Sie sind auf Seite 1von 10

J Mater Sci (2015) 50:73747383

DOI 10.1007/s10853-015-9295-3

Novel photoactive inorganic polymer composites of inorganic


polymers with copper(I) oxide nanoparticles
Mahroo Fallah1 Kenneth J. D. MacKenzie1 John V. Hanna2 Samuel J. Page2

Received: 26 May 2015 / Accepted: 22 July 2015 / Published online: 29 July 2015
Springer Science+Business Media New York 2015

Abstract Novel photoactive composites of cubic Cu2O


nanoparticles with aluminosilicate geopolymers were prepared and their structure was shown by SEM/EDS and
TEM. XRD, FTIR, 27Al and 29Si MAS NMR to consist of a
well-reacted geopolymer matrix containing homogeneously dispersed Cu2O nanoparticles. Under dark conditions, both the geopolymer matrix and the composites were
shown to adsorb a model organic compound (methylene
blue, chosen for its colour stability at high pH) by a process
which follows pseudo-first-order kinetics and can be
described by Langmuir-type isotherms. At concentrations
[20 wt%, the oxide decreases the adsorption rate by
blocking the active adsorption sites of the geopolymer.
Under UV radiation, the composites remove the methylene
blue by a combination of adsorption and photodegradation,
without deterioration of the geopolymer structure or the
photoactive Cu2O component, as evidenced by 63Cu NQR
spectroscopy, suggesting that these geopolymer composites
should function as useful new materials for the removal of
organic pollutants from water or the atmosphere.

& Kenneth J. D. MacKenzie


kenneth.mackenzie@vuw.ac.nz
1

MacDiarmid Institute for Advanced Materials and


Nanotechnology, School of Chemical and Physical Sciences,
Victoria University of Wellington, Wellington, New Zealand

Department of Physics, Warwick University,


Coventry CV4 7AL, UK

123

Introduction
Oxide semiconductors are of considerable interest for
applications in the fields of catalysis, optics, optoelectronics, biosensors and as materials for environmental
protection such as water disinfection, hazardous waste
remediation and air purification. The energy gap in a
semiconductor is a useful characteristic since it defines the
wavelength required to excite an electron from the valence
band to the conduction band [1]. Oxidation reactions such
as the destruction of harmful organic substances take place
in the valence band, whereas the conduction band is more
typically the site for reduction reactions [1]. Semiconductor
materials with wide band gaps are usually more stable than
those with narrow band gaps, but materials with narrow
band gaps couple more efficiently with the solar spectrum
and are, therefore, more suitable as photocatalysts. To date
TiO2, with a bandgap of 3.03 eV, has been the most
commonly used semiconductor for photo-induced reactions
although its bandgap is slightly too large to make efficient
use of the visible solar spectrum. Photocatalysts consisting
of TiO2 supported on a variety of substrates have been
reported [27], including a photoactive composite with a
metakaolin-based geopolymer prepared by ion exchange
with solutions of (NH4)2TiO(C2O4)2H2O, resulting in the
insertion of TiO2 particles (anatase) into the geopolymer
[8]. A metakaolin geopolymer containing TiO2 inserted by
an ion exchange process is reported to efficiently photodegrade a model volatile organic compound (2-butanone)
[9].
By comparison with TiO2, nanostructured copper(I) oxide is even more attractive, being an inexpensive readily
available p-type semiconductor with a direct band gap of
2.2 eV and a high optical absorption coefficient [10, 11].
Cu2O is also a powerful absorber of molecular O2 and can

J Mater Sci (2015) 50:73747383

scavenge photoelectrons, thereby restraining the combination of electrons and holes [12]. This suggests that Cu2O
should act as a stable photocatalyst for the photochemical
decomposition of H2O with the generation of O2 and H2
under visible light irradiation and should also be a good
candidate for the photocatalytic degradation of organic
pollutants under visible light irradiation. A further major
attraction of Cu2O is its low toxicity and good environmental acceptability, with a band gap that can be tuned by
adjusting its particle size [12]. The combination of ecologically friendly inorganic polymers and nanocrystalline
Cu2O for absorption and photodegradation of harmful
organic substances has not previously been investigated in
detail.
In this work, several approaches for the preparation and
characterization of Cu2O nanoparticles were investigated,
and the most promising of these products were inserted at
various concentrations into a metakaolinite-based inorganic polymer prior to curing. The resulting composites
were characterised by SEM, XRD, FTIR, 27Al and 29Si
MAS NMR and the role of the copper was investigated by
63
Cu nuclear quadrupole resonance (NQR). The photoactivity of the composites was evaluated by measuring the
bleaching of a model organic compound (methylene blue
dye) under dark and UV illumination by UVvisible
spectroscopy. Methylene blue was preferred for these
experiments since its colour is stable at the high pH of the
inorganic polymer matrix.

Experimental
Sample synthesis
The Cu2O nanoparticles were synthesised by the precipitation method of Bai et al. [13]. This synthesis was chosen
over other syntheses since it produces cubic crystallites of
nanometer size. An aqueous solution of copper acetate
(May and Baker AR grade) and polyvinylpyrrolidone
(PVP) (grade K30, mw = 10,000, Sigma-Aldrich) was
stirred vigorously, producing a blue suspension upon
dropwise addition of NaOH solution. Upon the further
dropwise addition of a solution of ascorbic acid (Fisher,
AR grade), the colour of the suspension changed from blue
to green and finally to orange after 30 min. The Cu2O
precipitate was separated by centrifuging, washed twice
with deionized water and dried in air.
The inorganic polymer matrix was prepared from New
Zealand kaolinite-type halloysite clay (Imerys Premium
Grade) dehydroxylated at 600 C for 24 h, sodium
hydroxide (BDH reagent grade) and sodium silicate solution (FERNZ Chemical Co, NZ, Type D, Na2O/
SiO2 = 0.48, solids content = 41.1 mass%). To these

7375

constituents, formulated to the molar oxide ratio SiO2/


Al2O3 = 3.14, Na2O/SiO2 = 0.34, Na2O/Al2O3 = 1.01
and H2O/Na2O = 10.12 was added 1030 wt% of the presynthesised cubic Cu2O nanoparticles and mixed for
15 min before transferring to cylindrical plastic moulds
and vibrating for 15 min to remove entrained air. The
composites were then cured for 12 h at ambient temperature, finely ground and stored in sealed plastic bags. In an
alternative approach, an attempt was made to synthesise
the geopolymer/Cu2O composites by mixing the matrix and
Cu2O-forming reagents in situ in a single step; this was
abandoned, however, because of the high water content
which militated against the formation of a well-reacted
geopolymer as determined by 27Al and 29Si MAS NMR
spectroscopy.
Sample characterization
The morphology of the Cu2O nanocrystals and the composites coated with a 7- to 9-mm carbon film using a
Quorum Q150T turbo-pumped carbon coater were
observed using a JEOL 6500F field emission SEM or a
JEOL JSM-6610 LA analytical SEM with an EDS attachment operated at 15 keV. TEM observations of the
geopolymer-Cu2O samples deposited on a gold grid were
made using a JEOL 2100F transmission electron microscope operated at 200 kV. The pore sizes and distributions
of the samples were determined from nitrogen adsorption/
desorption measurements at 77 K using a Micromeritics
model ASAP 2010 analyzer. XRD patterns were determined in the region 1080 2h with a 0.013 step using a
Bruker D8 Avance X-ray diffractometer with Ni-filtered
Cu Ka radiation operated at 45 kV and 40 mA. FTIR
spectroscopy was carried out on powdered samples suspended in KBr pellets in the range 4004000 cm-1 using a
Perkin-Elmer model 2000 FTIR spectrometer. Solid-state
27
Al and 29Si MAS NMR spectroscopy was used to study
the chemical environment of these elements in the system
before and after their use as photocatalysts. The spectra
were acquired at a magnetic field of 11.7 T using a Bruker
Avance III 500 spectrometer operating at a 27Al frequency
of 130.24 MHz and a 29Si frequency of 99.29 MHz. The
27
Al solid-state spectra were acquired using a 4-mm Doty
MAS probe with a silicon nitride rotor spun at 1012 kHz,
a 1 ls pulse and a 1 s recycle time, the spectra referenced
29
with respect to Al(H2O)3?
Si spectra were acquired
6 . The
with a 5-mm Doty MAS probe and a zirconia rotor spun at
*6 kHz. The excitation pulse for 29Si was 7 ls with a
recycle time of 30 s, and the spectra were referenced with
respect to tetramethyl silane (TMS).
The state of the copper in the samples before and after
their use as photocatalysts was studied by 63Cu NQR
spectroscopy. 63Cu (and 65Cu) quadrupole frequencies (mQ)

123

7376

J Mater Sci (2015) 50:73747383

were obtained at ambient temperature through the use of a


Chemagnetics CMX console pulsing into a probe
arrangement that was well removed from the magnet
([5 m) and shielded from extraneous magnetic and radio
frequency interference by a mumetal container. The
quadrupole frequency range scanned was determined from
previous 63,65Cu NQR and NMR studies of two, three and
four coordinate Cu (I) systems [1417]. The location of
both 63Cu and 65Cu isotope resonances (related by the ratio
mQ(63Cu)/mQ(65Cu)) = 1.081) verified that true copper NQR
frequencies were being observed. A scan of *500 kHz
around the central Cu2O frequency was undertaken for
each sample by stepping the spectrometer frequency in
100 kHz increments through this range and retuning the
probe at each frequency. A Hahn echo (hs2hs-acquire)
sequence with extended phase cycling was employed to
obtain undistorted echoes [18], where the recycle delay was
0.5 s and hard h/2h pulses of 2/4 ls duration, respectively,
were implemented.
Photodegradation measurements
The interaction of the composites with the chosen model
organic compound (methylene blue dye, MB) was carried
out by the batch method in which 100 mg of the powdered
sample was mixed with 100 ml of 0.1 mM aqueous MB
solution in a round-bottom flask and magnetically stirred
throughout the experiment. The experiments were carried
out both in the dark and under UV illumination using a
150-W xenon lamp as the light source. After varying
reaction periods, 5 ml aliquots of the sample were taken,
the suspension centrifuged off and the concentration of the
dye remaining in the solution determined at the absorbance
maximum (664 nm) using a UVvisible spectrophotometer
(Agilent 8453) operating in single-beam mode. The samples were measured in 1-cm wide quartz cuvettes, and the
spectra were recorded between 200 and 800 nm. The
removal efficiency g of the dye was then calculated by
Eq. 1:
g

C0  Ct
 100;
C0

where C0 and Ct are the concentrations of the dye solution


at the initial time and after time t.

Results and discussion


Characterization of the Cu2O nanoparticles
and Cu2O/geopolymer composites
SEM micrographs of the geopolymer matrix (Fig. 1d)
show a typically non-crystalline morphology, by contrast

123

with the as-synthesised Cu2O nanocrystals (Fig. 1a, b),


which occur as well defined but slightly truncated cubes
with sides of about 80160 nm. SEM of the Cu2
O/geopolymer composite containing 20 wt% Cu2O
(Fig. 1c) shows the Cu2O nanocrystals (the bright spots)
are well dispersed throughout the matrix, and the global
EDS spectrum of the composite (not shown) also confirms
the presence of the copper, together with the Si, Al, Na and
O components of the geopolymer matrix. TEM micrographs of the Cu2O/geopolymer composites (Fig. 1e) show
that the *100 nm Cu2O nanocubes have entered into the
pore structure of the geopolymer matrix even at Cu2O
concentrations as low as 10 wt%. The 0.22-nm lattice
spacings of the infiltrated crystallites (Fig. 1f) confirm their
identity as Cu2O.
A selection of representative powder XRD patterns of
the starting Cu2O and geopolymer composites with different Cu2O contents are shown in Fig. 2. The as-synthesised Cu2O (Fig. 2a) shows only the reflections of this
cubic phase (JCPDF file no. 0.1-078-2076), while the
narrowness of the peaks indicates a highly crystalline oxide
of small crystallite size. No reflections of CuO or Cu are
present. The XRD trace of the geopolymer matrix (Fig. 2b)
shows the broad background feature at about 2729 2h
typical of an X-ray amorphous phase, with superimposed
sharp reflections of crystalline quartz (JCPDF file no.
01-078-2076) and cristobalite (PDF no. 01-075-0923),
arising from impurities in the starting halloysite which
survive the geopolymerization reaction. The geopolymer/
Cu2O composites (Fig. 2ce) all show a combination of the
sharp crystalline Cu2O reflections superimposed on the
broad baseline feature of the geopolymer; the latter
becomes less apparent as the geopolymer matrix is
increasingly diluted by an increasing Cu2O content of the
composites.
The FTIR spectra of the matrix and composites (Fig. 3)
are all quite similar, containing peaks at about 1000 cm-1
arising from the SiO and AlO asymmetric stretching
vibrations [19]. The addition of Cu2O to the geopolymer
matrix exerts only a slight effect on its structure, as evidenced by a small shift from 1002 cm-1 in the reference
geopolymer (Fig. 3a) to 997 cm-1 in the composite containing 30 wt% Cu2O (Fig. 3d). The broad features in all
the samples at 34552851 cm-1 and at 16001700 cm-1
are related to the presence of water and hydroxyl groups
[19], while the only unique feature in the Cu2O (apart from
water vibrations) is a narrow CuO vibration at 629 cm-1
(Fig. 3e) [20] which is obscured by other vibrations in this
region in the composites; another diagnostic CuO vibration at 1384 cm-1 is, however, clearly visible in the
composite containing 30 wt% Cu2O (Fig. 3d).
Langmuir surface areas derived from the adsorption/
desorption isotherms (Fig. 4) show a decrease from

J Mater Sci (2015) 50:73747383

7377

Fig. 1 Representative SEM and TEM micrographs of a, b the as-synthesised cubic Cu2O nanocrystals, c 20 wt% Cu2O-geopolymer composite,
d the as-synthesised geopolymer matrix, e, f TEM micrographs of the 10 wt% Cu2O-geopolymer composite

21.9 m2/g in the geopolymer matrix to 6.8 m2/g in the


composite containing 10 wt% Cu2O, returning to 21.6 m2/g
in the sample containing 20 wt% Cu2O. Similar trends are
found in the BET surface areas and in the cumulative
surface areas of the pores calculated by the BJH method.
By contrast, the average pore diameters, calculated by the
BJH method, increase slightly from 41.0 nm in the
geopolymer matrix to 42.4 nm in the composite containing
10 wt% Cu2O, remaining approximately constant at
38.9 nm in the composite containing 20 wt% Cu2O. Thus,
the effect of incorporating the oxide nanoparticles into the
pore structure of the geopolymer matrix is to decrease the
surface area of the pores available for adsorption but
increasing the average pore size.
63
Cu NQR spectra of the Cu2O and Cu2O/geopolymer
composites (Fig. 5) show the expected single line in the
reported position for the pure oxide (Fig. 4a) [16] which is
unchanged by its incorporation in the geopolymer matrix
(Fig. 4c), indicating that the Cu2O is not incorporated into
the structure of the geopolymer matrix.
Representative 27Si MAS NMR spectra of the
geopolymer and geopolymer composites (Fig. 6) show that
these are all well-formed geopolymers with a broad

principal resonance at about -85 to -87 ppm, corresponding to a framework structure significantly saturated in
Al, but broadened by the presence of several slightly different SiOAl environments [21, 22]. A feature at about
-100 to -107 ppm which broadens into a shoulder in the
composites with higher Cu2O content (Fig. 5) denotes the
presence of sites more rich in silica [22]. The corresponding 27Al MAS NMR spectra (Fig. 7) all contain a
predominant sharp resonance at about 59 ppm corresponding to the tetrahedral AlO units of a well-developed
geopolymer structure [22]. The small broad feature in some
of the samples arises from AlO in octahedral coordination
[22], as a residual trace of unreacted metakaolin [22].
To summarize, these results confirm the successful
formation of aluminosilicate geopolymer composites containing homogeneously distributed nanosized cubic Cu2O
particles.
Reactions under dark conditions
Geopolymers are known to be efficient absorbers of
organic dyes in their own right, since their surface hydroxyl
groups can attract and hold cationic organic species [23].

123

7378

Fig. 2 XRD patterns of a the as-synthesised geopolymer matrix, b 10


wt% Cu2O-geopolymer composite, c 20 wt% Cu2O-geopolymer
composite, d 30 wt% Cu2O-geopolymer composite, e 30 wt% Cu2Ogeopolymer composite after MB adsorption in dark, f 30 wt% Cu2Ogeopolymer composite after MB degradation under UV. g assynthesised Cu2O. Key: Q quartz (PDF no. 01-075-8322), C cristobalite (PDF no. 01-075-0923), All unmarked peaks are Cu2O (PDF
no. 01-073-6237)

Fig. 3 Absorbance FTIR spectra of a the as-synthesised geopolymer


matrix, b 10 wt% Cu2O-geopolymer composite, c 20 wt% Cu2Ogeopolymer composite, d 30 wt% Cu2O-geopolymer composite, e 30
wt % Cu2O-geopolymer composite after MB adsorption in dark, f 30
wt% Cu2O-geopolymer composite after MB degradation under UV.
g as-synthesised Cu2O

Since geopolymer materials based on flyash have been


demonstrated to perform well in this application [23], the
present composites are expected to remove MB from

123

J Mater Sci (2015) 50:73747383

Fig. 4 Nitrogen adsorption/desorption isotherms for the geopolymer


matrix (full and open squares) and the 20 wt% Cu2O-geopolymer
composite (full and open triangles)

Fig. 5 Representative 63Cu NQR spectra of nanoparticle Cu2O and


Cu2O/geopolymer composites. a as-synthesised Cu2O, b Cu2O after
UV irradiation, c Cu2O/geopolymer composite before UV irradiation,
d Cu2O/geopolymer composite after UV irradiation

solution by a combination of adsorption and photodegradation processes. The UVvisible spectrum of MB contains
two absorption peaks, but all the present intensity measurements were made on the major adsorption at 664 nm.
Rapid initial adsorption of the dye under dark conditions
by the geopolymer matrix alone (Fig. 8) may be

J Mater Sci (2015) 50:73747383

Fig. 6 11.7 T 29Si MAS NMR spectra. a geopolymer matrix, b 10


wt% Cu2O-geopolymer composite, c 20 wt% Cu2O-geopolymer
composite, d 30 wt% Cu2O-geopolymer composite

understood in terms of the initial availability of the active


surface sites available for adsorption. The rate of MB
adsorption is seen to depend on the Cu2O content of the
composites, the most rapid adsorption occurring in the
sample containing 20 wt% Cu2O (Fig. 8). At longer times,
the accessible sites become saturated, resulting in the
slower absorption of the dye by diffusion into the porous
structure of adsorbent. Adsorption of the dye by the Cu2O
alone did not occur, but incorporation of the Cu2O into the
geopolymer matrix results in a greater effective surface
area and adsorption capacity of the Cu2O/geopolymer
composites (Fig. 8); this is probably due to the increased
size of the pore size of the matrix resulting from the entry
of the oxide nanoparticles, since the oxide itself is nonabsorbent. The most rapid adsorption of the MB dye occurs
in the geopolymer composite containing 20 wt% Cu2O
(approaching maximum adsorption after 90 min. and levelling out thereafter, Fig. 8); the decreased adsorption rate
of samples containing [20 wt% Cu2O is probably due to
blocking of the active adsorption sites by the additional
oxide.

7379

Fig. 7 11.7 T 27Al MAS NMR spectra. a geopolymer matrix, b 10


wt% Cu2O-geopolymer composite, c 20 wt% Cu2O-geopolymer
composite, d 30 wt% Cu2O-geopolymer composite

Fig. 8 Residual concentration (C/C0) of methylene blue upon


exposure to the geopolymer matrix (GP) and the Cu2O-geopolymer
composites versus time under dark conditions

123

7380

J Mater Sci (2015) 50:73747383

Dye adsorption kinetics under dark conditions


Two kinetic models, pseudo-first-order (Eq. 2) and pseudosecond-order (Eq. 3) have previously been used [24] to
analyse the kinetics of dye adsorption on to substrates:
q qe 1  expk1 t

k2 q2e t
;
q
1k2 qe t

2
3

where q is the amount of dye adsorbed (mg/g) at time


t (min), qe is the amount of dye adsorbed at equilibrium
(mg/g), k1 is the equilibrium rate constant for pseudo-firstorder kinetics (min-1) and k2 is the equilibrium rate constant for pseudo-second-order kinetics (g/mg/min). The
adsorption data (Fig. 8) were fitted to these two equations
by a nonlinear method with successive interactions calculated by the LevenbergMarquardt algorithm [25] using
OriginPro 2015 software. The resulting plots of the data
fitted to these two equations are shown in Fig. 9, with the
resulting pseudo-first- and pseudo-second-order rate constants, the qe values and the correlation coefficients R2
shown in Table 1. Comparison of the qe values calculated
from these equations with the experimental values qexp
(Table 1) shows that for all samples the pseudo-first-order
model provides a better fit to the experimental data. This fit
to pseudo-first-order kinetics is at variance with a result
reported for the simultaneous adsorption and photodecomposition of methylene blue dye on a flyash geopolymer, which followed pseudo-second-order kinetics [26].
This suggests that these processes are strongly influenced
by the nature of the geopolymer matrix which in the case of
flyash can contain photoactive constituents such as iron
oxides. The first-order rate constants (Table 1) increase
slightly with increasing Cu2O content of the geopolymer/

Cu2O composites, reflecting the improved adsorption by


the composites.
The adsorption mechanism of a cationic dye such as MB
can be related to its point of zero charge (PZC) with respect
to that of the adsorbent. No PZC values for geopolymers
are available in the literature, but those of zeolites, with
which geopolymers are closely related, are known to vary
widely with the Al/Si ratio. At higher pH values corresponding to the alkaline geopolymers, electrostatic
adsorption of cationic species is likely to occur on the
silanol groups present, and these are probably the adsorption sites in the present composites [27].
Dye adsorption isotherms under dark conditions
The relationship between the amount of solute adsorbed
and the concentration of the solution in the liquid phase at a
given constant temperature can be described by an
adsorption isotherm. Three equations commonly used to
describe adsorption isotherms are the Langmuir isotherm,
which assumes that the adsorption process occurs at
specific homogeneous sites within the adsorbent [24], the
Freundlich isotherm which is an empirical equation used to
describe heterogeneous systems and the LangmuirFreundlich isotherm which provides a more flexible analytical
modelling framework by use of a three-parameter equation
that can give a better fit to the data.
The Langmuir isotherm takes the following form:
q

Qs kC
;
1kC

where q is the adsorbed amount of the dye (mg/g), C is the


equilibrium concentration of the dye in solution (mg/l), Qs
is the adsorption capacity (mg/g) and k is a constant related
to the free energy of adsorption (l/mg).

Fig. 9 a Experimental data for the adsorption of methylene blue under dark conditions on to the geopolymer matrix and the Cu2O/geopolymer
composites fitted to a pseudo-first-order kinetic model, b pseudo-second-order kinetic model

123

J Mater Sci (2015) 50:73747383


Table 1 Kinetic parameters for
adsorption of methylene blue
under dark conditions

7381

Sample
GP

Kinetic model

qexp (mg/g)

qe (mg/g)

K1 (min-1)

K2 (g/mg/min)

1st-order

14.74

15.33

0.015

0.90

19.85

7.05e-04

0.89

13.59

0.026

0.75

15.69

0.002

0.75

13.91

13.10

0.021

0.92

16.14

0.001

0.92

14.76

15.51

0.028

0.73

17.73

0.001

0.73

2nd-order
GP/10 %Cu2O

1st-order

14.05

2nd-order
GP/20 %Cu2O

1st-order
2nd-order

GP/30 %Cu2O

1st-order
2nd-order

The Freundlich isotherm can be written as follows:


n

q kC ;

where k and n are constants specifying the extent of the


adsorption and the degree of nonlinearity between the
solution concentration and the adsorption.
The LangmuirFreundlich equation takes the following
form:

Qs kC n
:
1 kC n

R2

When the present dye adsorption data were fitted to


these three isotherm equations, the best fit, evidenced by
the R2 values, was provided by the Freundlich and LangmuirFreundlich equations (R2 [ 0.90 and [0.96, respectively). Figure 10 shows the experimental methylene blue

Fig. 10 Adsorption isotherms for the experimental data of methylene blue under dark conditions on to the geopolymer matrix and the
Cu2O/geopolymer composites fitted to a Freundlich, b Langmuir, c LangmuirFreundlich models

123

7382

adsorption data for the geopolymer matrix and all the


geopolymer/Cu2O composites fitted to the Freundlich,
Langmuir and LangmuirFreundlich models (Fig. 10ac,
respectively). These plots indicate that the more empirical
Langmuir-type isotherms are the most appropriate to
describe the absorption of the dye on the geopolymer and
geopolymers composites.
Reactions under UV irradiation
The two processes operating in the decoloration of the MB
dye under UV irradiation can be seen in Fig. 11, which
illustrates the difference between the behaviour of the
geopolymer matrix and the composite containing 30 wt%
Cu2O in the dark and under UV illumination. The
geopolymer matrix behaves essentially similarly in dark
and UV conditions, showing the adsorption of the dye but
no evidence of photoactive properties, as would be
expected. By contrast, the Cu2O-containing composite is
more efficient in removing the dye, due to a combination of
adsorption and photodegradation processes.
After adsorption or photodegradation of the methylene
blue, the structures of the geopolymer matrix and the
Cu2O/geopolymer composites are essentially unchanged,
according to their XRD traces (Fig. 2e, f) and their FTIR
spectra (Fig. 3e, f). The photoactive Cu2O itself is unaffected by exposure to methylene blue, either in the dark or
under UV irradiation, as evidenced by 63Cu NQR spectroscopy (Fig. 4b, d). Thus, these composites should continue to remove organic substances from their immediate
environment by a combination of adsorption and

J Mater Sci (2015) 50:73747383

photodegradation without deterioration of their structure or


of the photoactive component.

Conclusions
Novel composites of cubic Cu2O nanoparticles with aluminosilicate geopolymers were prepared and the homogeneity of the oxide nanoparticle dispersion within the
geopolymer matrix was demonstrated by SEM/EDS and
TEM. XRD, FTIR, 27Al and 29Si MAS NMR and 63Cu
NQR spectroscopy confirmed the formation of a well-reacted geopolymer matrix that was unaffected by the
insertion of the Cu2O nanoparticles or by the adsorption or
photodegradation of a model organic compound (methylene blue, chosen for its colour stability at high pH). Under
dark conditions, the geopolymer matrix, and in particular,
the Cu2O/geopolymer composites, remove the dye by an
adsorption process which follows pseudo-first-order kinetics and can be described by Langmuir-type isotherms.
Although the nano-Cu2O itself does not adsorb the
methylene blue dye, its incorporation up to 20 wt% into the
geopolymer composites increases their adsorption ability,
probably by increasing the pore size of the matrix. At
concentrations [20 wt%, the oxide decreases the adsorption by blocking the active adsorption sites of the
geopolymer. Under UV radiation, the composites remove
the methylene blue by a combination of adsorption and
photodegradation, without deterioration of the geopolymer
structure or the photoactive Cu2O component, suggesting
that these geopolymer composites should function as useful
new materials for the removal of organic pollutants from
water or the atmosphere.
Acknowledgements We are indebted to Grant OSullivan for supplying the halloysite clay, to David Flynn for assistance with the
electron microscopy, to Diego del Puerto for the BET measurements
and to Dr. Ruth Knibbe for assistance in interpreting the TEM
micrographs. Mahroo Fallah acknowledges the financial support of a
PhD Fellowship from the MacDiarmid Institute for Advanced Materials and Nanotechnology. JVH thanks the University of Warwick,
EPSRC and the Birmingham Science City for contribution to the
solid-state NQR instrumentation used in this research. The latter was
partially supported through the Birmingham Science City Advanced
Materials Project 1: Creating and Characterising Next Generation
Materials Project, with support from Advantage West Midlands
(AWM) and partial funding from the European Regional Development Fund (ERDF).

References
Fig. 11 Residual concentration (C/C0) of methylene blue upon
exposure to the geopolymer matrix (GP) and the 30 wt% Cu2Ogeopolymer composite versus time in the dark and under UV
illumination

123

1. Serpone N, Pelizzetti E (eds) (1989) Photocatalysis: fundamentals and applications. Wiley, New York
2. Kibanova D, Trejo M, Destaillats H, Cervini-Silva J (2009)
Synthesis of hectoriteTiO2 and kaoliniteTiO2 nanocomposites

J Mater Sci (2015) 50:73747383

3.

4.

5.

6.

7.

8.

9.

10.
11.

12.

13.

with photocatalytic activity for the degradation of model air


pollutants. Appl Clay Sci 42:563568
Dvininov E, Popovici E, Pode R, Cocheci L, Barvinschi P, Nica
V (2009) Synthesis and characterization of TiO2-pillared Romanian clay and their application for azoic dyes photodegradation.
J Hazard Mater 167:10501056
Zhang Q, Shan A, Wang D, Jian L, Cheng L, Ma H, Li J (2013) A
new acidic Ti sol impregnated kaolin photocatalyst: synthesis,
characterization and visible light photocatalytic performance.
J Sol Gel Technol 65:204211
Mamulova Kutlakova K, Tokarsky J, Kovar P, Vojteskova S,
Kovarova A, Smetana B, Kukutschova J, Capkova P, Matejka V
(2011) Preparation and characterization of photoactive composite
kaolinite/TiO2. J Hazard Mater 188:212220
Chong MN, Vimonses V, Lei S, Jin B, Chow C, Saint C (2009)
Synthesis and characterisation of novel titania impregnated
kaolinite nano-photocatalyst. Micropor Mesopor Mater
117:233242
Okada K, Yoshizawa A, Kameshima Y, Isobe T, Nakajima A,
MacKenzie KJD (2011) Adsorption and photocatalytic properties
of TiO2/mesoporous silica composites from two silica sources
(acid-leached kaolinite and Si-alkoxide). J Porous Mater
18:345354. doi:10.1007/s10934-010-9384-2
Gasca-Tirado JR, Manzano-Ramrez A, Villasenor-Mora C,
valos JC,
Muniz-Villarreal MS, Zaldivar-Cadena AA, Rubio-A
Borras VA, Mendoza RN (2012) Incorporation of photoactive
TiO2 in an aluminosilicate inorganic polymer by ion exchange.
Micropor Mesopor Mater 153:282287
Gasca-Tirado JR, Manzano-Ramrez A, Vasquez-Landaverde
valos JC,
PA, Herrera-Diaz EI, Rodriguez-Ugarte ME, Rubio-A
Amigo-Borras V, Chavez-Paez M (2014) Ion exchanged
geopolymer for photocatalytic degradation of a volatile organic
compound. Mater Lett 134:222224
Pollack G, Trivich D (1975) Photoelectric properties of cuprous
oxide. J Appl Phys 46:163172
Yang H, Ouyang J, Tang A, Xiao Y, Li X, Dong X, Yu Y (2006)
Electrochemical synthesis and photocatalytic property of cuprous
oxide nanoparticles. Mater Res Bull 41:13101318
Huang L, Peng F, Yu H, Wang H (2009) Preparation of cuprous
oxides with different sizes and their behaviors of adsorption,
visible-light driven photocatalysis and photocorrosion. Solid
State Sci 11:129138
Bai Y, Yang T, Gu Q, Cheng G, Zheng R (2012) Shape control
mechanism of cuprous oxide nanoparticles in aqueous colloidal
solutions. Powder Technol 227:3542

7383
14. Kroeker S, Wasylishen RE, Hanna JV, Wasylishen RE, Ainscough EW, Brodie AM (1998) Anisotropy in the 31P, 63/65Cu
indirect spin-spin coupling and 31P nuclear shielding tensors of
copper (I) phosphines. J Magn Reson 135:208215
15. Kroeker S, Wasylishen RE, Hanna JV (1999) The structure of
solid copper(I) cyanide: a multinuclear magnetic and quadrupole
resonance study. J Am Chem Soc 121:15821590
16. Vega AJ (1992) 63Cu nutation spectroscopy and spin counting in
copper oxides. Isr J Chem 32:195204
17. Bowmaker GA, Boyd SE, Hanna JV, Hart RD, Healy PC, Skelton
BW, White AH (2002) Structural and spectroscopic studies on
three-coordinate complexes of copper (I) halides with tricyclohexylphosphine. J Chem Soc Dalton Trans 13:27222730
18. Kunwar AC, Turner GL, Oldfield E (1986) Solid-state spin echo
Fourier transform NMR of 39K and 67Zn salts at high field.
J Magn Reson 69:124127
19. Barbosa VFF, MacKenzie KJD, Thaumaturgo C (2000) Synthesis
and characterisation of materials based on inorganic polymers of
alumina and silica: sodium polysialate polymers. Int J Inorg
Mater 2:309317
20. Papadimitropoulos G, Vourdas N, Em Vamvakas V, Davazoglou
D (2005) Deposition and characterization of copper oxide thin
films. J Phys Conf Ser 10:182185
21. Rowles MR, Hanna JV, Pike KJ, Smith ME, OConnor BH
(2007) 29Si, 27Al, 1H and 23Na MAS NMR study of the bonding
character in aluminosilicate inorganic polymers. Appl Magn
Reson 32:663689
22. MacKenzie KJD, Smith ME (2002) Multinuclear solid state NMR
of inorganic materials. Pergamon materials series, vol 6. Pergamon/Elsevier, Oxford
23. Li L, Wang S, Zhu Z (2006) Geopolymeric adsorbents from fly
ash for dye removal from aqueous solution. J Colloid Interface
Sci 300:5259
24. Cardoso NF, Lima EC, Pinto IS, Amavisca CV, Royer B, Pinto
RB, Alencar WS, Pereira SFP (2011) Application of cupuassu
shell as biosorbent for the removal of textile dyes from aqueous
solution. J Environ Manag 92:12371247
25. Pujol J (2007) The solution of nonlinear inverse problems and the
Levenberg-Marquardt method. Geophysics 72:W1W16
26. Zhang Y, Liu L (2013) Fly ash-based geopolymer as a novel
photocatalyst for degradation of dye from wastewater. Particuology 11:353358
27. Schreier M, Teren S, Belcher L, Regalbuto JR, Miller JT (2005)
The nature of overexchanged copper and platinum on zeolites.
Nanotech 16:S582S591

123

Das könnte Ihnen auch gefallen