• 沒有找到結果。

Chapter 4 Effects of Nitrogen-Doping on the Microstructure, Bonding and

4.4 Summary

Vertically aligned CNx NTs have been synthesized using microwave plasma

enhanced chemical vapor deposition with different nitrogen flow rate. The

microstructure, bonding and electrochemistry properties have been investigated

utilizing Raman, XPS, CV and AC impedance measurements to establish

correlations among the N content, electronic structure, microstructure and EC

activities. The results show that nitrogen incorporation provides a simple pathway to

engineering/modifying the microstructure and electronic bonding structure, which is

crucial to the ET kinetics of the CNx NT electrode. In-situ N doping of the CNTs at

an optimal 3.5 at. % N is effective in converting the nanotubes into bamboo-type

morphology and promoting substitutional graphite-like defect structure, which is

favorable for fast ET in electrochemistry. The surface defects thus created further

enhance anchoring of Pt atom in the subsequent selective EC deposition to produce

ultrahigh density Pt NPs to form nanocomposite, which is ideal for many

highly desirable in many EC based applications.

Figure 4.1 Cross-sectional SEM images of the vertically aligned CNTs synthesized

at different flow rate of N2: (a) 0, (b) 80, and (c) 120 sccm.

( ( a a ) )

(c ( c) )

(b ( b) )

Figure 4.2 (a) Comparison of the peak intensities and the full width at half

maximum (FWHM) of the first-order Raman spectra for the vertically aligned CNx

NTs prepared with different N2 flow rate during growth. (b) D-band position as a

function of N2 flow rate. (c) ID/IG as a function of N2 flow rate.

Figure 4.3 (a) The C1s XPS spectra of the vertically aligned CNTs prepared with

various N2 flow rate during growth. The inset is the FWHM of the C-N bonding

component as a function of N2 flow rate. (b) The N1s XPS spectra of the vertically

aligned CNTs prepared with different N2 flow rate. (c) The IP and IG as a function of

the N2 flow rate for the N1s peak. (d) The N-doping concentration as a function of

Figure 4.4 (a) Cyclic voltammetry of the vertically well-aligned CNTs modified with

different N-doping level in 1 M KCl and 5 mM K4Fe(CN)6. (b) The ferricyanide

peak current versus the scan rate (v)1/2 plot for CNTs using various flow rate of N2

for both anode and cathode. (★: 0 sccm, ■: 40 sccm, ●: 80 sccm, ◆: 120 sccm, ▲:

Figure 4.5 AC impedance analysis of the vertically aligned CNTs with different

N-doping level in 1 M KCl and 5 mM K4Fe(CN)6. The inset shows the internal

resistance as a function of the N2 flow rate.

Figure 4.6 TEM images of the Pt NP-CNx NT hybrid nanostructures synthesized at

different flow rate of N2: a) 0, b) 40, and c) 120 sccm.

( ( a a ) ) ( ( b b ) )

( ( c c ) )

10 nm 5 nm

5 nm

0.0 0.2 0.4 0.6 0.8 1.0 0

10 20 30

Cur ren t D e ns ity ( mA /c m

2

)

Potential(V vs. Ag/AgCl)

0 sccm 40 sccm 120 sccm

Figure 4.7 Typical CV curve of the arryed Pt NP-CNx NT nanocomposites with

different N2 flow rate at a scan rate of 50 mV/s in 1 M CH3OH + 1 M H2SO4

solution.

Chapter 5

Nanostructured ZnO Nanorod@Cu Nanoparticle as Catalyst for Microreformers

5.1 Introduction

The use of hydrogen for energy generation has attracted significant attention in the

recent years as a clean, sustainable and transportable fuel alternative, and has

consequently sparked a rapid global development of hydrogen fuel cells for electric

power generation.[1] Catalytic reformation of hydrocarbons, with careful attention

to avoid storage and safety issues,[75,76] is currently the predominant process for

hydrogen generation. One of the leading and most promising techniques for

hydrogen generation is catalytic reformation of methanol.[77,78] Cu/ZnO-based

catalysts are, therefore, of great importance for industrial scale catalytic production

of reformate hydrogen.[13] Owing to their wide commercial relevance,

Cu/ZnO-based catalysts, prepared via several preparation routes, are being

extensively investigated; and substantial improvements in their efficiency of

catalytic activity brought about by addition of suitable promoter/support,

hetero-nanostructures as reforming catalysts is still lacking to date. This inspired us

to design the core-shell nanostructured catalyst, consisting of a ZnO nanorod (NR)

core and an outer shell of Cu nanoparticles (NPs), i.e. NR@NPs, for achieving high

efficiency of catalytic conversion.

In addition, the idea of using microreformers is highly attractive for several

applications such as on-board hydrogen sources for small vehicles and portable fuel

cells (FCs).[79,80] However, two key issues have hindered the realization of

microreformers for catalysis, viz. poor adhesion between the catalyst layer and the

microchannels; and poor utilization of catalyst layer deposited in the form of thick

film.[81] Notwithstanding the several approaches investigated to overcome these

issues, catalyst immobilization and its efficient utilization inside the microchannel

remains a challenge.[82,83] Most of these approaches involve a two-step process,

wherein active catalysts are prepared in the first step, followed by its immobilization

on the surface of the microchannels in the second step. In this communication, we

report a simple and reliable method for integrating in-situ synthesis of catalyst and

its immobilization for microreformer applications.

5.2 Structural Characterization of Cu Nanoparticle/ZnO Nanorod

Nanocomposites

ZnO NR arrays at low temperature subsequently resulted in spontaneous formation

of the cable-like nanostructures. Since the ZnO NR@Cu NP nanocomposites were

in-situ synthesized directly on microreactor, the arrayed ZnO@Cu nanocomposites

were strongly anchored onto the microchannel. This was evidenced by observation

of no material loss after sonication in the water for several hours, which highlighted

the strong mechanical anchorage of nanostructured catalysts on the surface of

microchannel. The interaction between Cu NP and ZnO NR was studied by several

analytical techniques, including electron microscopy (EM), X-ray diffraction (XRD),

X-ray photoelectron spectroscopy (XPS), X-ray absorption spectroscopy (XAS), and

temperature-programmed reduction (TPR). The structure of the microreformer

design based on the ZnO NR@Cu NP nanocomposite is illustrated in Figure 5.1,

which is evidenced by the photographs comparing the microchannels before and

after the deposition of the ZnO NR@Cu NP nanocomposite (see Figure 5.2).

One of the most significant advantages of the core-shell nanocomposites, which

are clearly distinct from the traditional catalysts, is the large surface area they offer

for effective surface contact between the reactants and catalysts. Figure 5.3a

illustrates the cross-sectional SEM image of vertically aligned ZnO NRs grown on

the inner surface of the microchannel. The size of the NRs range from 35 to 50 nm

uneven surface with arrow-marked stacking faults, which is shown in greater detail

in Figure 5.3c. Figure 5.3d displays the typical TEM image of ZnO@Cu hybrid

nanocomposites at Cu decoration concentration of 2 mM. Close attachment of Cu

NPs on the ZnO NR cores can clearly be observed. More detailed TEM images with

EDX elemental mapping of Cu and Zn are shown in Figure 5.4, which further

confirms the attachment of Cu NPs on the ZnO NR cores. Further, a high-resolution

TEM image, shown in Figure 5.3e, yielded the spacing of the{111}lattice planes of

the fcc copper crystal to be 0.21 nm. A histogram of the diameter of Cu NPs

determined from the TEM measurement is shown in Figure 5.3f, indicating a

diameter range of 3 to 8 nm with an average of 5 nm. Moreover, a high-resolution

TEM image of the Cu NP/ZnO NR heterostructures reveals that the (111) plane of

the Cu NP is immobilized on the (002) plane of ZnO NR with clear evidence of the

lattice-mismatched region (see Figure 5.5).

Figure 5.6a shows XRD patterns of Cu deposited on ZnO NRs prepared with

different Cu decoration concentrations from 1 to 3 mM. A large range XRD patterns

can be as reference in the Figure 5.7. No obvious Cu peak could be detected below 1

mM decoration concentration. Figure 5.6a also shows that as the decoration

concentration gradually increases, Cu(111) peak at 43.5 degree becomes more

size of Cu particles in the nanocomposite samples. Furthermore, the FWHM

increases as the decoration concentration decreases, indicating formation of finer

particles at lower decoration concentration, which is also confirmed in the results of

larger Cu surface area and higher dispersion determined via N2O decomposition

method (see Table 5.1). In addition, in Figure 5.6a, a shift in the position of this peak

to higher 2θ values as compared to the metallic Cu peak is also observed with

decreasing decoration concentration. This can be attributed to the existence of

defects at the interface between Cu NP and ZnO NR or partial dissolution of Cu into

the ZnO lattice. The defects, either microstrain or structural disorder, can originate at

the interface due to the lattice mismatch between Cu and ZnO, but can be quickly

overwhelmed by the strong metallic Cu(111) peak from larger sized particles

deposited at higher decoration concentrations, as can be observed from Figure 5.6a.

A detailed inspection of the electronic states of the surface metal species was

carried out through XPS analysis. Figure 5.6b depicts the Cu 2p core level XPS of

NR-Cu composite nanostructures prepared at different decoration concentrations. In

order to differentiate between the oxidation states of Cu, the main peak of the Cu

2p3/2 core level spectra was fitted with two components at 932.8 and 933.7 eV

corresponding to Cu0/Cu+ and Cu2+ species, respectively.[86] It can be seen that a

binding energies above 940eV further indicates a lower Cu oxidation state.[87]

When the decoration concentration was decreased, the position of the Cu main peak

shifted slightly toward higher binding energy. This is strongly related to the

modification of the electron density on smaller Cu species at lower decoration

concentrations.[86] Moreover, it appears that the fraction of atomic Cu on the

surface, as estimated from the ratio of Cu/(Cu + Zn + O) peaks tends to increase

with increasing decoration concentration. In contrast, the fraction of atomic Cu on

the surface is lower than 8 at. % while the decoration concentration is decreased to 1

mM, which is consistent with the XRD results in Figure 5.6a. Meanwhile, similar

trend can be obtained for the surface ratios of Cu/Zn and ZnO/Cu through XPS

measurements (see Table 5.1). However, it is rather unfortunate that Cu0 state cannot

be distinguished from Cu+ state by XPS analysis due to their spectral overlap. X-ray

absorption spectroscopy (XAS) measurements were hence used to resolve this issue.

The X-ray absorption near edge structure (XANES) spectra are associated with the

excitation of a core electron to bound and quasi-bound states. To determine the

valence of copper in the arrayed ZnO NR@Cu NP nanocomposites, the Cu K-edge

XANES spectra are compared with those of Cu foil in Figure 5.6c. The XANES

spectra of arrayed nanocomposites exhibit the edge absorption at 8979 eV together

nanocomposites. A more detailed understanding may be gained from the derivative

of the XANES spectra (see Figure 5.8). The intensity of the edge absorption for

arrayed nanocomposites is relatively low compared to that of Cu foil. This could be

related to the smaller (nano-sized) dimensions of the Cu particles.[88] Additionally,

it may be noted that the nanocomposite samples display lower intensity as well as a

shift towards higher energy of the peak at ca. 8990 eV, together with the shoulder

around 8984 eV, as compared to the derivative of Cu foil. These features suggest

some changes in the chemical environment around the copper species in these

nanocomposites. The reason may be attributed to a further distortion of the copper

lattice due to the presence of microstrain at the interface between Cu NPs and ZnO

NRs. Thus, the XAS results seem to agree well with the XRD results.

To investigate the reducibility of NR@Cu NP nanocomposites prepared at

different decoration concentrations, TPR measurements were carried out in the

temperature range of 140 to 300 oC, as shown in Figure 5.6d. It is known that the

reduction of bulk CuO is indicated by a single reduction peak at a considerably

higher temperature of 320 oC.[89] Clearly, the Cu NP-immobilized ZnO NR

nanoarchitectures with 2 mM decoration concentration exhibits a significant shift in

the position of the reduction peak to lower temperature in addition to the reduction

NR@Cu NP catalysts can be emphasized by the lower reduction temperature than

the commercial catalysts (Cu-ZnO-Al2O3; MDC-3: Sud-Chemie) as shown in the

same figure. The reason can be attributed to the enhanced dispersion of fine Cu NPs

and strong metal-support interaction (SMSI) effect under NR@NP nanosystems,

which facilitates the reduction of the supported Cu species. However, at decoration

concentrations of higher than 2 mM, the increase in the size of Cu particles plays a

predominant role in the structural change of ZnO@Cu nanocomposites, which also

weakens the interaction between Cu and ZnO and consequently undermines the

catalytic activity.

5.3 Test of Methanol Reforming Reaction

Methanol conversion and hydrogen production rate for the arrayed ZnO@Cu

nanocomposites and the commercial catalysts are shown in Figure 5.9a and 5.9b,

respectively. Each data point represents a 12-hour experimental run with very small

deviation observed with repeated runs, which is indicated by the small error bars in

the figures. The methanol conversion rate over the arrayed ZnO@Cu

nanocomposites is as high as 93 % with a hydrogen production rate of 183 mmol

gcat-1 h-1 at 250 oC. Both these rates are significantly higher than those obtained with

the commercial catalysts. Small amounts of CO were detected by CO detector in the

than ca. 1000 ppm formed by the commercial catalysts. The present results thus

provide a promising alternative to the commercial catalyst for use in catalytic

generation of high purity hydrogen without necessitating additional processes to

remove CO down stream. For a better comparison of the catalytic activities of

NR@NP nanoarchitectures and commercial catalysts, kinetic parameters were

evaluated from the conversion data.[90] The kinetic constant at a given temperature

can be calculated from the equation:

κ = 1/τ [εx + (1+ε)ln (1-x)],

where κ, τ, x, and ε are the kinetic constant, contact time, fractional conversion,

and fractional expansion of the system, respectively. As shown in Figure 5.9c, the

kinetic constant data manifest higher activity of the nanostructured NR@NP arrays

than the commercial catalysts. This can be attributed to the greater external surface

area, the existence of SMSI effect, the presence of microstrain at the interface, and

the modification of electronic structures in the nanocomposites. These effects are

consistent with the TEM, XRD, XPS, XANES, and TPR data. Based on the reaction

temperature dependence of the kinetic constant, the apparent activation energy was

calculated from the slope of the Arrhenius plot (Figure 5.9d). It is evident that the

reaction activation energy for the nanostructured NR@NP arrays is noticeably lower

nanostructured catalysts than on the commercial catalysts. Last but not least, the

advantage of the nanostructured NR@NP arrays catalyst is further exemplified by

the higher stability (11.5 % reduction after 36 hours operation) in methanol

reforming reaction, which is significantly superior to the 33.8 % reduction for the

commercial catalysts (see Figure 5.10).

5.4 Summary

In summary, we have successfully demonstrated an easy-to-fabricate route to

prepare arrayed ZnO NR@Cu NP heterostructures on the inner surface of

microchannels via direct in-situ synthesis for microreformer applications. The

superb catalytic performance and stability of the Cu NP-decorated ZnO NR

nanostructures can be attributed to the larger surface area and enhanced dispersion

of fine Cu NPs, formation of microstrain, the modification of electronic structure of

Cu species, and the existence of SMSI effect. These results present new

opportunities in the development of highly active and selective NR@NP

nanoarchitectures for a wide range of different catalytic reaction systems.

Microchannel Reactor

Gas In

Gas

Out Arrayed Arrayed ZnOZnONRsNRs Cu NPs

Out Arrayed Arrayed ZnOZnONRsNRs Cu NPs

Figure 5.1 Schematic diagram of the novel catalyst ZnO NR@Cu NP arrays grown

on the inner surface of microchannel reactor.

Figure 5.2 Photographs of the microchannels a) before and b) after homogeneously

depositing the ZnO NR@Cu NP nanocomposites.

20 nm

Figure 5.3 a) Cross-sectional SEM image of vertically well-aligned ZnO NRs. b)

Typical TEM image of a single ZnO NR showing the presence of stacking faults as

marked with arrows. c) HRTEM image of the ZnO NR. d) Typical TEM image of

the ZnO NR@Cu NP nanocomposites. e) HRTEM image of Cu NPs on the surface

of one single ZnO NR. f) The size histogram of Cu NPs analyzed from the HRTEM

20 nm

a)

20 nm

b)

20 nm

c)

20 nm

a)

20 nm 20 nm

a)

20 nm

b)

20 nm 20 nm

b)

20 nm

c)

20 nm 20 nm

c)

Figure 5.4 a) TEM image and corresponding EDX elemental mapping of b) Cu and c)

Zn.

ZnO NR 2 nm

Cu NP

ZnO(002) Cu(111)

mismatch

ZnO NR 2 nm

Cu NP

ZnO(002) Cu(111)

mismatch

Figure 5.5 HRTEM image of ZnO NR@Cu NP heterostructures.

41 42 43 44 45 46

8960 8980 9000 9020 9040

2 mM

925 930 935 940 945 950

Cu 2p3/2

8960 8980 9000 9020 9040

2 mM

925 930 935 940 945 950

Cu 2p3/2

Figure 5.6 a) XRD patterns of Cu NPs on the surface of ZnO NRs prepared with

different decoration concentrations from 1 to 3 mM. b) Cu 2p XPS core level spectra

of arrayed ZnO NR@Cu NP nanocomposites prepared with different Cu decoration

concentrations from 0.5 to 3 mM. c) Cu K-edge XANES spectra of arrayed ZnO

NR@Cu NP nanocomposites and bulk reference sample Cu foil. d) Comparative

TPR profiles of ZnO NR@Cu NP nanocomposites and commercial catalysts

(MDC-3: Sud-Chemie).

30 35 40 45 50 55 60 65

In te n s it y ( a .u .)

2 θ (degree)

3 mM

2 mM

1 mM

▲ ▲

▲▲

30 35 40 45 50 55 60 65

In te n s it y ( a .u .)

2 θ (degree)

3 mM

2 mM

1 mM

▲ ▲

▲▲

Figure 5.7 Large range of XRD patterns of Cu NPs on the surface of ZnO NRs

prepared with different decoration concentrations from 1 to 3 mM. Peaks marked

with ▲ are due to ZnO, the other marked with ■ is due to Cu.

4.4

8976 8980 8984 8988 8992

Cu K-edge

Table 5.1 Microstructure properties of Cu NPs on the surface of ZnO NRs prepared

with different decoration concentrations.

Figure 5.8 First derivative of the Cu K-edge XANES spectra of arrayed ZnO

NR@Cu NP nanocomposites and bulk reference sample Cu foil.

140 160 180 200 220 240 260 280 300

140 160 180 200 220 240 260 280 300 0

H2 production rate (mmol gcat -1 h-1 )

Temperature (oC)

140 160 180 200 220 240 260 280 0

140 160 180 200 220 240 260 280 300 0

140 160 180 200 220 240 260 280 300 0

H2 production rate (mmol gcat -1 h-1 )

Temperature (oC)

140 160 180 200 220 240 260 280 0

Figure 5.9 a) Methanol reforming reaction profiles for 2 mM arrayed Cu NPs/ZnO

NRs nanocomposites (□) and commercial catalysts (●). b) Hydrogen production rate

as a function of reaction temperature for 2 mM arrayed Cu NPs/ZnO NRs

nanocomposites (□) and commercial catalysts (●). c) Kinetic constants as a function

of reaction temperature for 2 mM arrayed Cu NPs/ZnO NRs nanocomposites (□)

and commercial catalysts (●). d) Arrhenius plots for methanol reforming reaction for

2 mM arrayed Cu NPs/ZnO NRs nanocomposites (□) and commercial catalysts (●).

0 4 8 12 16 20 24 28 32 36 0

20 40 60 80 100

M e tha nol c o nv e rs io n ( % )

Time on Stream (hr)

11.5 % loss 33.8 % loss

0 4 8 12 16 20 24 28 32 36 0

20 40 60 80 100

M e tha nol c o nv e rs io n ( % )

Time on Stream (hr)

11.5 % loss 33.8 % loss

Figure 5.10 Stability tests in methanol reforming reaction over 2 mM arrayed Cu

NPs/ZnO NRs nanocomposites (□) and commercial catalysts (○). Reaction

conditions: H2O/O2/MeOH = 1/0.125/1, Temperature = 250 oC, W/F = 21 kgcat s

mol-1methanol.

Chapter 6

Microwave-Activated CuO Nanotip/ZnO Nanorod Nanoarchitectures for Efficient Hydrogen Production

6.1 Introduction

In the field of development of advanced energy-related devices, nanostructured

materials have attracted great interest in recent years due to their intriguing

physical and chemical properties that are significantly different from their bulk

counterparts.[91-93] Nanostructured catalysts are expected to be the key

components in the advancement of future energy technologies. Hence, new

strategies for the synthesis of high-performance nanomaterials are widely

pursued.[94] In particular, Cu/ZnO-based catalysts have been extensively

investigated because of their importance in several industrial applications, such

as methanol synthesis, CO removal, and hydrogen generation.[17,35,78,95-99]

Nanostructured catalysts consisting of microporous structures are attractive for

flow-type gas-solid reactions due to their effective diffusion of reactants and heat.

In our previous work, strong metal-support interaction (SMSI) effect in arrayed

These findings led us to design and develop a much effective technique to tailor

the Cu/ZnO interface.

In recent years, microwave (MW) irradiation has become an effective tool in

synthetic organic chemistry, yielding dramatic increases in reaction rates and

yields.[101-104] Moreover, since the magnitude of MW absorption is related to

dipole oscillation in a material, MW annealing can provide selective heating and

lead to a new route for material processing. Recently, application of MW as selective

annealing technique has attracted tremendous attention for improving the efficiency

of polymer organic photovoltaic devices.[105] In this communication, we report

MW treatment of nanoarchetectures of CuO nanotip (NT) on ZnO nanorod (NR)

framework as catalyst precursors for methanol reforming reaction (MRR).

6.2 Structural Characterization of CuO Nanotip/ZnO nanorod Nanocomposites

Our initial attempts employing a surfactant-free approach

(ammonia-evaporation-induced synthetic method) for synthesis of CuO

nanostructures[106] resulted in formation of radiating nanosheets as shown in

Figure 6.1a. The high-resolution transmission electron microscopy (HRTEM)

image in Figure 6.2 provides specific structural information about an individual

different morphologies of CuO were observed, ranging from nanosheet to NT

(Figure 6.1b). Figure 6.3 presents the TEM and HRTEM images of CuO NT/ZnO

NR catalyst precusors synthesized in this work, showing that NTs not only

adhere to but also directly conjoin with the NR. While detailed understanding of

adhere to but also directly conjoin with the NR. While detailed understanding of