Sunteți pe pagina 1din 17

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/226334416

Copper and Copper Oxide Nanoparticle


Formation by Chemical Vapor Nucleation From
Copper (II) Acetylacetonate
ARTICLE in JOURNAL OF NANOPARTICLE RESEARCH NOVEMBER 2001
Impact Factor: 2.18 DOI: 10.1023/A:1012508407420

CITATIONS

READS

62

213

5 AUTHORS, INCLUDING:
Albert G Nasibulin

Igor Altman

Skolkovo Institute of Science and Technology

49 PUBLICATIONS 526 CITATIONS

241 PUBLICATIONS 3,991 CITATIONS

SEE PROFILE

SEE PROFILE

All in-text references underlined in blue are linked to publications on ResearchGate,


letting you access and read them immediately.

Available from: Albert G Nasibulin


Retrieved on: 06 March 2016

Journal of Nanoparticle Research 3: 385400, 2001.


2001 Kluwer Academic Publishers. Printed in the Netherlands.

Copper and copper oxide nanoparticle formation by chemical vapor


nucleation from copper (II) acetylacetonate
Albert G. Nasibulin1 , P. Petri Ahonen1 , Olivier Richard1 , Esko I. Kauppinen1, and Igor S. Altman2
VTT Chemical Technology, Aerosol Technology Group, P.O. Box 1401, FIN-02044 VTT, Finland;
2
Institute of Combustion & Advanced Technologies, Odessa National University, Dvoryanskaya 2,
Odessa, 65026, Ukraine; Author for correspondence (Tel.: +358 9 456 6165; Fax: +358 9 456 7021;
E-mail: Esko.Kauppinen@vtt.fi)

Received 22 January 2001; accepted in revised form 20 June 2001

Key words: Cu, Cu2 O, nanoparticle, copper acetylacetonate, thermal vapor decomposition, chemical nucleation

Abstract
Crystalline nanometer-size copper and copper (I) oxide particle formation was studied by thermal decomposition of
copper acetylacetonate Cu(acac)2 vapor using a vertical flow reactor at ambient nitrogen pressure. The experiments
were performed in the precursor vapor pressure range of Pprec = 0.06 to 44 Pa at furnace temperatures of 431.5 C,
596.0 C, and 705.0 C. Agglomerates of primary particles were formed at Pprec > 0.1 Pa at all temperatures. At
431.5 C the number mean size of the primary particles increased from Dp = 3.7 nm (with geometric standard
deviation g = 1.42) to Dp = 7.2 nm (g = 1.33) with the increasing precursor vapor particle pressure from 1.8 to
16 Pa. At 705.0 C the primary particle size decreased from Dp = 24.0 nm (g = 1.57) to Dp = 7.6 nm (g = 1.54),
respectively.
At furnace temperatures of 431.5 C and 596.0 C only crystalline copper particles were produced. At 705.0 C the
crystalline product of the decomposition depended on the precursor vapor pressure: copper particles were formed
at Pprec > 10 Pa, copper (I) oxide at Pprec 1 Pa, and a mixture of the metal and its oxide at intermediate vapor
pressures. A kinetic restriction on copper particle growth was shown, which leads to the main role of Cu2 molecule
participation in the particle formation. The formation of copper (I) oxide particles occurs due to the surface reaction
of the decomposition products (mainly carbon dioxide). For the explanation of the experimental results, a model is
proposed to build a semiempirical phase diagram of the precursor decomposition products.

Introduction
Copper and copper oxide particles are of significant
technological interest. Applications for copper powder include bronze self-lubricating bearings, conductive epoxys, metal-bonded abrasive wheels and cutting
tools, and braking systems. Ultra-fine copper particles
are a base for developing technologies such as metal
injection molding as well as for electronics, ceramics and for thick/thin-film applications. Copper oxides
have applications in thin-film oxygen pressure sensors,
as a binder in pastes for thick-film microelectronic

circuits, as a p-type semiconductor and they exhibit


luminescence (Majumdar et al., 1996; Holzschuh &
Suhr, 1990).
The importance of producing copper and copper
oxide particles is exemplified by applications such as
high surface area catalysts that are used in diverse
chemical processes, for example, the water-gas shift
reaction (Campbell et al., 1987), the butanol dehydrogenation reaction (Shiau & Tsai, 1997) and the carbon monoxide oxidation (Van der Meijden, 1981; Du
et al., 1997). Copper-based catalysts are used as a key
intermediate in the industrial synthesis of methanol

386
(Klier, 1982; Campbell et al., 1987; Yurieva et al.,
1993; Klenov et al., 1998), which is promising as
an environmentally friendly fuel for the power industry. In addition, copper and copper-based materials
have applications as catalysts in traditional and new
organic syntheses, for example, the condensation of
aromatic halides, known as Ullmann reaction (Dhas
et al., 1998), synthesis of cyclic amines from aminoalcohols and their alkyl derivatives (Hammerschmidt
et al., 1986; Kijenski et al., 1989), synthesis of methylamines (Gredig et al., 1997), thermal cracking of plastics and many others (Kijenski et al., 1984; Runeberg
et al., 1985; Vultier et al., 1987; Pereia et al., 1994;
Shannon et al., 1996). Copper has been also identified as a good catalyst for the combustion of methane
(Tijburg, 1989) and selective oxidation of hydrocarbons (Adams & Jennings, 1964; Voge & Adams,
1967).
Thus, studies of the formation of copper and copper oxide particles are important. Accordingly, it has
been a subject of much research during the last decades.
Copper and copper oxide particle formation have
been studied by solution reaction: reduction of copper (II) acetate in ethanol (Ayaappan et al., 1997)
and in water and 2-ethoxyethanol using hydrazine
(Huang et al., 1997), the reaction of copper (II)
chloride with organolithium compounds (Takahashi
et al., 1988), reduction of copper dodecylsulfate by
sodium borohydrate (Lisieski et al., 1996), by using
reverse micelles (Lisieski & Pileni, 1993; 1995),
thermal and sonochemical reductions of copper (II)
hydrazine carboxylate (Dhas et al., 1998), chemical deposition in two-phase system octanewater
(Vorobyova et al., 1997), by means of electrolysis
(Folmanis & Uglov, 1991; Kirchheim et al., 1991;
Pietrikova & Kapusanska, 1991) and others (Herley
et al., 1989). Ding et al. (1996) prepared copper
nanoparticles by a mechanochemical process and
studied the influence of milling conditions on particle structure and size. Nanocrystalline copper was
prepared by consolidation of mechanically milled
powder by Weins et al. (1997). Copper oxide formation was studied via oxidation of copper particles (Kaito et al., 1973; 1993; Kellerson et al.,
1995).
Much work has been devoted to the investigation of
aerosol formation of copper or its oxide particles by
physical methods, including molecular beams (Bowles
et al., 1981), direct laser vaporization (Moini and Eyler,
1988), by using gas evaporation (Kashu et al., 1974;

Peoples et al., 1988; Xu et al., 1992), sputtering


(Haas & Birringer, 1992), melting in a cryogenic
liquid (Champion & Bigot, 1996) and others (Long
& Petford-Long, 1986; Girardin & Maurer, 1990;
Bouland et al., 1992). Another and a very attractive way
to obtain aerosol particles is via the chemical route.
This method might be the least expensive for aerosol
particle formation under controlled conditions. Little
work has been devoted to copper and its oxide formation by chemical vapor nucleation. Majumdar et al.
(1996) generated CuO powder by spray-pyrolysis from
Cu(NO3 )2 solution. Daroczi et al. (1998) studied the
production of copper and iron nanocomposites by thermal decomposition of copperferrocyanide in an open
vertical tube. The obvious advantage of the chemical
vapor nucleation method is the possibility to produce
nanosized particles at relatively low temperatures and
ambient pressure.
The current work is devoted to the investigation
of copper and copper oxide particle formation from
metalorganic compound, copper (II) acetylacetonate
(Cu(acac)2 ). A suitable equilibrium vapor pressure
(P = 13.1 Pa at t = 150.0 C) and a convenient
decomposition temperature (tdec = 286 C) were the
reasons for the choice of this precursor. The selection
of this metalorganic substance was also based on its
popularity as a precursor for chemical vapor deposition processes (e.g., Pelletier et al., 1991; Pauleau &
Fasasi, 1991; Gerfin et al., 1993; Marzouk et al., 1994;
Hammadi et al., 1995; Maruyama & Shirai, 1995) and
on the knowledge of decomposition reaction kinetics
(Tsyganova et al., 1992). The goals of the investigation
are to produce copper and/or copper oxide nanoparticles at ambient pressure and at a temperature as low
as possible, to characterize the obtained nanoparticles
synthesized with various reactor conditions, and to
discuss nanoparticle formation based on experimental
results.

Experimental
Materials
For this study copper (II) acetylacetonate, Cu(acac)2 ,
(Aldrich Chemical Company, 97%) has been used as
a precursor. The decomposition of Cu(acac)2 vapor
leads to the formation of copper vapor that is supersaturated at the experimental conditions. The Cu(acac)2

387
decomposition reaction can be presented as
CH3

CH3
O

HC

O
Cu

O
CH3

CH
O
CH3

O
Cu + CuO + H2C CH2(CO) + CO2 + H3C C CH3
36.0
22.3
38.0
O
+ H2O + H3C C C2H5 ,
3.2
0.3

(1)

where numbers below the decomposition reaction


products indicate molar percentage fraction of the
gaseous reaction products which were measured using
mass-spectrometry analysis by Tsyganova et al. (1992).
Such reaction products indicate significant destruction of the ligand, acetylacetone, which is possibly
formed on the initial stage of the thermal decomposition. The Cu(acac)2 vapor decomposition was studied
using the manometric method in static conditions in
clear ampoules at 290335 C and at initial pressures of
98173 hPa (Tsyganova et al., 1992).
In our study two sources of nitrogen (AGA,
99.9 vol.% and 99.999 vol.%) have been used as a
carrier gases. Inert nitrogen was used in order to prevent the additional reoxidation of the formed copper vapor by atmospheric oxygen. The compaction of
Cu(acac)2 powder in a saturator and hence, blocking the
flow were prevented by mixing the powder with inert
chromatographic carrier, silicon dioxide, SiO2 (Balzers
Materials, 99.9%) with a grain size of 0.20.7 mm.
Experimental methods
A vertical laminar flow reactor for experimental investigation of Cu(acac)2 decomposition under controlled
conditions was designed and constructed (Figure 1).
The experimental device consisted of a saturator, a laminator, and a furnace. The saturator and the laminator consisted of a stainless steel tube with an internal
diameter of 22 mm. A removable cartridge, to hold the
mixture of Cu(acac)2 powder and the chromatographic
carrier, was inserted inside the tube. An absolute filter
(Munktell, MK 360) was used to clean the vaporgas
flow downstream of the saturator. The saturation mixture and the filter were retained on a stainless steel net.
In order to laminarize the flow and to avoid turbulence

of the flow proceeding from small diameter to large


one after the saturator a laminator has been used. It was
constructed as a cylindrical cone in the junction part.
Between the laminator and the furnace, a Teflon thermoinsulator was used. A ceramic tube, with external
and internal diameters of 28 and 22 mm, respectively,
inserted inside the furnace (Entech, Sweden) has been
used as a reactor.
The flow of pure filtered nitrogen carrier gas was
supplied from a high-pressure cylinder to the saturator. Then the gas passed through the heated Cu(acac)2
powder and the vapor saturation by the precursor was
reached. Inside the laminator, a steady state laminar flow was established. Then the vaporgas mixture entered to the furnace where the temperature
is maintained higher than the Cu(acac)2 decomposition temperature. The formation of supersaturated
copper vapor, as a result of the precursor decomposition reaction, led to the nucleation process and further
growth of particles via condensation, coalescence, and
agglomeration processes.
The flow rate of the gas-carrier was measured by
a flow meter (DC-2, BIOS) and was referred to the
standard condition (t = 25 C, P = 101325 Pa).
Temperatures were measured by nichromenickel thermocouples (K-type) which had been calibrated with an
accuracy of 0.1 C by using an oil bath and thermoresistors calibrated against the Finnish National Standard.
The aerosol number size distributions in the range of
3200 nm were measured by a differential mobility
analyzer (DMA) system consisting of a charger, a classifier (Winklmayr et al., 1991, modified Hauke, length
of 11 cm), a condensation particle counter (CPC, TSI
3027), and a supporting software. A sheath flow rate
for DMA system was maintained at Qsh = 14.5 lpm.
The morphology, the primary particle size, and the
crystallinity of the particles were investigated with
a field emission scanning electron microscope (Leo
Gemini DSM982) and a field emission transmission
electron microscope (Philips CM200 FEG), respectively. An electrostatic precipitator (Combination electrostatic precipitator, InTox Products, Albuquerque,
NM, USA) was used to collect the aerosol particles
on a carbon-coated copper grid (SPI Holey Carbon
Grid). Electron diffraction patterns of the particles
were used for determination of the crystalline phase.
The samples for X-ray diffraction (XRD, Philips MPD
1880 powder X-ray diffractometer) spectrometry were
collected on silver filter disks (Millpore AG4502500,
45 m pore) and studied with Cu K ( = 0.154 nm)

388

Figure 1. Schematic presentation of the experimental setup.

radiation. A Mettler Toledo TA8000 system equipped


with TGA850 thermobalance was used for thermogravimetric analysis (TGA) of the samples under
flowing nitrogen atmosphere with the heating rate of
10 C/min for sample sizes of approximately 3 mg.
Inside the reactor a known temperature gradient
was maintained, which gives the possibility to determine the location where the decomposition of the precursor occurs. Tsyganova et al. (1992) reported that
the Cu(acac)2 vapor decomposition was a first-order
rate reaction. The rate constant of this reaction can
be calculated by using the Arrhenius equation k =
k0 exp (Ea /RT ), where the pre-exponential factor is
k0 = 3.02107 s1 , the activation energy is Ea =
115.4 kJ/mole, and R is the universal gas constant. It is
obvious that most of the copper acetylacetonate vapor

thermolysis occurs in the vicinity of the highest temperature zone in the furnace. In Figure 2 the dependence of temperature and the rate constant of Cu(acac)2
decomposition reaction inside the reactor are shown.
On the basis of data presented in Figure 2, the decomposition of the metalorganic precursor occurs at location x = 255 12 mm, and at the temperature of
t = 431.5 0.5 C.
Determination of Cu(acac)2
vapor concentration
It is known that the vaporization rate of the solid
precursor changes with time due to decreasing surface area of the powder (Chou & Tsai, 1994;

389

Figure 2. The dependence of temperature and the rate constant of Cu(acac)2 decomposition reaction inside the reactor.

Figure 3. The dependence of the evaporation rate at different conditions.

Kodas & Hampden-Smith, 1999). Certainly, the vapor


pressure of the precursor is a crucial factor in the
decomposition reaction. It can be a reason for the irreproducibility of the experimental results and, as it will
be found out later, it is even the reason for a change of
decomposition reaction products. Therefore, studies of
gas flow saturation by Cu(acac)2 vapor are necessary.
The usage of the removable cartridge (Figure 1)
allows determining of the saturation degree of the carrier gas flow by Cu(acac)2 vapor. If the flow rate of
the gas and the mass difference of the cartridge with

Cu(acac)2 powder in a certain time interval are known,


it is possible to calculate the evaporation rate and hence,
the vapor pressure of the metalorganic compound at
the entrance of the reactor. From Figure 3 it can be seen
that the evaporation rate decreases with experimental
time. The gas saturation by Cu(acac)2 vapor decreases
very rapidly at a saturator temperature of tsat = 190.0 C
and at a flow rate of Q = 2000 cm3 /min. Because of
high gas velocity, the saturation time is too short to heat
the carrier gas in the upper part of the saturator. The
observation of saturator mixture color showed that the

390
precursor was consumed in the lower part where gas
temperature reached the saturator temperature. At the
flow rate of Q = 330 cm3 /min and at tsat = 170 C,
the evaporation decreases less rapidly and depends on
the decrease of the powder surface area. Decreasing
the saturator temperature down to 150 C leads to the
significant decrease of the evaporation rate dependence
on time. Four days of continuous operation using the
flow rate of Q = 330 cm3 /min revealed that a significant change in the results (concentration and number
size distribution of agglomerated particles) was found
only after the first 24 h period.
The effect of the composition of the copper acetylacetonate and silicon dioxide mixture on the evaporation rate was also investigated. Variation of the mass
ratio of Cu(acac)2 and SiO2 from 1 : 3 to 1 : 150 did
not significantly affect on the precursor vapor pressure at the entrance to the reactor. The influence of
using a small precursor fraction (1 : 150 relation) was
found only after about 7 h because of the exhaustion
of Cu(acac)2 material, while duration of an experiment was only about 3040 min. During the following experiments only newly prepared powder mixtures
consisting of 4 g (1.53102 mole) of copper (II) acetylacetonate and 16 g (0.27 mole) of silicon dioxide have
been used as the saturator mixture.
The dependence of vapor pressure, as determined
from the cartridge mass difference, on the saturator temperature at Q = 330 cm3 /min is presented in
Figure 4. Also, a fitted curve and literature data of

Cu(acac)2 equilibrium vapor pressure (Teghil et al.,


1981; Tonneau et al., 1995) are shown. From the figure,
a difference between the literature data and our results
can be seen. It is necessary to note that the measured
saturator temperature does not coincide with the powder temperature inside the saturator and has been used
only as a reference temperature during the experiments.
The main crucial experimental parameter is the precursor concentration, which has been determined after
each experiment. Thus, during our experiments, wellcontrolled parameters of the setup regarding the contents of vaporgas phase have been maintained.
Results and discussion
Experimental results
Experiments on particle formation were carried out at
two fixed temperature profiles inside the furnace. One
of these profiles is shown in Figure 2. Experiments
were also performed with a similar temperature profile
with a temperature maximum at t = 705.0 0.5 C,
where this condition was valid at x = 255 15 mm.
In the following section, we refer to those temperature
profiles as 431.5 C and 705.0 C.
First, experiments on the influence of residence
time (flow rate) on particle number size distributions were carried out. At the furnace temperature of
tfurn = 431.5 C, at flow rates higher than 400 cm3 /min,

Figure 4. The dependence of the precursor vapor pressure on the saturator temperature.

391

Figure 5. Number particle size distributions at the different flow rates.

a bimodal particle size distribution was measured


(Figure 5). SEM and TEM micrographs showed two
kinds of particles small agglomerated copper particles
and needle-like particles of Cu(acac)2 . Thus, the mode
with the smaller mean size is connected to copper particle formation and the second one is due to nucleation
of undecomposed Cu(acac)2 vapors downstream of the
furnace. The variation of the flow rates at the furnace
temperature of 705.0 C showed that the optimum flow
rate was about 330 cm3 /min. The choice of this flow rate
was determined by the concentration range of the condensation particle counter (1105 particles/cm3 ) and by
reasonable time of a sample collection for XRD analysis. Hereafter, only experimental results obtained at
the flow rate of Q = 330 cm3 /min are presented.
TEM micrographs of the nanoparticles synthesized
at reactor temperature of 431.5 C and at the precursor vapor pressures of Pprec = 1.8 and Pprec = 16 Pa
are presented in Figure 6. As one can see, the primary
aerosol particles are fairly monodisperse and the size
of the particles increases with increasing saturator temperature. In the same figure, the electron diffraction
patterns of the agglomerated particles are presented.
The electron diffraction ring pattern simulations performed for copper fits with the experimental results.
These results have been confirmed by XRD analysis.
TGA showed that at the furnace temperature of 431.5 C
the decomposition of the precursor was not complete.

At the furnace temperature of 431.5 C only 20% of


Cu(acac)2 was thermolyzed. The remaining undecomposed Cu(acac)2 forms a layer around copper particles
that can be seen on the TEM images (Figure 6). It is
worth noting that the electron diffraction ring pattern
enclosed in Figure 6(a) is not so clear as the others
due to the large amount of amorphous unreacted precursor. Microdiffraction patterns for this experimental
condition have been also performed in order to determine unambiguously that the particles presented in
Figure 6(a) are crystalline copper.
TEM results for the nanoparticles synthesized
at 705.0 C and at the precursor vapor pressures of
Pprec = 1.8 and Pprec = 16 Pa are presented in Figure 7.
With these conditions, the size of the primary particles decreases with increasing the vapor pressure
of the precursor, that is, there is an inverse situation
compared to the preceding samples synthesized at
the furnace temperature of tfurn = 431.5 C. Electron
diffraction patterns of the agglomerated particles produced at the furnace temperature of tfurn = 705.0 C
indicate the presence of crystalline copper particles
at the vapor pressure of Pprec = 16 Pa and copper
oxide (Cu2 O, cuprite) particles at the vapor pressure
of Pprec = 1.8 Pa. In Figure 8 XRD diffractograms of
the particles collected on a silver filter are shown. The
variation of the crystalline phase with changing the
precursor vapor pressure can be seen. At the vapor

392

Figure 6. Transmission electron microscopy images and electron


diffraction patterns of particles produced at the precursor vapor
pressure of Pprec = 1.8 Pa (a) and Pprec = 16 Pa (b). The furnace
temperature is 431.5 C.

pressure of Pprec = 1.8 Pa and Pprec = 6 Pa, a mixture


of crystalline copper and copper (I) oxide particles
was synthesized and at higher saturator temperatures
only copper particles were formed.
The additional experiments of XRD phase identification were carried out at the intermediate furnace
temperature of tfurn = 596.0 C. At the precursor vapor
pressures of Pprec = 6 and 44 Pa only copper crystalline
product was found.
The effect of the experimental conditions on product
compositions is presented in Table 1. As one can see
from Table 1, the dependence of the decomposition

Figure 7. Transmission electron microscopy images and electron


diffraction patterns of particles produced at the precursor vapor
pressure of Pprec = 1.8 Pa (a) and Pprec = 16 Pa (b). The furnace
temperature is 705.0 C.

product on the precursor vapor pressure is revealed only


at highest experimental temperature of tfurn = 705.0 C.
The possible reasons of this phenomenon are examined
in the discussion part of the article.
In an attempt to reduce agglomeration, the aerosol
flow was diluted by excess nitrogen flow immediately downstream of the furnace. The particle number size distribution was not significantly changed,
indicating that agglomeration of the particles occurred
inside the furnace. Another way to avoid the formation of the agglomerated particles is to decrease
the number concentration of the primary particles by

393

Figure 8. XRD spectra of particles synthesized at the furnace temperature of 705.0 C and collected on the silver filter.
Table 1. List of experimental conditions, corresponding
crystalline product compositions, and methods used for the
phase identification
Furnace
temperature, C

Vapor pressure
of precursor, Pa

Crystalline
products

Method

431.5
596.0
705.0
705.0
705.0

0.1544
644
0.061.0
1.810
1644

Cu
Cu
Cu2 O
Cu/Cu2 O
Cu

XRD, ED
XRD
ED
XRD, ED
XRD, ED

resolution picture shows the possibility to produce


single crystalline individual nanoparticles.

Discussion

decreasing the Cu(acac)2 vapor pressure. In Figure 9


and Figure 10 the number size distribution of the
produced particles at different saturator temperatures
of tfurn = 431.5 C and tfurn = 705.0 C are presented. It
is known that the area limited by the curve determines
the particle number concentration. At tfurn = 431.5 C
the concentration grows smoothly when the saturator temperature increases, but at tfurn = 705.0 C the
change of the concentration is not smooth. This is
probably in connection with the formation of different
decomposition products.
At precursor vapor pressures below 0.1 Pa, single
crystalline individual particles are formed, the formation of agglomerated particles is observed at higher
vapor pressures. In Figure 11, TEM micrograph of
copper (I) oxide nanoparticle that was produced at
the furnace temperature of 705.0 C and at the vapor
pressure of Pprec = 0.04 Pa is presented. This high

The size distribution of primary particles obtained


from TEM micrographs is presented in Figure 12. At
431.5 C the number mean size of the primary particles
is increased from Dp = 3.7 (with geometric standard
deviation of g = 1.42) to Dp = 7.2 nm (g = 1.33)
with increasing precursor vapor pressure from 1.8 to
16 Pa. At tfurn = 705.0 C, the primary particle size
is decreased from Dp = 24.0 nm (g = 1.57) to
Dp = 7.6 nm (g = 1.54), respectively. At the furnace temperature of tfurn = 431.5 C, increasing the
vapor pressure increased the size of the primary particles, but at tfurn = 705.0 C an inverse situation was
observed. Apparently, it is connected with the different products of the precursor decomposition. The
diameter of the primary copper (I) oxide particles
(tfurn = 705.0 C, Pprec = 1.8 Pa) is about three times
larger than that of the copper particles (tfurn = 705.0 C,
Pprec = 16 Pa). The qualitative explanation of the primary size dependence on the crystalline products can
be obtained from the consideration of physical properties of these compounds. It is known that the sintering
rate, which controls the size of primary particles,
is dependent on the self-diffusion coefficient of the
elements contained in the substance: with the larger

394

Figure 9. Number size distributions of agglomerated particles at the furnace temperature of tfurn = 431.5 C.

Figure 10. Number size distributions of agglomerated particles at the furnace temperature of tfurn = 705.0 C.

coefficient the sintering rate is enhanced and the size


of primary particles becomes larger. According to the
literature data (Smithells & Brandes, 1983; Kofstad,
1972), the self-diffusion coefficients of copper atoms
are 1.91012 cm2 /s in pure crystalline copper and
1.51010 cm2 /s in crystalline copper (I) oxide at
temperature t = 705 C. The almost two orders of
magnitude of difference in the diffusion coefficients
is most likely the reason for the formation of such a

different size of copper and copper (I) oxide primary


particles.
As it was found, the product of the decomposition
reaction depended on the precursor vapor pressure only
at tfurn = 705.0 C: copper particles were formed at
the higher vapor pressures (Pprec > 10 Pa), copper (I)
oxide particles at the pressures of Pprec < 1 Pa, and a
mixture of Cu and Cu2 O was formed at the intermediate vapor pressures. At the furnace temperatures of

395
431.5 C and 596 C, only copper particles were formed
(Table 1). At the first glance, the revealed dependence
of oxygen content in particles upon the precursor pressure seems to be anomalous. Indeed the formation of

Figure 11. High resolution transmission electron micrograph of


Cu2 O particle produced at the furnace temperature of 705.0 C
and at the vapor pressure of Pprec = 0.04 Pa.

Figure 12. Particle number size distributions of primary particles.

primary particles can be presented by two sequential


steps. At the first stage, small clusters are formed by
means of homogeneous nucleation process, and secondly, the particles grow because of a coalescence of
the clusters and vapor condensation on the particles.
Increasing the precursor pressure leads to the increase
of the oxygen-containing gas pressure in the system
and its participation in these two stages. The presence
of oxygen in the particles with increasing the precursor pressure at least cannot vanish. Let us demonstrate
that this prima facie consideration is incorrect for the
explanation of the experimental results.
A few possible reasons for the change of decomposition products were examined. The first reason
could be the presence of impurities in the carrier
gas. The appearance of the copper (I) oxide in the
crystalline products occurs at the precursor vapor
pressure of 10 Pa, which corresponds to a concentration of 2.61015 molecules/cm3 . Meanwhile, the
maximum concentration of admixtures in the nitrogen
carrier gas sample is about 2.51016 molecules/cm3 ,
that is, the concentration of the main component in
the reaction is of the same order or less than the
impurity concentration. In order to check this hypothesis, the nitrogen gas cylinder was changed to a
liquid nitrogen tank to obtain ultra-pure carrier gas
(99.999 vol.% with maximum oxygen concentration
of 8.11013 molecules/cm3 ). Decreasing the content of
admixtures by two orders of magnitude did not change

396
the product of the Cu(acac)2 decomposition. Moreover,
the results by Tonneau et al. (1995) and Hammadi et al.
(1995) on chemical vapor deposition are also contrary
to this assumption. They showed that when the vapor
mole fractions of oxygen and Cu(acac)2 were about
the same, only crystalline copper was formed. Accordingly, the presence of impurities in the gas-carrier is
not likely to be the reason for the different products of
the reaction.
Another possible reason is connected with the kinetics of the decomposition. It is well known that changing the gas phase concentration of the reactant can
lead to the change of the decomposition mechanism
(Kondratev, 1964). For example, at lower precursor concentrations a unimolecular decomposition can
occur and as a result of the decomposition Cu2 O is
formed. Increasing the vapor pressure leads to the
increase in the probability of collisions of two or more
molecules during the Cu(acac)2 thermolysis reaction,
which leads to the decomposition product of pure
copper. However, in order to form Cu2 O molecule
it is necessary to have two precursor molecules, but
for copper formation only one precursor molecule
is needed. Thus, this explanation is correct for the
inverse behavior. Moreover, it is hard to believe that
switching the mechanism from a unimolecular to a collision reaction can occur in a such small vapor pressure range. Therefore, the explanation of switching the
decomposition mechanism does not seem to be justified for the elucidation of the variation of the reaction
products.
Also, a possible reason of the condensed product
variation may be related to the kinetics of particle
formation. Then the explanation of the phenomena is
possible only in the case of any second-order reaction. Indeed, the ratio of partial pressures of all species
formed in the first-order reactions is the same at all precursor pressures. Thereby, the relative role of different
gaseous species in the condensed particle growth cannot depend on the precursor pressure. The possible reason for the crystalline product change is the existence
of a secondary reaction. We assume that this reaction
is a formation of copper dimers that are quite stable at
the experimental conditions (Petrov, 1986):
2Cu(g) Cu2 (g).

(2)

In order to explain the revealed transition between the


crystalline products let us consider the mechanism of
the formation of different condensed substances from
gas phase in detail.

It is worth noting that there is a restriction prohibiting the growth of copper particles from gas via single
Cu atom adsorption on the cluster surface. The physical
fundamentals of such restrictions for different materials are discussed by Altman et al. (2001). They are
based on the mechanism of energy transfer during a
phase transition. The energy release during Cu atom
condensation is about 3.5 eV, and this energy has to
be dissipated directly during the adsorption. There are
two possible ways for this process: an electron excitation and a phonon creation. Since this energy value
is large compared to a typical Debye energy value
(0.05 eV), only the electron excitation for energy dissipation is left. However, the energy band structure of
copper contains the sp-band gap (Knoesel et al., 1998).
The value of the sp-band gap energy at the experimental conditions is about 4.64.7 eV. The existence of this
gap leads to impossibility of the direct energy dissipation from an adsorbing copper atom and to a prohibition of the atom condensation as a consequence.
The energy release for Cu2 molecule condensation is
about 5 eV, which is larger than the copper energy
gap. Therefore, this process becomes feasible and the
growth of the particles is made possible by condensation of Cu2 molecules. Then the reason of the crystalline product change can be understood. Indeed, due
to the second-order reaction (2), the Cu2 partial pressure increases with the precursor pressure as its second
power instead of all other gas partial pressures. Thus,
since the condensation of Cu2 molecules determines
the copper particle formation, the relative role of this
process increases with increasing precursor pressure
faster when compared to other processes.
The second question to be discussed is the possibility
of copper (I) oxide formation. Because of low reagent
concentration, the variety of possible pathways can be
limited by bimolecular reactions. In this case, the only
way of copper (I) oxide formation is the reaction on the
surface of a growing oxide particle:
Cu2 + CO2 Cu2 O(s) + CO.

(3)

It is obvious that this reaction should be activated.


Let us propose the physical model of the transition.
The formation of copper (I) oxide particle occurs if the
flux of CO2 molecules (with an energy larger than the
barrier Eo ) jCO2 on a particle surface is larger than
the flux of Cu2 molecules jCu2 , in the opposite case
copper particle formation occurs. Thereby, transition
between the two products might occur due to the variation of these two fluxes. At a given temperature of T

397

(the transition pressure of Cu2 O product


pressure Pprec
identification) can be written as

the fluxes may be defined as


jCu2 = 
jCO2

PCu2

,
2mCu2 kB T


PCO2
Eo
=
,
exp
kB T
2mCO2 kB T

(4)

where mCu2 and mCO2 are masses of Cu2 and CO2


molecules, respectively, and kB is Boltzmanns constant. It is obvious that the copper particle growth
leads to the exhaustion of the copper vapor and to the
decrease of relative value of jCu2 flow. The initial ratio
of jCu2 and jCO2 flows is more than unity (in case of
pure copper formation), but at the end of the precursor
decomposition this ratio becomes smaller (in case of
copper (I) oxide formation). That is why there is a transition region instead of the sharp border between the
different crystalline products. The absence of copper (I)
oxide (at Pprec > 10 Pa) most likely means that the ratio
of amount of the oxide to pure copper is less than the
relative sensitivity f of XRD phase identification. In
this case, at the precursor pressure of Pprec = 16 Pa, the
copper (I) oxide particles are formed only at the end of
the precursor decomposition, where the jCu2 /jCO2 flux
ratio becomes smaller than unity due to the exhaustion
of copper vapor and an accumulation of the decomposition products (carbon dioxide). In order to obtain the
characteristic value of the precursor pressure where the
transition between products occurs, let us consider the
final location of the decomposition. The partial pressure of carbon dioxide can be written as
PCO2 = 2Pprec ,

(5)

 2



Kp fPprec
2Pprec
Eo
=
exp
.

mCu2
mCO2
kB T

(8)

Based on the experimental data of the disappearance

= 16 Pa at tfurn =
of copper (I) oxide phase (Pprec

705.0 C) and thermodynamic data (Kp = 1.84 Pa1 ), at


the sensitivity of f = 0.01, the value of the energy barrier for reaction (3) is Eo = 0.595 eV (57.3 kJ/mole).
This value seems to be realistic. Then, it is easy to
show that the boundary pressure of copper (I) oxide
formation with mole fraction of less than 0.01 may be
calculated as


const
Eo

Pprec =
,
(9)
exp
Kp
kB T
where const = 3.4 104 . The semiempirical temperature dependence of the boundary pressure obtained
on the basis of Eq. (9) is presented in Figure 13. The
area on the diagram above the boundary pressure line
is related to the region of the absence of copper oxide
products. As one can see the proposed semiempirical
phase diagram is in agreement with the experimental results of the current work. The boundary pressure

Pprec
decreases by decreasing the furnace temperature.
That is why at lower temperatures (tfurn = 431.5 C at
Pprec = 0.06 to 44 Pa and tfurn = 596 C at Pprec =
1644 Pa) only pure copper particles in the final product of Cu(acac)2 decomposition were obtained. It is
worth noting that the diagram could be useful for the

where the factor 2 is obtained on the basis of


mass-spectrometry results of gaseous decomposition
products (Tsyganova et al., 1992). The copper vapor
pressure at the end of the decomposition, when
copper (I) oxide starts to be formed, can be written by
taking into account the copper exhaustion and phase
identification sensitivity f :
PCu = fPprec .

(6)

The partial vapor pressure PCu2 can be found using equilibrium constant Kp of the reaction (2):
2
PCu2 = Kp PCu
.

(7)

Combining Eqs. (4)(7) with the condition jCu2 =


jCO2 , the expression for the boundary precursor

Figure 13. The phase diagram of Cu(acac)2 decomposition crystalline products. Marks and correspond to the experimentally determined copper and copper (I) oxide products,
respectively.

398
the precursor decomposition and copper and copper (I)
oxide particle formation. After the Cu(acac)2 vapor
evaporation and heating the vapor up to a high enough
temperature, the formation of copper vapor and decomposition products, as a result of Cu(acac)2 decomposition reaction, occurs inside the furnace. The next
important stage is the formation of gaseous copper
dimers that participate in nucleation and condensation processes. Formation of copper (I) oxide particles
occurs at low precursor vapor pressures due to the surface reaction of Cu2 vapor and the products of the
decomposition (mainly, carbon dioxide). The last stage
is the agglomeration process of the formed primary particles, which exist at the precursor vapor pressure of
Pprec 0.1 Pa.
Conclusion

Figure 14. The schematic presentation of copper and copper (I)


oxide particle formation by the Cu(acac)2 vapor decomposition.

prediction of the product of the precursor vapor decomposition at other experimental conditions. The proposed approach for the building of the decomposition
product phase diagram can be used for other experimental systems with chemical reactions.
Figure 14 summarizes our viewpoint of the main
stages occurring inside the furnace at 705.0 C during

For this study, a vertical laminar flow reactor has been


constructed and tested for the investigation of nanoparticle formation via chemical vapor nucleation. It has
been shown that crystalline nanometer-size copper and
copper oxide particles can be produced by thermal
vapor decomposition of a metalorganic precursor,
Cu(acac)2 , at relatively low temperatures. Individual
primary particles are formed at the precursor vapor
pressure of Pprec < 0.1 Pa. At higher vapor pressures,
particles form aggregates. At tfurn = 431.5 C, the number mean size of the primary particles increased from
Dp = 3.7 nm (with geometric standard deviation of
g = 1.42) to Dp = 7.2 nm (g = 1.33) with increasing precursor vapor particle pressure from 1.8 to 16 Pa.
At tfurn = 705.0 C, the primary particle size decreased
from Dp = 24.0 nm (g = 1.57) to Dp = 7.6 nm
(g = 1.54), respectively.
At the furnace temperatures of 431.5 C and 596.0 C,
only crystalline copper particles were produced. At the
furnace temperature of tfurn = 705.0 C, the product of
the decomposition reaction depended on the precursor
vapor pressure: copper particles were formed at vapor
pressures higher than 10 Pa, copper (I) oxide at pressures lower than 1 Pa, and a mixture of the metal and
its oxide at intermediate vapor pressures. For the explanation of the obtained results, the kinetic restriction on
copper particle growth was proposed. It leads to the
main role of a Cu2 molecule participation in the particle
formation. The formation of copper (I) oxide particles
occurs due to the surface reaction of the decomposition
products of which carbon dioxide is the most important. For the explanation of the experimental results,

399
a model is proposed to build a semiempirical phase diagram of the precursor decomposition crystalline products. The phase diagram, showing the boundary of the
appearance of copper (I) oxide, is in agreement with
the experimental results.
Acknowledgements
This work has been supported by TEKES, Finland
and VTT Chemical Technology via the MATRA
research program. Mr. P. Raisanen, Dr. M. Ritala and
Prof. M. Leskela are gratefully acknowledged for help
during the XRD analyses and Mr. Timo Hatanpaa
for carrying out TGA analysis. The authors thank
Dr. J.K. Jokiniemi, Dr. P.V. Pikhitsa, Dr. V. Mikheev
and Prof. T. Ward for fruitful discussions and Dr.
D.P. Brown for discussions and proofreading the
manuscript.
References
Adams C.R. & T.J. Jennings, 1964. Mechanism studies of the
catalytic oxidation of propylene. J. Catal. 3, 549.
Altman I.S., P.V. Pikhitsa & M. Choi, 2001. Key effects at
nanoparticle formation by combustion techniques. In: Kish L.,
Granqvist C.G., Marlow W., Siegel R.W. eds. Fundamentals of
Gas-Phase Nanotechnology (in press).
Ayaappan S., R.S. Gopalan, G.N. Subbanna & C.N.R. Rao, 1997.
Nanoparticles of Ag, Au, Pd, and Cu Produced by Alcohol
Reduction of the Salts. J. Mater. Res. 12, 398.
Bouland D., J.C. Chouard, A. Briand, F. Chartier, J.L. Lacour,
P. Mauchien & J.M. Mermet, 1992. Experimental study of
aerosol production by laser ablation. J. Aerosol Sci. 23(1), 225.
Bowles R.S., J.J. Kolstad, J.M. Calo & R.P. Andres, 1981.
Generation of molecular clusters of controlled size. Surf. Sci.
106, 117.
Campbell C.T., K.A. Daube & J.M. White, 1987. Cu/ZnO(0001)
and Cu(111): Model Catalysts for Methanol Synthesis. Surf.
Sci. 182, 458.
Champion Y. & J. Bigot, 1996. Characterization of nanocrystalline copper powders preparing by melting in a cryogenic
liquid. Mater. Sci. Eng. A217/218, 58.
Chou K.-S. & G.-J. Tsai, 1994. Dynamic evaporation behaviour
of diketonate compounds of yttrium, copper and barium.
Thermochim. Acta 240, 129.
Daroczi L., M.T. Beck, D.L. Beke, M. Kis-Varga, L. Harasztosi &
N. Takacs, 1998. Production of Fe and Cu nanocrystalline particles by thermal decomposition of ferro- and copper-cyanides.
Mat. Sci. Forum 269272, 319.
Dhas N.A., C.P. Raj & A. Gedanken, 1998. Synthesis, characterization, and properties of metallic copper nanoparticles. Chem.
Mater. 10, 1446.
Ding J., T. Tsuzuki, P.G. McCormick & R. Street, 1996. Ultrafine
Cu articles prepared by mechanical process. J. Alloys and
Comp. 234, L1.

Du F.-L., Z.-L. Cui, Z.-K. Zhang & S.-Y. Chen, 1997. Behavior
of supported nano-copper catalyst in CO oxidation. J. Nat. Gas
Chem. 6, 135.
& V.A. Uglov, 1991. Nanocrystalline copper powFolmanis G.E.
der produced by electrolysis. Poroshkovaya Metallurgiya 2, 5.
Gerfin T., M. Becht & K.-H. Dahmen, 1993. Preparation of copper and copper oxide film by metal-organic chemical vapour
deposition using -ketoiminato complexes. Mater. Sci. Eng.
B17, 97.
Girardin D. & M. Maurer, 1990. Ultrafine metallic powders prepared by high pressure plasma: synthesis and characterization.
Mat. Res. Bull. 25, 119.
Gredig S.V., R.A. Koppel & A. Baiker, 1997. Synthesis of methylamines from carbon dioxide, hydrogen and ammonia over supported copper catalysts. Influence of support. J. Mol. Catal. A:
Chemical 127, 133.
Haas V. & R. Birringer, 1992. The morphology and size of nano
structured Cu, Pd, and W generated by sputtering. Nanostruct.
Mater. 1, 491.
Hammadi Z., B. Lecohier & H. Dallaporta, 1993. Chemical vapor
deposition of metallic copper film in the presence of oxygen.
J. Appl. Phys. 73, 5213.
Hammerschmidt W., A. Baiker, A. Wokaun & W. Fluhr, 1986.
Copper catalyzed synthesis of cyclic amines from aminoalcohols. Appl. Catal. 20, 305.
Herley P.J., W. Jones & G.R. Millward, 1989. Electron beam
decomposition of copper hydride and the generation of ultrafine particles of copper. J. Mater. Sci. Lett. 8, 1013.
Holzschuh H. & H. Suhr, 1990. Deposition of copper oxide
(Cu2 O, CuO) thin films at high temperatures by plasmaenhanced CVD. Appl. Phys. A51, 486.
Huang H.H., F.Q. Yan, Y.M. Kek, C.H. Chew, G.Q. Xu, W. Ji,
P.S. Oh & S.H. Tang, 1997. Synthesis, characterization, and
nonlinear optical properties of copper nanoparticles. Langmuir
13, 172.
Kaito C., K. Fujita & H. Hashimoto, 1973. Electron-microscopy
study of oxidation processes by metal fine particles. Jpn. J.
Appl. Phys. 12, 489.
Kaito C., T. Watanabe, K. Ochtsuka, F. Chen & Y. Saito, 1993.
A new study on the production of clusters from copper ultrafine
particles. J. Cryst. Growth 128, 267.
Kashu S., M. Nagase, C. Hayashi, R. Uyeda, N. Wada & A. Tasaki,
1974. Preparation and properties of ultrafine metal powders.
Proc. 6th Int. Vac. Congr. 1974. Jpn. J. Appl. Phys. 14 (Suppl.
2), 491.
Kellerson A., E. Knozinger, W. Langel & M. Giersig, 1995. Cu2 O
quantum-dot particles prepared from nanostructured copper.
Adv. Mater. 7, 652.
Kijenski J., J. Burger & A. Baiker, 1984. Copper catalyzed disproportionation of benzylamine methyl derivatives. Appl. Catal.
11, 295.
Kijenski J., P.J. Niedzielski & A. Baiker, 1989. Synthesis of cyclic
amines and their alkyl derivatives from amino-alcohols over
supported copper catalysts. Appl. Catal. 53, 107.
Kirchheim R., X.Y. Huang, P. Cui, R. Birriner & H. Gleiter, 1991.
Free energy of active atoms in grain boundaries of nanocrystalline copper, nickel and palladium. Nanostruct. Mater. 1, 167.
Klenov D.O., G.N. Kryukova & L.M. Plyasova, 1998. Localization of copper atoms in the structure of the ZnO catalyst for
methanol synthesis. J. Mater. Chemistry 8, 1665.

400
Klier K., 1982. Methanol synthesis. Adv. Catal. 31, 243.
Knacke O., O. Kubaschewski & K. Hesselmann, 1991.
Thermochemical Properties of Inorganic Substances.
Springer-Verlag, Berlin.
Knoesel E., A. Hotzel & M. Wolf, 1998. Temperature dependence
of surface state lifetimes, dephasing rates and binding energies
on Cu(111) studies with time-resolved photoemission. J. Elect.
Spectr. Rel. Phen. 8891, 577.
Kodas T.T. & M.J. Hampden-Smith, 1999. Aerosol Processing of
Materials. Wiley-VCH, New York.
Kofstad P., 1972. Nonstoichiometry, Diffusion, and Electrical
Conductivity in Binary Metal Oxides. Wiley-Interscience,
New York.
Kondratev V.N., 1964. Chemical Kinetics of Gas Reactions.
Persimmon Press Oxford, London.
Kulikov I.S., 1969. Thermal Dissociation of Compounds.
(Termicheskaja dissociatciya soedinenij) Metallurgia, Moscow
(in Russian).
Lisieski I. & M.P. Pileni, 1993. Synthesis of copper clusters using
reverse micelles as microreactors. J. Am. Chem. Soc. 115,
3887.
Lisieski I. & M.P. Pileni, 1995. Copper metallic particles synthesized in situ in reverse micelles: influence of various parameters on the size of particles. J. Phys. Chem. 99, 5077.
Lisieski I., F. Billoudet & M.P. Pileni, 1996. Control of the shape
and the size of copper metallic particles. J. Phys. Chem. 100,
4160.
Long N.J. & A.K. Petford-Long, 1986. In situ Electron-beaminduced reduction of CuO: a study of phase transformation in
cupric oxide. Ultramicroscopy 20, 151.
Majumdar D., T.A. Shefelbine, T.T. Kodas & H.D. Glicksman,
1996. Copper (I) oxide powder generation by spray pyrolysis.
J. Mater. Res. 11, 2861.
Maruyama T. & T. Shirai, 1995. Copper thin film prepared by
chemical vapour deposition from copper (II) acetylacetonate.
J. Mater. Sci. 30, 5551.
Marzouk H.A., J.S. Kim, P.J. Reucroft, R.J. Jacob,
J.D. Robertson & C. Eloi, 1994. Evaluation of copper
chemical-vapor-deposition films on glass and Si(100)
substrates. Appl. Phys. A58, 607.
Moini M. & J.R. Eyler, 1988. Formation of small negative and
positive cluster ions of gold, silver, and copper by direct laser
vaporization. J. Chem. Phys. 88, 5512.
Pauleau Y. & A.Y. Fasasi, 1991. Kinetics of sublimation of
copper (II) acetylacetonate complex used for chemical vapor
deposition of copper films. Chem. Mater. 3, 45.
Pelletier J., R. Pantel, C. Oberlin, Y. Pauleau & P. Gouy-Pialler,
1991. Preparation of copper films at ambient temperature
by microwave plasma-enhanced chemical vapor deposition
from the copper (II) acetylacetonate-argon-hydrogen system.
J. Appl. Phys. 70, 3862.
Peoples S.M., J.F. McCarthy, L.C. Chen, D. Eppelsheimer &
M.O. Amdur 1988. Copper oxide aerosol: generation and
characterization. Am. Ind. Hyg. Assoc. J. 49, 271.
Pereia R., M. Rufo & U. Schuchardt, 1994. Copper (II) catalyzed
oxidation of cyclohexane by tert-butyl hydroperoxide. J. Braz.
Chem. Society 5, 83.

Petrov Yu.I., 1986. Clusters and Small Particles. (Klastery i malye


chastitcy) Nauka, Moscow (in Russian).
Pietrikova A. & E. Kapusanska, 1991. Production of very fine
copper powder and control of its properties. Metallic Mater.
29, 626.
Runeberg J., A. Baiker & J. Kijenski, 1985. Copper catalyzed
amination of ethylene glycole. Appl. Catal. 17, 309.
Shannon I.J., F. Rey, G. Sankar, J.M. Thomas, T. Maschmeyer,
A.M. Waller, A.E. Palomares, A. Corma, A.J. Dent &
G.N. Greaves, 1996. Hydrotalcite-derived mixed oxides containing copper: catalysts for the removal of nitric oxide.
J. Chem. Soc. Farad. Trans. 92, 4331.
Shiau C.-Y. & J.C. Tsai, 1997. Cu/r-Al2O3 Catalyst prepared by
electrolyses method. J. Chin. Ist. Chem. Eng. 28, 55.
Smithells C.J. & E.A. Brandes, 1983. Smithells Metals Reference
Book. Butterworths, London.
Takahashi T., J. Suzuki, M. Saburi & Y. Uchida, 1988. Selective
preparation of copper, copper (I) oxide, or copper (II) oxide
fine particles from organocopper compounds. J. Mater. Sci.
Let. 7, 1251.
Teghil R., D. Ferro, L. Bencivenni & M. Pelino, 1981. A thermodynamic study of the sublimation processes of aluminium and
copper acetylacetonates. Thermochim. Acta 44, 213.
Tijburg I.I.M., Ph.D. Thesis, Utrecht, The Netherlands, 1989.
Tonneau D., R. Pierrisnard, H. Dallaporta & W. Marine, 1995.
Growth kinetics of copper films from photoassisted CVD of
copper acetylacetonate. Journal de Physique IV 5(C5), 629.
Tsyganova E.I., G.A. Mazurenko, V.N. Drobotenko,
L.M. Dyagileva & Yu.A. Aleksandrov, 1992. Kinetic regularities of thermolysis of yttrium, barium, copper acetylacetinates.
J. Gen. Chem. 62, 407 (Translated from Zh. Obshch. Khim.
(USSR) 62, 499).
Voge H.H. & C.R. Adams, 1967. Catalytic oxidation of olefins.
Adv. Catal. 17, 151.
Vorobyova S.A., V.V. Mushinsky & A.I. Lesnikovich, 1997.
Copper oxide produced by two-phase synthesis in octanewater system. Dokl. Acad. Sci. Belarus 41, 62.
Vultier R., A. Baiker & A. Wokaun, 1987. Copper catalyzed
amination of 1,6-hexanediol. Appl. Catal. 30, 167.
Van der Meijden J., Ph.D. Thesis, Utrecht, The Netherlands, 1981.
Weins W.N., J.D. Makinson, R.J. De Angelis & S.C. Axtell,
1997. Low-frequency internal friction studies of nanocrystalline copper. Nanostruct. Mater. 9, 509.
Winklmayr W., G.P. Reischl, A.O. Linder & A. Berner, 1991.
A new electromobility spectrometer for the measurement of
aerosol size distributions in the size range from 1 to 100 nm.
J. Aerosol Sci. 22, 289.
Xu J., X. Sun, W. Chen, X. Fan & W. Wei, 1992. Synthesis
of copper ultrafine particles by using gas evaporation. Cailiao
Kexue Jinzhan 6, 209.
Yurieva T.M., L.M. Plyasova, T.A. Krieger, V.I. Zaikovskii,
O.V. Makarova & T.P. Minyukova, 1993. State of copper containing catalyst for methanol synthesis in the reaction medium.
React. Kinet. Catal. Lett. 51, 495.

S-ar putea să vă placă și