporation in CNTs on the

morphology, microstructures, electronic states, and electrochemical properties have

been carried out. CNx NTs with different nitrogen content have been produced using

a simple in-situ nitrogen doping in a microwave plasma enhanced chemical vapor

deposition (MPECVD) reactor.[7] Comparative studies to correlate the nitrogen

content, microstructure, electronic structure, and EC performance of the CNx NTs

have been carried out. Subsequent loading of Pt NPs on the CNx NTs to study the

correlation of N dpoing with surface defect density and distribution has been carried

out.

Figure 4.1 shows the cross-sectional scanning electron microscopy (SEM) images

of the vertically aligned CNTs synthesized using different flow rate of N2 gas but

otherwise identical growth parameters. The diameters of the CNTs thus produced are

in the range of 20-50 nm with an approximate length of 2-4 μm. Bamboo-like

structure of the CNx NTs is observed, as reported in our previous paper.[7] However,

as the N2 gas flow is higher than 80 sccm the average diameter of the CNTs goes

beyond 100 nm with a reduced length of 1-2 μm. This may be attributed to the

higher growth temperature at higher N2 flow rate, which is deviated from the

nominal growth condition.

The Raman spectra of the CNx NTs prepared under different nitrogen flow rate are

presented in Figure 4.2. The 1350 cm-1 peak (D band) in Figure 4.2(a) corresponds

to the disorder-induced feature due to the finite particle size effect or lattice

distortion, while the 1580 cm-1 peak (G band) corresponds to the in-plane stretching

vibration mode E2g of single crystal graphite.[66] A D′ band around 1620 cm-1 at the

shoulder of the G band is attributed to the symmetry breaking duo to the

microscopic sp2 crystallite size.[67] The D-band position and the ratio of the D- and

G-band integrated intensities as a function of N2 flow rate are depicted in Figure

4.2(b) and 4.2(c), respectively. As the N2 flow rate increases from 0 to 40 sccm the

dependence on the N2 flow rate. Interestingly, the intensity ratio I(D)/I(G) increases

strongly as the N2 flow rate is increased to 40 sccm and also decreases above the 40

sccm optimal flow rate. In principle, the I(D) increases rapidly as a result of the

enhanced defect density. Therefore, the up-shift of the D band and the increase in the

ratio of the integrated intensities can be attributed to the increase of defect density in

the graphitic structure and/or the enhancement of the edge plane by N-induced

deformation in CNx NTs. However, further increase of N2 flow rate over 40 sccm

leads to a higher growth temperature and results in enhanced graphitization degree.

The growth temperature determined by pyroelectric thermal detector is 420, 450,

and 750 oC at the N2 flow rate of 0, 40, 160 sccm, respectively.

In order to obtain better understanding on the electronic structures of the CNx NTs,

X-ray photoemission spectroscopy (XPS) is applied to samples grown using

different N2 flow rates. Figure 4.3(a) shows the C 1s spectra of the CNx NTs after

background subtraction using Shirley’s method.[68] The C 1s spectra thus obtained

can be decomposed into three Gaussian peaks including C1 peak at 284.4±0.1 eV

representing the delocalized sp2-hybridized carbon or graphite-like C-C bonding,[69]

C2 peak at 285.1±0.1 eV reflecting defect-containing sp2-hybridized carbon

associated with the trigonal phase with a sp2 bonding,[69,70] and the C3 peak at

peak position shows a slight up-shift from 0 to 40 sccm optimal N2 flow rate

followed by a down-shift at higher flow rate. This result is in agreement with the

variation in the degree of structural disorder, which is associated with the electron

density of the defect-containing sp2-hybridized carbon and reflected in the FWHM

of the C2 peak as illustrated in the inset of Figure 4.3(a). The C2 FWHM reaches a

maximum value at 40 sccm manifesting that the most pronounced disruption in the

sp2 carbon framework due to the incorporation of N atoms into the graphene lattice

occurs at a N2 flow rate of 40 sccm.

Likewise, the N 1s XPS spectra of the CNx NTs with increasing N2 flow rate are

fitted by two Gaussian lines as shown in Figure 4.3(b). The peaks at 398.1±0.2 eV

(denoted by IP) and 400.8±0.2 eV (IG) are assigned to tetrahedral nitrogen bonded to

a sp3-hybridized carbon (so-called substitutional pyridine-like dopant structure) and

trigonal nitrogen bonded to a sp2-coordinated carbon (so-called substitutional

graphite-like dopant structure), respectively.[69] Figure 4.3(c) displays the fraction

of pyridine-like and graphite-like defects as a function of the N2 flow rate.

Increasing N2 flow rate is shown to give rise to an increase of the peak intensity

ratios of the N-sp3 C bonding (IP), which is strongly related to graphene sheets.

However, increasing N2 flow rate, keeping the rest of the process parameters the

graphene planes through the sp3-coordinated carbons. Meanwhile, the peak intensity

ratios of the N-sp2 C bonding decreases slightly over 40 sccm nominal N2 flow rate.

In general, N-doping is shown to lower the energy of pentagon defects in the

graphite-like dopant structures. Thus, through the N-sp2 C bonding, N atoms are

easily incorporated in the graphene sheets. The total N-doping levels for the peak

ratio of N/(N + C) are plotted against the respective N2 flow rate in Figure 4.3(d).

The CNx NTs with N2 flow rate of 40 sccm contain maximum N atomic ratio.

4.3 Nitrogen-Doping Effect on Electrochemical Activity

The ET behavior of CNx NTs are explored using a potassium ferrocyanide redox

probe (5 mM K4Fe(CN)6 in 1M KCl). A typical CV curve of CNx NTs

microelectrode in this redox couple system is shown in Figure 4.4(a). The

well-defined peaks obtained in the forward and reverse scans are due to the

Fe3+/Fe2+ redox couple. The reversible redox reaction of the CNx NT

microelectrodes is further evidenced by the linear Ipa and Ipc vs υ1/2 plots shown in

Figure 4.4(b), where IpaIpc and υ are the corresponding peak current densities of the

cathodic and anodic reactions and the scan rate, respectively. These data indicate

that the whole reactions are limited by semi-infinite linear diffusion of the reactants

to the electrode surface. Moreover, the effective surface area of CNx NT arrays can

CNx NTs surface to the geometrical electrode surface area, as a function of N2 flow

rate is depicted in the inset of Figure 4.4(b). It is noted that the roughness factor of

CNx NT microelectrode prepared with N2 flow rate of 40 sccm is ~33, showing

significant enhancement in the effective surface area. Clearly, the N-doping induced

electrochemically active sites on the surface of CNx NT microelectrodes prepared

with N2 flow rate of 40 sccm are optimized.

The peak-to-peak separation ( Ep△ ) of potassium ferrocyanide redox probe is

strongly dependent on the ET rate, namely, the reactivity of electrode materials to

the electrolyte. In Figure 4.4(c), Ep△ of CNx NTs with N2 flow rate of 40 sccm is

around 59 mV, reflecting excellent ET reactions. This is also related to the reduced

internal resistance of the CNx NT structures, which was determined by EC

impedance (EIS) in 1M KCl solution containing 5 mM K4Fe(CN)6 at an AC

frequency varying from 0.1 to 100 kHz as shown in Figure 4.5. From the point

intersecting with the real axis in the range of high frequency, the internal resistance

of the electrode is obtained. As shown in the inset of Figure 4.5, the arrayed CNx

NTs microelectrode prepared with N2 flow rate of 40 sccm shows the lowest

resistance of all, which is in good agreement with the above mentioned

measurements. Moreover, the Nyquist complex plane plot of the CNx NT

capability with selective N dopant.

The enhanced ET kinetics observed at CNx NTs surface may in part be attributed

to higher electronegativity of the CNx NTs surface. The attractive interaction

between the C-N dipoles present at the surface may attract the negatively charged

members of the Fe(CN)63-/4- and accelerates the redox reactions. In fact, the rate

constant for ET from the reactant to the electrode can be expressed as

kox = ∫ dє (1 - f(є,T))wox(є),

where wox is the rate of ET from an occupied level of the reactant to an empty

where wox is the rate of ET from an occupied level of the reactant to an empty