• 沒有找到結果。

Chapter 2 Literature Review

2.2 Methanol reforming

Hydrocarbon fuels have very high hydrogen contents, e.g. methane (CH4) has

hydrogen content as high as 25 wt%.[34] They may be of use for automobile and

and diesel.[35] There are three ways to transform hydrocarbons into hydrogen, i.e. (i)

direct decomposition, (ii) partial oxidation, and (iii) steam reforming.[11] The direct

decomposition and partial oxidation of hydrocarbons require an elevated

temperature and, at the same time, they produce a considerable amount of CO as a

product or a byproduct (Table 2.1).[11] In a hydrogen FC, even a trace of CO can

deteriorate the Pt electrode, e.g. on introducing 20 ppm CO into the polymer

electrolyte membrane FC (PEMFC), the current density decreases to about 18%

within 210h. In comparison with other impurities, the poisoning effect of CO is

found to exert the largest impact on FC performance.

Steam reforming of hydrocarbons produces a lower CO content relative to the

direct decomposition and partial oxidation of hydrocarbons, as listed in Table 2.1;

hence, of the above three methods, steam reforming is a promising method to

generate hydrogen for automobile and portable applications.[11] As the reforming of

methanol requires a low temperature and produces one order of magnitude less CO

than the other hydrocarbons (0.8% vs 10%-20%; see Table 2.1), it has attracted

considerable attention in the past few decades in this research field.[35] In fact, the

U.S. army is focusing on the development of the technology of reforming methanol

to generate hydrogen for portable PEMFCs, which are used as energy source for

battlefield.[36]

In the following, the reforming of methanol is specifically introduced. Methanol is

a single chemical compound (CH3OH) and commercially is primarily formed from

natural gas through a syngas route. It is a liquid under ambient conditions and has a

low boiling point of 65℃, which allows for facile vaporization in roughly the same

temperature range as that for water. For the typical range of operation in methanol

reforming (45-60 wt% methanol), the freezing point of the fuel mixture ranges from

-44 to -74℃, a distinct advantage for cold-weather development of methanol-fueled

systems. However, the high toxicity of methanol may pose a problem for large-scale

applications. Unlike gasoline or diesel, methanol does not cause vomiting when

ingested. This means that any ingestion that is not deal with quickly will result in the

formic-acid metabolism route internally.[35]

Methanol reacts with steam in the presence of Cu-based catalysts, e.g.

Cu/ZnO/Al2O3, at temperatures higher than 150℃ to form a hydrogen-rich gas.[13]

The main products are H2, CO2, and CO. The formation of methane is

thermodynamically favored, but Cu-based catalysts usually do not promote the

formation of this byproduct. The methanol steam reforming process involves the

following reactions:[11]

CO + H2O→CO2 + H2, △H0298 = -41.1 kJ/mol (3)

Equation (1) generates 12 wt% hydrogen, which is the algebraic summation of

Eqs. (2) and (3). Equation (2) represents methanol decomposition. Equation (3)

represents a water-gas shift (WGS) reaction, which is an important process to reduce

CO content in the reformate gas. The steam reforming reaction is endothermic, and

an external heat supply is required to maintain the reaction. Steam reforming of

methanol over Cu-based catalysts was originally thought to have involved the above

process, i.e. methanol decomposition, followed by WGS. However, in recent years,

this is considerable evidence to suggest another pathway including a methyl formate

intermediate.[35] At the same time, the mechanism for the formation of the CO

byproduct is a controversial topic. Even so, it is generally observed that CO can be

minimized by decreasing the contact time, increasing the steam to carbon ratio to

facilitate the WGS reaction, and decreasing the temperature, which acts to suppress

CO thermodynamically.[35] But the increase of the steam to carbon ratio requires

more heating for vaporizing the additional water in the feed.[11]

As the CO tolerance of PEMFCs is about 10-20 ppm at the normal working

temperature of 80℃ in portable applications, further processing is needed to remove

the small amount of CO (~0.8 vol%) from the reformate gas of methanol.[37]

selective CO methanation.[35] Figure 2.7 displays a schematic representation of the

integrated methanol steam reformer system, which shows the main steps and

components of steam reforming of methanol to generate hydrogen for portable

applications.[11,38]

Compared with DMFC, steam reforming of methanol does not have the methanol

crossover and the high catalyst usage problems. Current DMFCs typically use a Pt

alloy at a loading of 2-8 mg/cm2 on the anode, which is much higher than 0.2-0.3

mg/cm2 used for both the anode and cathode in PEMFCs.[35] The drawback of

steam reforming of methanol is its thermal management, because the reactor works

at temperature above 150℃ and sometimes up to 300℃. For portable FCs, the

device not only needs to be well insulated to eliminate hot surface but also the

exhaust needs to be cooled sufficiently so that it does not burn anyone.[35] As

external heat is required for steam reforming (Figure 2.7), recently an

autothermal-reforming technology was developed, which is a combination of partial

oxidation and steam reforming processes.[35] In this technology, methanol reacts

with a mixture of steam and oxygen, and the partial oxidation and steam reforming

are carried out simultaneously, where the exothermic oxidation supplies the energy

for the endothermic reforming.

energy into chemical energy that can be stored in molecules, such as hydrogen,

through electrochemical reactions, e.g. water splitting. The generation of hydrogen

from PEC water splitting was first demonstrated in 1972 by Honda and

Fujishima.[23] Besides electrolytes, the key components in a PEC cell are the

electrodes (cathode and anode) on which redox chemical reactions involving

electron transfer take place; at least one electrode should be a semiconductor (SC).

Figure 2.8 shows a simple schematic of a typical PEC device. A conventional PEC

cell is established with a SC photoanode and a Pt electrode as the cathode in the

electrolyte solution. Under irradiation with the photon energy equal to or exceeding

the band-gap energy of the SC photoanode, the electrons are excited and promoted

from the valence band to the unoccupied conduction band. n-type SCs are preferred

for the photoanode, and the depletion layer formed at the n-type SC-electrolyte

interface will lead to energy-band bending as shown in Figure 2.9, which facilitates

the separation of photogenerated electrons and holes. The electrons transport to the

cathode and react with photons to generate hydrogen, while the holes accumulate on

the surface of the photoanode and react with water molecules to produce oxygen. In

the presence of sacrificial hole scavengers such as alcohols in the electrolyte

solution, the photoexcited holes can oxidize these reducing reagents without

To achieve efficient splitting of water, the SC photoanode should meet the

following criteria: (i) photochemically stable with good corrosion resistance in

aqueous solution; (ii) with a conduction band edge more negative than hydrogen

evolution potential and a valence band edge more positive than the oxygen evolution

potential; (iii) strong absorption in the solar spectrum region; (iv) high-quality

material with low density of defects for efficient charge transfer and reduction of the

electron-hole recombination; and (v) low cost.[41-43] Unfortunately, to date, there is

no such material that can meet all the requirements simultaneously. Among the

various candidates for the photoelectrode, SC metal oxides are relatively

inexpensive and have a better photochemical stability. Many metal oxides have been

extensively studied and considerable progress has been made in recent years.[26]

For a PEC cell, the conduction band of most metal oxide material is less negative

than the hydrogen evolution potential; thus, a small external potential needs to be

applied to facilitate the PEC reactions.

The SCs for PEC water splitting can be generally classified as metal oxide and

conventional photovoltaic (PV) material. The SC photoelectrode can be n-type

(Figure 2.10a), p-type (Figure 2.10b) or coupling of n-type and p-type (Figure 2.10c).

This can be a single photosystem as in the n-type (TiO2) or p-type (InP), but for the

solar spectrum or several p-types can also be done the same way (Figure 2.10d).

When involving more than one photosystem, it is important to match the currents

generated by the different layers to obtain better efficiency and this is achieved by

aligning complimentary band gaps and controlling the thickness or active area.[21]

In PEC water splitting, metal oxide and conventional PV material or their

combination are used. The anode and cathode are usually physically separated, but

can be combined into a monolithic structure either using a metal structure by

depositing the anode on one side and cathode on the other and sealing the edges

(Figure 2.10e) or stacking the anode on its own substrate with the cathode on its own

substrate and providing an electrical connection between the two (Figure 2.10f).[21]

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.