Sie sind auf Seite 1von 9

Ultrasonics Sonochemistry 21 (2014) 11081116

Contents lists available at ScienceDirect

Ultrasonics Sonochemistry
journal homepage: www.elsevier.com/locate/ultson

Ultrasound-based treatment approaches for intrinsic viscosity reduction


of polyvinyl pyrrolidone (PVP)
Indrajeet A. Pawar a, Prathmesh J. Joshi a, Akshay D. Kadam a, Nishant B. Pande a, Priyanka H. Kamble a,
Shruti P. Hinge a, Barnali S. Banerjee a, Ashish V. Mohod a,, Parag R. Gogate b,
a
b

Department of Chemical Engineering, AISSMS College of Engineering, Kennedy Road, Pune 411001, India
Chemical Engineering Department, Institute of Chemical Technology, Matunga, Mumbai 400019, India

a r t i c l e

i n f o

Article history:
Received 17 July 2013
Received in revised form 2 December 2013
Accepted 14 December 2013
Available online 21 December 2013
Keywords:
Polymer degradation
PVP
Intrinsic viscosity
Ultrasound
Ultraviolet irradiation
Titanium dioxide

a b s t r a c t
The present work deals with achieving viscosity reduction in polymer solutions using ultrasound-based
treatment approaches. Use of simple additives such as salts, or surfactants and introduction of air at varying ow rates as process intensifying parameters have been investigated for enhancing the degradation of
polyvinyl pyrrolidone (PVP) using ultrasonic irradiation. Sonication is carried out using an ultrasonic horn
at 36 kHz frequency at an optimized concentration (1%) of the polymer. The degradation behavior has
been characterized in terms of the change in the viscosity of the aqueous solution of PVP. The intrinsic
viscosity of the polymer has been shown to decrease to a limiting value, which is dependent on the operating conditions and use of different additives. Similar extent of viscosity reduction has been observed
with 1% NaCl or 0.1% TiO2 at optimized depth of horn and 27 C, indicating the superiority of titanium
dioxide as an additive. The combination of ultrasound and ultraviolet (UV) irradiation results in a
signicantly faster viscosity reduction as compared to the individual operations. A kinetic analysis for
the degradation of PVP has also been carried out. The work provides a detailed understanding of the role
of the operating parameters and additives in deciding the extent of reduction in the intrinsic viscosity of
PVP solutions.
2013 Elsevier B.V. All rights reserved.

1. Introduction
Worldwide production of synthetic polymers is approximately
140 million pounds annually. Subjecting polymers to degradation
(mostly for achieving a reduction in viscosity) is a signicantly
important processing step due to a wide variety of applications
of polymers also requiring specic characteristics. Polymer
degradation techniques based on the use of energy (thermal or
radiation), chemical (acid or alkali) and microorganisms or enzymes are particularly important and applied generally for a range
of polymers. Pyrolysis, photolysis, biological action and shear can
be effectively used for the degradation of different polymers and
these processes have been widely investigated for polymers such
as poly (ethylene oxide), polyethylene, polypropylene, etc. Prolonged exposure of solutions of macromolecules to high energy
waves (ultrasound) has been shown to produce a permanent
reduction in viscosity [18]. Schmidt and Rommel [2] rst
observed the permanent reduction in the viscosity of polymer
Corresponding authors. Tel.: +91 20 26058587; fax: +91 20 26059843 (A.V.
Mohod). Tel.: +91 22 33612024; fax: +91 22 33611020 (P.R. Gogate).
E-mail addresses: ashishmohod2004@gmail.com (A.V. Mohod), pr.gogate@
ictmumbai.edu.in (P.R. Gogate).
1350-4177/$ - see front matter 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.ultsonch.2013.12.013

solutions and attributed it to the breakage of covalent bonds in


the polymer chain.
In recent years, it has been observed that degradation of
polymers is of primary importance due to widespread uses of the
polymer in many industrial applications. Synthetic polymers such
as poly(glycolic acid), poly(lactic acid) and their copolymers,
poly(p-dioxanone), and copolymers of trimethylene carbonate
and glycolide have been used in a number of clinical applications
[912]. The major applications include resorbable sutures, drug
delivery systems and orthopaedic xation devices such as pins,
rods, and screws [13,14].
It has also been observed that many opportunities exist for the
application of synthetic polymers in the biomedical area, particularly, in the eld of tissue engineering and controlled drug delivery.
In biomedical applications, the criteria for selection of biomaterials
are based on their functional groups, molecular weight, solubility,
shape and structure, hydrophilicity/hydrophobicity, lubricity, surface energy, water absorption, and erosion mechanism. In tissue
engineering, polymeric scaffolds play a unique role in tissue regeneration and repair. Scaffolds can be fabricated from either synthetic
or biological polymers, and may be degradable or non degradable,
depending upon the intended use [15]. Biological scaffolds are derived from humans and animal tissues whereas synthetic scaffolds

I.A. Pawar et al. / Ultrasonics Sonochemistry 21 (2014) 11081116

are derived from synthetic polymers. The degradation of scaffolds


can occur through mechanisms that involve physical, chemical or
biological processes.
Reduction in viscosity would also be required in mechanical
synthesis of various block and graft copolymers, in the plasticization of rubber, and in the production of valuable low-molecular
weight substances e.g. glucose from natural polymers. Thus, the
study of degradation is essential in many ways and can be one of
the practical methods for stabilizing the polymers [16].
Use of ultrasonic irradiation, or in general, cavitation phenomenon has also been looked upon as a promising technology not only
for physical processing, but also for chemical processing applications [17,18]. Ultrasound can also be looked upon as an effective
means to introduce the desired changes in polymer solutions.
Ultrasonic irradiation has a signicant effect on polymers in terms
of their mechanical, mechano-chemical and morphological properties. Understanding the effect of different operating parameters on
the extent of degradation of polymers, thus, becomes very important. The polymer degradation can be quantied either by molecular weight distribution or by change in the intrinsic viscosity [19].
It is believed, that, ultrasonic degradation, unlike chemical or thermal decomposition, is a non-random process, with cleavage taking
place at the center of the polymer chain. It is important to note
here that there is a certain molecular weight of polymer below
which degradation does not take place. The polymer degradation
in the solution using ultrasonication is due to the motion of the
wall of violently collapsing bubble causing movements of solvent
molecules around these bubbles. These movements set up large
shear elds due to which the polymer chain get shorter or contracted [19,2026]. In general, as the concentration of polymer decreases, the viscosity of the solution lowers and thus enhancement
in degradation can be accomplished by reducing the corresponding
viscosity. When high intensity acoustic energy waves travel
through a medium, rapid and successive compression and rarefaction cycles occur. The decrease in the pressure in the negative pressure cycle of the ultrasonic shock wave spontaneously generates
small cavities. Theses cavities collapse in the positive pressure cycle and produce highly turbulent ow conditions and extremely
high pressure and temperatures, a process that is better known
as cavitation. It is known that the cavitation in the liquid requires
that the negative pressure in the rarefaction region of wave function must overcome the natural cohesion forces acting within the
liquid. It is difcult to use cavitation in highly viscous liquids,
due to strong forces, as it requires waves with greater amplitude
and intensity that limits the collision activity of bubbles. Due to
above fact, dilute solutions of polymers have been generally used
for the study of ultrasonic degradation [27].
A variety of different models have been proposed to explain the
way in which the factors such as frequency, intensity, solvent, temperature, nature of dissolved gas, external pressure and molecular
mass distribution inuence the rate of degradation and nal
molecular mass of the degraded species [2830]. Akyuz et al.
[30] investigated the polymer degradation using ultrasound and
compared different theoretical and phenomenological models of
evolution of the molecular weight during sonication. Typically,
these models share the property of an initial rapid drop in molecular weight followed by a slowing down of the rate of decrease in
the molecular weight. Different models are proposed to explain the
degradation phenomena based on cavitation induced by
ultrasound.
Polyvinyl pyrrolidone (PVP), which is also called as Polyvidone
or Povidone, is a water-soluble polymer. PVP is used as a binder
in many pharmaceutical tablets and also as an adhesive in glue
sticks and hot-melts. It is also used as a special additive for batteries, ceramics, berglass, inks and also acts as an emulsier and disintegrant for solution polymerization. PVP is also used in personal

1109

care products, such as shampoos and toothpastes. In molecular


biology, PVP can be used as a blocking agent during Southern blot
analysis as a component of Denhardts buffer. It is also exceptionally good at absorbing polyphenols during DNA purication. The
application of PVP depends on the different properties such as
molecular weight, density, viscosity, solubility etc and hence it is
important to study the polymer degradation in terms of achieving
the viscosity reduction as well as molecular weight and this step
can be considered as an essential preprocessing requirement for
applications of PVP in any eld. Due to large number of applications, PVP is considered as a very important polymer and hence
was considered as the model polymer in the present work. There
are not many studies in the literature on detailed investigation of
sonochemical degradation of PVP. The work of Akyuz et al. [30]
dealt with a fundamental understanding into the degradation process and reported the comparison of different kinetic models for
degradation of PVP. Taghizadeh et al. [31] investigated the effect
of initial molecular weight of PVP on the extent of degradation
using ultrasonic irradiations where the extent of degradation was
quantied using the method of viscometry and a kinetic model
was developed to estimate the degradation rate constant. It has
been reported that the rate constant decreases as the initial concentration increases. The objective of the present study is to establish new experimental data for the ultrasonic degradation of
polyvinyl pyrrolidone (PVP) in terms of the optimization of the different operating parameters such as depth of horn and operating
temperature at laboratory scale operation using ultrasonic horn
reactor. After the completion of initial laboratory scale optimization studies, intensication studies using different additives such
as air, sodium chloride, titanium dioxide and surfactant have been
carried out. The work is also rst of its kind to report the inuence
of additives and combination of US and ultraviolet irradiations for
the degradation of PVP. The extent of degradation has been quantied in terms of the change in the intrinsic viscosity of the polymer solution, which is a simple method for monitoring the rate
of degradation and used generally in monitoring the rate of polymer degradation [3134].
2. Experimental work
2.1. Materials
The material used was polyvinyl Pyrrolidone (PVP), obtained
from S. D. Fine Chemicals, Mumbai, India. The molecular weight
of polyvinyl pyrrolidone is approximately 125,000 g/mol with values of the constants, k (related to the particular solventsolute
pair) as 1.30  102 l/kg and a (related to the shape of the solute
molecule) as 0.68, for 86.8% hydrolysis [35]. k and a are the
Mark-Houwink parameters that relate intrinsic viscosity (g) with
the molecular weight (M), specically for polymer solution and
whose values are dependent on the particular polymer-solvent
system. The dilute solutions of different chemicals with required
concentrations were obtained using demineralized water (DM
water) for experimental studies. Titanium oxide (TiO2), sodium
sulphate (Na2SO4), sodium chloride (NaCl), sodium carbonate
(Na2CO3), sodium lauryl sulphate (SLS) were obtained from Merck
Specialties Pvt. Ltd., Mumbai, India.
2.2. Sonochemical reactors
For the present experimental studies, ultrasonic horn (Fig. 1)
obtained from M/s Dakshin, Mumbai operating at 36 kHz and a
rated power dissipation of 120 W has been used. The tip of the
horn is 2 cm in diameter. The actual power dissipated in the
system was investigated using calorimetric measurements [36],

1110

I.A. Pawar et al. / Ultrasonics Sonochemistry 21 (2014) 11081116

The intrinsic viscosity [g] values can be related to the specic


viscosity and relative viscosity by the Huggins and Kramer equations [37]. The conditions used in this work (a = 0.55 and
K = 6.67  105 l g/l) were adopted on the basis of previous ndings reported in the literature [38].
Reproducibility of the efux time measurements was within
0.3 s. Experiments have been also repeated twice to check the
reproducibility of the obtained data for the variation of viscosity
against time for all the sets. It has been observed that experimental
errors were within 2%. The g values for the PVP solution at different parameters were calculated by the one-point intrinsic viscosity
equation [39].

g 2gsp  ln gr 0:5 =C

Fig. 1. Schematic representation of ultrasonic horn.

and was observed to be equal to 29.3 W giving an energy efciency


of 24.5%. Calorimetric measurements for each run also indicated
that the ultrasonic power dissipated in the solution is same for
each of the viscosities and is not affected due to the use of different
additives over the range of loadings considered in the work.
The solutions of PVP having concentrations were prepared by
using demineralized water and after initial optimization studies
with varying concentration, initial concentration of 1% (w/v) was
kept constant for all the remaining experimental runs. The degradation was investigated at different operating temperatures and
in the presence of different additives. During all the experimental
runs, any external bubbling was not introduced expect for the runs
involving presence of air. Whenever any solvent or solution is irradiated by ultrasonic energy, heat is generated and temperature of
the system increases. In the present work, the desired temperature
of the reaction vessel (glass beaker) was maintained by using an ice
bath around the glass beaker. For experiments involving the ultraviolet irradiations, a UV lamp was used which was procured from
Philips India Pvt. Ltd., Mumbai (Model PL-S). Three different lamps
with a power rating of 5 W, 9 W and 11 W respectively, have been
used with an objective of investigating the dependency on the
power dissipation of UV tube.

where C is concentration of PVP solution in g/l.


The variation of either molecular weight or the intrinsic viscosity in the presence of ultrasonic irradiation reects the ultrasonic
degradation of polymer. The extent of ultrasonic degradation of
polymer solution has been quantied by using a parameter:

In a typical run for optimization studies, the polymer solution


was taken in a glass beaker with a maximum capacity of 500 ml.
Periodically samples of sonicated solutions were removed for viscosity measurements which was carried out using a Ubbelohde viscometer (Model 9721-K56) procured from Cole-Parmer Pvt. Ltd.,
Mumbai, India. Before the withdrawal of the samples, it was ensured that the reaction mixture was well mixed. The location of
the sampling point was near the bottom of the reactor, i.e. away
from the location of the ultrasonic horn and the same location
was used in all experiments for the removal of the sample. Due
to distribution of cavitational activity in the reactor (maximum
near to the horn surface and reduced activity away from the surface of the ultrasonic horn), maximum extent of degradation is obtained very near to the ultrasonic horn and hence sampling point at
this location should be avoided.
Relative and specic viscosities (gr and gsp, respectively) were
calculated using Eq. (2):

g0   g
 100
g0   g1 

g k  M a

where [g] is the intrinsic viscosity, k and a are constants, values of


which depend on the type of polymer, solvent and the temperature
of viscosity determinations [40].
2.4. Kinetic analysis of polymer degradation process
Baramboim [41] suggested that the kinetics of polymer degradation under stress could be expressed as Eq. (6), which is applied
to describe the kinetics of ultrasonic degradation in this work:
dM t M 1
M1



Mt  M1
M1

where M1 and Mt are the limiting molecular weight and average


molecular weight at irradiation time t, respectively, and k is the rate
constant of degradation reaction.
By integrating and considering that at t = 0, Mt = M0 (where M0
is the initial molecular weight), Eq. (6) could be expressed as:

Mt M0  M 1 ekt M 1

If the polymer degradation process is monitored in terms of the


intrinsic viscosity of the polymer solution, similar equation in
terms of intrinsic viscosity can be written as:

gt g0  g1 ekt g1

According to Eq. (8), the magnitude of the rate constant has


been calculated for all the runs knowing the initial viscosity and
the limiting intrinsic viscosity of the polymer solution by plotting
a graph of ln (A) against time [42] where:

A g0  g1 =gt  g1

gr t=t0

3. Results and discussion

gsp gr  1

3.1. Effect of reaction volume

where t and t0 are the efux time for polymer solution and solvent,
respectively.

where [g] and [g1] is actual intrinsic viscosity and limiting intrinsic
viscosity, respectively. g0 is the initial intrinsic viscosity.
For calculating the molecular weight of PVP solution, MarkHouwink equation has been used which describes the dependence
of the intrinsic viscosity of a polymer on its relative molecular
mass (molecular weight) and can be given as follows:

dt

2.3. Experimental methodology

As the reaction volume plays a key role in determining the extent of degradation, the experiments were carried out at different

I.A. Pawar et al. / Ultrasonics Sonochemistry 21 (2014) 11081116

reaction volumes with same power dissipation of the ultrasonic


reactor at 1% concentration of PVP. Fig. 2 shows the effect of reaction volume on the extent of degradation in terms of the change in
intrinsic viscosity at a 1% concentration of PVP solution. The extent
of degradation decreases with an increase in the reaction volume
at same supplied ultrasonic power dissipation. After 180 min of
irradiation time, the extent of degradation for 200 ml at 1% concentration of PVP solution is two-times higher than that obtained at
300, 400 and 500 ml. As it can be perceived that increase in the
reaction volume decreases the power dissipation per unit volume,
resulting in a corresponding decrease in the cavitational activity.
The decreased cavitational activity naturally reduces the extent
of degradation of polymer. Sivakumar and Pandit [43] have reported similar trends for the effect of power density on the extent
of degradation of Rhodamine B whereas Harkal et al. [44] have reported similar results for the degradation of polyvinyl alcohol. It
has been also observed that the degradation rate constant for optimum volume, i.e. 200 ml, was 2.6  102 (min1).

3.2. Effect of initial concentration of polymer


The effect of initial concentration of PVP on extent of degradation has been studied at different initial concentrations including
1%, 1.5%, and 2% and the obtained results have been given in
Fig. 3. It has been inferred that the extent of degradation was higher at lower concentration of PVP and hence the lower initial concentration of 1% has been selected for the further work on
parameter optimization and investigating the effect of additives.
Similar results in terms of lower extent of degradation at higher
concentrations have been reported earlier [45,46]. The degradation
rate constant was also found to be higher (about 3 times) for 1%
concentration as compared to the higher concentration of 2.0%.
The observed results can be attributed to the fact that concentration of polymer solution decides the viscosity of polymer solution. With an increase in concentration, viscosity also increases
and at high viscosity of solution, the cavitational intensity is lower.
Independent bubble dynamics studies have clearly indicated that
an increase in the viscosity results in decrease in the collapse pressure generated due to cavitation [47]. In highly viscous solution,
the molecules present in the solution are less mobile. Therefore
velocity gradients around the collapsing bubble become smaller,
which results in lower extents of degradation. Similar results for
the effect of initial concentration have been reported by Taghizadeh et al. [31] considering PVP of different initial molecular
weights (160,000, 360,000 and 1,300,000 g/mol) at xed temperature. It was established that the degradation rate reduced with

Fig. 2. Effect of reaction volume on degradation of PVP using ultrasound (sonication


time 180 min; 1% PVP; depth of horn 0.5 cm; 300 K).

1111

Fig. 3. Effect of initial concentration of PVP on degradation using ultrasound


(sonication time 180 min; volume 200 ml; depth of horn 0.5 cm; 300 K).

increasing solution concentration or with an increase in the molecular weight of polymer.


3.3. Effect of sonication time on intrinsic viscosity (only sonication)
Fig. 4 shows the change in intrinsic viscosity of PVP solution
with sonication time at 1% initial concentration of PVP for 200 ml
reaction volume, depth of horn as 0.5 cm, and operating temperature as 27 C. Using only ultrasound, it has been observed that the
viscosity of the polymer solution decreases with an increase in the
ultrasonic irradiation time. The study also shows that the
degradation of molecules continued only to a certain limiting
molecular weight or intrinsic viscosity. Below this intrinsic viscosity value, the polymer chain was too short and thus, cleavage at the
center of the PVP chain did not take place any further. From Fig. 4,
it has been observed that the limiting intrinsic viscosity of PVP
solution was 8.50 (l/g), which was achieved after 200 min of
treatment.
A number of different rate models have been proposed for the
degradation of polymers [2,8], but in this study a simple model described by Eq. (6) was employed. The obtained data was observed
to be consistent with Eq. (8) by considering the order of reaction
with respect to molar concentration as 1. The plot of ln (A) versus
sonication time has been given in Fig. 5. The apparent degradation
rate constant can be estimated from the slope of the plot using Eq.
(9). Table 1 gives the value of the rate constant obtained from the
kinetic studies and it can be seen that the observed initial kinetic

Fig. 4. Effect of sonication time on intrinsic viscosity on degradation of PVP using


ultrasound (sonication time 180 min; 1% PVP; volume 200 ml; depth of horn
0.5 cm; 300 K).

1112

I.A. Pawar et al. / Ultrasonics Sonochemistry 21 (2014) 11081116

rate constant for PVP solution degradation using ultrasound alone


was 1.8  102 min1.

limiting intrinsic viscosity is marginally affected but the time for


reaching the limiting viscosity decreases with an increase in the
depth of the horn. At a depth of 0.5 cm, the required time for reaching the limiting viscosity was around 175 min but the time reduced
to only 120 min when the depth of horn in the reactor was increased to 2 cm. The results can be attributed to the changes in
the ow pattern of the liquid depending on the distance of horn
tip immersed in the polymer solution. The direct circulation
currents in the liquid solution are due to the acoustic ow in the
reactor when horn is mostly immersed in the liquid solution
whereas the reverse ow i.e. reection occurred from the bottom
of the reactor also contribute signicantly at enhanced depth of
horn. The extent of mixing in the reactor is also dependent on
the immersion depth of the horn tip [48]. The results reported in
this study are consistent with the results reported for ultrasonic
precipitation of calcium carbonate, which is also controlled by
the physical effects of cavitation phenomena [48]. As it is observed
that 0.5 cm immersion depth of horn gives similar results in terms
of reaching the nal limiting viscosity, this preferred value of
0.5 cm depth of horn based on the efcient operation of current
conguration of ultrasonic horn has been kept constant for all
the remaining experiments.

3.4. Effect of depth of horn

3.5. Effect of air sparging at different ow rate

The effect of depth of horn on the extent of degradation has


been investigated at a constant power dissipation of the ultrasonic
horn. The experiments were conducted for varying depth of horn
over the range of 0.52 cm for a constant concentration of 1%
and volume as 200 ml. Fig. 6 shows the effect of depth of horn
on the intrinsic viscosity at different depths of horn. It can be easily
seen from the gure that the rate of decrease in the intrinsic viscosity increases with an increase in the depth of horn. The nal

Experiments were performed to study the introduction of air on


viscosity reduction of PVP solution. The air was introduced with
the help of sh pond aerator at different ow-rates. The obtained
results for effect of varying air ow rate as 15, 25 and 35 cm3/s under the conditions of normal temperature and pressure i.e. 1 atm
and 25 C on the extent of degradation of the PVP has been
depicted in Fig. 7. It can be seen from the gure that the rate of viscosity reduction is maximum at ow rate of air as 15 cm3/s. It can

Fig. 5. Kinetic rate constant for degradation of PVP using ultrasound (sonication
time 180 min; 1% PVP; volume 200 ml; depth of horn 0.5 cm;300 K).

Table 1
Kinetic rate constant for different sets of PVP degradation.
Sr. no.

Parameters

Extent of degradation (%) after 180 min

Kinetic rate constant (min1)

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33

Reaction volume 200 ml


Reaction volume 300 ml
Reaction volume 400 ml
Reaction volume 500 ml
Concen. of PVP at 1%
Concen. of PVP at 1.5%
Concen. of PVP at 2%
Air ow rate at15 cm/s2
Air ow rate at 25 cm/s2
Air ow rate at 35 cm/s2
NaCl at 1% concen.
NaCl at 2% concen.
NaCl at 4% concen.
NaCl at 6% concen.
NaSO4 at 1% concen.
NaSO4 at 2% concen.
NaSO4 at 4% concen.
NaSO4 at 6% concen.
Na2CO3 at 1% concen.
Na2CO3 at 2% concen.
Na2CO3 at 4% concen.
Na2CO3 at 6% concen.
0.5 cm depth of horn
1.0 cm depth of horn
2.0 cm depth of horn
SLS at 0.1%
SLS at 1.0%
TiO2 at 0.1%
TiO2 at 1.0%
UV only (5 W)
US + UV (5 W)
US + UV (9 W)
US + UV (11 W)

83.89
68.62
77.25
64.40
83.89
72.90
50.05
88.40
92.98
97.64
83.89
83.31
80.43
70.38
82.23
81.34
76.34
74.34
80.34
78.36
75.45
72.34
83.89
83.89
83.89
79.99
64.30
98.44
95.20
89.52
89.76
90.45
98.86

2.6  102
1.8  102
2.4  102
1.6  102
2.6  102
2.2  102
1.4  102
3.0  102
7.1  102
7.3  102
2.6  102
2.9  102
2.4  102
1.9  102
2.8  102
2.7  102
2.3  102
2.2  102
2.9  102
2.6  102
2.3  102
1.9  102
1.8  10-2
1.9  102
2.6  102
2.4  102
2.3  102
7.3  102
7.1  102
6.9  102
3.1  102
3.2  102
7.3  102

(@ 200 min)
(@ 220 min)
(@ 180 min)

(@
(@
(@
(@
(@
(@

80 min)
100 min)
80 min)
80 min)
80 min)
80 min)

I.A. Pawar et al. / Ultrasonics Sonochemistry 21 (2014) 11081116

Fig. 6. Effect of depth of horn on degradation of PVP using ultrasound (sonication


time 180 min; 1% PVP; volume 200 ml;300 K).

Fig. 7. Effect of air sparging at different ow rate on degradation of PVP using


ultrasound (sonication time 180 min; 1% PVP; volume 200 ml; depth of horn
0.5 cm; 300 K).

be seen from gure that viscosity reduction rapidly falls at lower


ow rate of air and limiting viscosity is reached around 140 min
of sonication time. It was also observed that the limiting viscosity
was marginally lower at higher ow rate of air i.e.35 cm3/s. The enhanced rate of decrease in viscosity in the presence of aeration (at
lower ow rate) can be attributed to the fact that the presence of
air ow signicantly increases the cavitational effect by supplying
more number of nuclei for the process [49]. It has been observed
that at early time periods, the intrinsic viscosity decreases rapidly
in the presence of air. In absence of external aeration, the degassing
effect of ultrasound ensures that there is less amount of dissolved
gas in the liquid. This gives a lesser quantum of nuclei and hence
lowers cavitational intensity resulting in slower rate of decrease
in the intrinsic viscosity. At signicantly higher ow rates of air,
the degradation was observed to be slower attributed to the fact
that too much cavitation leads to lower cavitational intensity due
to the cushioning effect.
3.6. Effect of different salts
The effect of presence of salt such as sodium sulphate, sodium
chloride and sodium carbonate used as catalyst on viscosity reduction of PVP solution (1% initial loading) at constant ultrasonic
power has been investigated over varying concentration range of
16% (by weight) of the salt over a xed time of treatment as
180 min for comparison purpose. Fig. 8 shows the effect of different
concentration of salts on the extent of degradation of PVP solution

1113

Fig. 8. Effect of different salts on degradation of PVP using ultrasound (sonication


time 180 min; 1% PVP; volume 200 ml; depth of horn 0.5 cm; 300 K).

after treatment of 180 min. It has been observed that by adding


the salts (at optimum concentration of 1%), the extent of reduction
in the intrinsic viscosity of PVP solution changes signicantly. This
is due to the fact that the presence of salt increases the presence of
polymer molecules at the site of cavity collapse and also alters the
surface tension, vapour pressure and ionic strength of the aqueous
phase [50]. All these factors help in a more violent collapse of
bubbles, resulting in enhanced degradation of polymer. However,
the effect is only marginal at higher loadings of salts. It has been
observed that with an increase in concentration of salts such as
sodium sulphate, sodium chloride, the extent of viscosity reduction
decreases marginally. The maximum extent of degradation (83.8%)
was observed with 1% NaCl at 27 C constant temperature.
Previous studies [51] illustrated that the addition of salt increases
the ionic strength of the aqueous phase which drives the organic
pollutants towards the bubblebulk interface and gives signicant
effects for the hydrophobic compounds. For hydrophilic compounds, marginal enhancement due to the presence of NaCl can
be attributed to enhanced cavitational activity due to the changes
in physicochemical properties leading to lower attenuation of
sound waves and higher cavitational intensity [52].
3.7. Effect of surfactant on extent of viscosity reduction
The effect of surfactant on the viscosity reduction at constant
reaction volume of 200 ml of PVP was investigated for different
concentrations of surfactant i.e. sodium lauryl sulfate, SLS (0.1%
and 1%). The obtained results are illustrated in Fig. 9. It has been
observed that, at same ultrasonic power supply, the polymer degradation is marginally affected at 0.1% concentration of sodium
lauryl sulfate (SLS). The extent of degradation is slightly lower in
the presence of SLS as compared to that obtained in the absence
of SLS, which is also conrmed by the observed value of kinetic rate
constant (Table 1). In addition to the quantication of the viscosity
reduction, percentage degradation calculation is very important
due to the variation in the initial viscosity of the polymer solution
by the presence of surfactant. The extent of degradation of PVP at
higher concentration of SLS (1%) loading was about 65% whereas
the extent of degradation of PVP in the absence of SLS was 84%.
The observed results can be attributed to the fact that the presence
of surfactant in higher concentration affects the growth of bubble
adversely. Crum [53] investigated the effect of surface active solutes on rectied diffusion growth of bubble during cavitation. It
was established that surfactant molecules are adsorbed at bubble/solution interface restricting the growth of the bubbles. Lee
et al. [54] also conrmed and observed that surfactant molecules

1114

I.A. Pawar et al. / Ultrasonics Sonochemistry 21 (2014) 11081116

Fig. 10. Effect of TiO2 on degradation of PVP using ultrasound (sonication time
180 min; 1% PVP; volume 200 ml; depth of horn 0.5 cm; 300 K).
Fig. 9. Effect of SLS on degradation of PVP using ultrasound (sonication time
180 min; 1% PVP; volume 200 ml; depth of horn 0.5 cm; 300 K).

are adsorbed at the bubble/solution interface affecting the rate of


growth of acoustic bubbles. Ashokkumar et al. [55] also reported
that bubble growth by coalescence pathway is restricted leading
to a relatively slower growth of the bubbles by rectied diffusion
pathway. At gasliquid interface, the thermodynamic stability of
the solution is critical, whereas adding SLS into the solution results
in altering the stability of acoustic bubble and the process of intramolecular energy transfer (macroscopic effects) as well as the
intermolecular energy transfer (microscopic effects) during the rapid collapse of the bubbles [52,53,56].
3.8. Effect of titanium dioxide on extent of viscosity reduction
The presence of solid particles such as titanium dioxide provides an additional phase in the system which can increase the
cavitational activity by allowing more bubbles to form and grow
into active cavitation bubbles. Hence, the number of cavitation
events occurring in the reactor is enhanced resulting in a subsequent enhancement in the cavitational activity. It must be noted
that the presence of solids also have a negative impact on the cavitational activity as the solid particles result in scattering of the
sound waves thereby decreasing the focused energy transferred
into the system. The net effect of these two phenomena will be
dependent on the system in question and hence optimization is
the required before operating parameters are selected for the actual operation [5759].
The obtained results for effect of TiO2 on viscosity reduction of
PVP are depicted in Fig. 10. It can be seen that intrinsic viscosity
reduction is higher and signicantly faster at lower concentration
(0.1%) as compared to the operation involving only ultrasound. It
has been observed that the nal limiting viscosity is also reduced
in the presence of 0.1% loading of TiO2. The observed enhancement
can be attributed to the fact that solid particles provides additional
nuclei which promote cavitational activity [18]. Similar results
were obtained by Guo et al. [60] for the degradation of 2,4-dinitrophenol (at 20 mg/l) which showed higher degradation with cupric
oxide than in the absence of CuO. Shirgaonkar and Pandit [61] reported that the extent of degradation of trichlorophenol was higher (16.8%) at 0.1 g/l of TiO2 loading as compared to that in the
absence of TiO2 (10.2%).
Results with higher concentration of TiO2 (1%) revealed that the
rate of viscosity reduction was lower as compared to the use of
lower TiO2 concentration (0.1%), although better as compared to
the sonication alone. The observed results can be attributed to
the fact that at higher loadings of solid particles in the system,
the incident sound waves are scattered depending on the surface

Fig. 11. Effect of US and UV irradiation on degradation of PVP using ultrasound


(sonication time 180 min; 1% PVP; volume 200 ml; depth of horn 0.5 cm; 300 K,
5 W of UV tube).

characteristics and the projected area, resulting in a lower amount


of energy dissipated into the system due to the ultrasonic irradiation. If the scattering of sound waves is dominant, as the case may
be in the use of TiO2 at higher concentrations, slower viscosity
reduction is observed due to lower degree of cavitational effects
[58,62].

3.9. Effect of combination of ultrasound (US) and ultraviolet


irradiations (UV)
The results for this combination approach have been presented
in Fig. 11. It can be seen that the extent of intrinsic viscosity reduction is higher for the combination mode as compared to the individual operations. Both US and UV irradiations have some
potential to degrade PVP and UV has slightly better efcacy as
compared to the use of ultrasound. Use of UV irradiations with
11 W power rating results in faster reduction in the viscosity
though the nal limiting viscosity is about 10% higher as compared
to the use of ultrasonic irradiations alone. This is due to the fact
that the UV irradiation involves abstraction of hydrogen atom from
the molecule by a radical which is produced from the water molecule. These radicals react with oxygen, forming other radicals
which are responsible for chain scission into smaller chains and
other molecules [63]. Thus the presence of ultraviolet radiations
provides an additional mechanism for polymer degradation.
A more careful observation leads to the fact that when both the
types of irradiations are operated in combination, the value of the

I.A. Pawar et al. / Ultrasonics Sonochemistry 21 (2014) 11081116

limiting viscosity is lowered and at a higher rate indicating the


superiority of the combined operation. Thus, for the lower viscosity
reduction of PVP, use of US and UV irradiation in the combined
manner would be much better option as compared to individual
operating mode. Quantitatively speaking, maximum extent of degradation i.e. 98.8% has been observed at 100 min of operation using
combination of US and UV whereas in the case of ultrasound operated individually, the extent of degradation was only 83%. The
important objective of the present work is to highlight the utility
of combined operation for the degradation of PVP polymeric solution. The obtained results are very important as there have been no
prior studies that elucidate the use of combined irradiation approach for the polymer degradation (especially for PVP).

3.10. Effect of power intensity of UV irradiations


Addition of energy via UV photons can produce signicant
absorption of energy in the molecule at sufcient levels that can
lead to effective breaking of the molecular bonds or effective cutting of molecular chains which leads to a reduction in the intrinsic
viscosity. The results of the experiments, with varying power
intensity of the UV tube carried out using ultrasonic horn for
enhancing the degradation of PVP, are depicted in Fig. 12. The results indicate that intrinsic viscosity decreases with increasing
power intensity. It has been observed that at initial stages the viscosity reduction was higher at 5 W. With increasing treatment
time, the viscosity reduction seems to be similar upto 60 min for
all power intensities, and after that the viscosity reduction increases rapidly for the case of higher power rating i.e. 11 W. UV
radiation in the range of 290400 nm, when absorbed by polymer
molecule, is strong enough to rupture the most chemical bonds.
Also it has been said that the essential photochemical reactions
in polymers comprise of the cleavage of main chains, the combination of different macromolecules (cross linking), the generation of
unsaturated groups and cleavage of side groups resulting in formation of volatile products. Generally, the lateral groups are split off
in the primary process and further intermediates decompose
involving bond rupture in the main chain. Also, the main chain
bonds are split leading to a pair of terminal radicals. It is expected
that the these effects are higher at higher power dissipation levels
and hence the extent of degradation and the rate of polymer breakage will be higher at 11 W as compared to that obtained at 5 W and
9 W. Kinetic study revealed that with a change in the power dissipation from 5 W to 11 W, the rate constant was found to increase
from 3.1  102 to 7.3  102.

Fig. 12. Effect of power intensity on degradation of PVP using ultrasound


(sonication time 180 min; 1% PVP; volume 200 ml; depth of horn 0.5 cm; 300 K,
5 W,9 W and 11 W of UV tube).

1115

4. Conclusions
The degradation of polyvinyl pyrrolidone (PVP) has been successfully demonstrated using ultrasonic irradiations in combination with process intensifying approaches and it has been
conrmed that the use of ultrasound effectively results in the
breaking of polymer chains. It can be also concluded that the extent of degradation and the viscosity reduction is strongly dependent on the operating parameters such as reaction volume, depth
of horn, and initial concentration. The molecular weight or intrinsic
viscosity decreases with an increase in the sonication time and
reaches to a limiting value, below which no further degradation occurs. It has been also established that the use of simple additives
such as salts, titanium dioxide and introduction of air at different
ow rates can enhance the rate of ultrasonic degradation of PVP
effectively. Higher enhancement in rates of degradation is achieved
when optimized amounts of additives are used. Presence of surfactants leads to marginally detrimental effects at lower concentration and in general should not be used as desired additives for
process intensication. Among the different additives investigated
in the work, titanium dioxide is most effective in depolymerizing
the PVP using ultrasonic horn. It has been observed that the maximum extent of degradation is obtained by combination of ultrasound and ultraviolet irradiation. It has been also established
that major extent of degradation takes place in the initial phase
of irradiation. This shows that a continuous operation with ultrasonic horn is indeed possible, which is important for treatment
of large volumes of the streams containing low concentrations of
PVP.

References
[1] T.J. Mason, J.P. Lorimer, Applied Sonochemistry, Wiley, VCH, 2002.
[2] J.P. Lorimer, T.J. Mason, T.C. Tuthbert, E.A. Broikeld, Effect of ultrasound on
the degradation of aqueous native dextrin, Ultrason. Sonochem. 2 (1) (1995)
5157.
[3] M.J. Cauleld, G.G. Quiao, D.H. Solomon, Degradation on polyacrylamides Part
II. Polyacrylamide gels, Chem. Rev. 102 (2002) 30673083.
[4] M.M. Fares, J. Hacaloglu, S. Suzer, Characterization of degradation products of
polyethylene oxide by pyrolysis mass-spectrometry, Eur. Polym. J. 30 (1994)
845850.
[5] A.M. Omar, F. Heatley, A 13C NMR study of the products and mechanism of the
thermal oxidative degradation of poly (ethylene oxide), Macromol. Chem.
Phys. 203 (2002) 22732280.
[6] H. Kaczmarek, A. Sionkowska, A. Kaminska, J. Kowalonek, M. Swiatek, A. Szalla,
The inuence of transition metal salts on photo-oxidative degradation of
poly(ethylene oxide), Polym. Degrad. Stab. 73 (2001) 437441.
[7] J.F. Rabek, L.A. Linden, H. Kaczmarek, B.J. Qu, W.F. Shi, Photodegradation of
poly(ethylene oxide) and its coordination-complexes with iron(III)chloride,
Polym. Degrad. Stab. 37 (1992) 3340.
[8] H. Kaczmarek, A. Kaminska, L.A. Linden, J.F. Rabek, Photo-oxidative
degradation of poly(ethylene oxide) copper chloride complexes, Polymer 37
(1996) 40614068.
[9] S.W. Shalaby, Bioabsorbable Polymers, in: J. Swarbrick, J.C. Boylan (Eds.),
Encyclopedia of Pharmceutical Technology, M. Dekker, Inc. New York, 1, 1988,
465476.
[10] S.J. Holland, B. Tighe, Biodegradable polymers, in: Advances in Pharmaceutical
Science, Academic Press, London, 1992, pp. 101164. vol. 6.
[11] T. Hayashi, Biodegradable polymers for biomedical uses, Prog. Polym. Sci. 19
(1994) 663702.
[12] J. Kohn, R. Langer, Bioresorbable and bioerodible materials, in: B.D. Ratner, A.S.
Hoffman, F.J. Schoen, J.E. Lemon (Eds.), An Introduction to Materials in
Medicine, Academic Press, San Diego, 1997, pp. 6573.
[13] E. Behravesh, A.W. Yasko, P.S. Engle, A.G. Mikos, Biodegradable polymers for
orthopaedic applications, Clin. Orthop. 367S (1999) 118185.
[14] J.C. Middleton, A.J. Tipton, Synthetic biodegradable polymers as orthopedic
devices, Biomaterials 21 (2000) 23352346.
[15] S. Ramakrishna, J. Mayer, E. Wintermantel, K.W. Leong, Biomedical
applications of polymer-composite materials: a review, Compos. Sci.
Technol. 61 (2001) 11891224.
[16] V.S. Papkov, The Great Soviet Encyclopedia, third ed., The Gale Group, Inc,
2010, 19701979.
[17] T.J. Mason, Practical Sonochemistry: Users Guide in Chemistry and Chemical
Engineering. Ellis Horwood Series in Organic Chemistry, Ellis Horwood Ltd.,
Chinchester, UK, 32, 1992, 453458.

1116

I.A. Pawar et al. / Ultrasonics Sonochemistry 21 (2014) 11081116

[18] P.R. Gogate, Cavitational reactor for the process intensication of chemical
processing applications: a critical review, Chem. Eng. Proc. 47 (2008) 515
527.
[19] G.J. Price, The use of ultrasound for the controlled degradation of polymer
solutions, in: T.J. Mason (Ed.), Advances in Sonochemistry, JAI Press,
Cambridge, 1990, pp. 231239. vol. 1.
[20] T.J. Mason, D. Peter, Practical Sonochemistry: Power Ultrasound Uses and
Application, second ed., Ellis Horwood series in organic chemistry, Ellis
Horwood Ltd., Chinchester, UK. 6, 2002, 313323.
[21] A. Gronroos, P. Pirkonen, J. Heikkinen, J. Ihalainen, H. Mursunen, H. Sekki,
Ultrasonic depolymerization of aqueous polyvinyl alcohol, Ultrason.
Sonochem 8 (2001) 259264.
[22] J.P. Lorimer, T.J. Mason, T.C. Cuthbert, E.A. Broikeld, Effect of ultrasound on
the degradation of aqueous native dextran, Ultrason. Sonochem 2 (1995) 55
57.
[23] V. Desai, M.A. Shenoy, P.R. Gogate, Ultrasonic degradation of low-density
polyethylene, Chem. Eng. Proc. 47 (2008) 14511461.
[24] A.M. Bisdow, K.H. Ebert, Mechanism of degradation of polymers in solution by
ultrasound, Makromol. Chem. 176 (1975) 745756.
[25] W. Schnable, Polymer Degradation: Principles and Practical Applications,
Hanser International, New York, 1981, pp. 227240.
[26] J.G. Price, T.F. Smith, Ultrasonic degradation of polymer solution, 2. The effect
of temperature, ultrasound intensity and dissolved gases on polystyrene in
toluene, Polymer 34 (1993) 41114117.
[27] A. Gronroos, P. Pentti, K. Hanna, Ultrasonic degradation of aqueous
carboxymethylcellulose: effect of viscosity, molecular mass, and
concentration, Ultrason. Sonochem. 15 (2008) 644648.
[28] A. Weissler, Depolymerization by ultrasonic irradiation: the role of cavitation,
J. Appl. Phys. 21 (1950) 171176.
[29] T.J. Mason, Chemistry with Ultrasound, Elsevier Applied Science, London, UK,
1991.
[30] A. Akyuz, H. Catalgilgiz, A.T. Giz, Kinetics of ultrasonic polymer degradation:
comparison of theoretical models with on-line data, Macromol. Chem. Phys.
209 (2008) 801809.
[31] M.T. Taghizadeh, A. Mehrdad, Calculation of the rate constant for the
ultrasonic degradation of aqueous solutions of polyvinyl alcohol by
viscometry, Ultrason. Sonochem. 10 (2003) 309313.
[32] H. Kim, J.W. Lee, Study of application of ultrasonic wave to injection molding,
Polymer 43 (2002) 25852587.
[33] S. Koda, H. Mori, K. Matsumoto, H. Nomura, Ultrasonic degradation of water
soluble polymers, Polymer 34 (1993) 3036.
[34] J. Chakraborty, J. Sarkar, R. Kumar, G. Madras, Ultrasonic degradation of
polybutadiene and isotactic polypropylene, Polym. Degrad. Stab. 85 (2004)
555558.
[35] A. Noor, B. Ahmed, A. Bhettani, Part 1: Viscometric light scattering studies of
poly (vinyl pyrrolidone), J. Chem. Soc. Pakistan 13 (1991) 153156.
[36] P.R. Gogate, I.Z. Shirgaonkar, M. Shivkumar, P. Senthilkumar, N.P. Vichare, A.B.
Pandit, Cavitation reactors: efciency analysis using a model reaction, AIChE J.
47 (2001) 25262538.
[37] D.W. van Krevelen, Properties of Polymers, third ed., Elsevier, Amsterdam,
1990.
[38] J. Brandrup, E.H. Immergut, Polymer Handbook, second ed., Wiley Interscience, New York, 1975.
[39] O. Solomo, Z.Z. Ciutta, De termination de la viscosite intrinseque de solutions
de polyme res par une simple de termination de la viscosite, J. Appl. Polym.
Sci. 6 (1962) 683686.
[40] IUPAC Compendium of chemical technology, second ed., Blackwell Science,
1997.

[41] N.K. Baramboim, Mechanochemistry of Polymers, Rubber and Plastic Research


Association of Great Britain, Maclaren, UK, 1964.
[42] J. Li, S. Guo, X. Li, Degradation kinetics of polystyrene and EPDM melts under
ultrasonic irradiation, Polym. Degrad. Stab. 89 (2005) 614.
[43] M. Sivakumar, A.B. Pandit, Ultrasound enhanced degradation of Rho-damine B:
optimisation with power density, Ultrason. Sonchem. 8 (2001) 233240.
[44] U.D. Harkal, P.R. Gogate, A.B. Pandit, M.A. Shenoy, Ultrasonic degradation of
poly (vinyl alcohol) in aqueous solution, Ultrason. Sonochem 423 (2006) 423
428.
[45] A. Bhatnagar, H.M. Cheung, Sonochemical destruction of chlorinated C1 and C2
volatile organic compounds in dilute aqueous solutions, Environ. Sci. Technol.
28 (1994) 14811486.
[46] H.M. Hung, M.R. Hoffmann, Kinetics and mechanism of the sonolytic
degradation of chlorinated hydrocarbons: frequency effects, J. Phys. Chem. A
103 (1999) 27342739.
[47] P.R. Gogate, A.B. Pandit, Sonochemical reactors: scale up aspects, Ultrason.
Sonochem. 11 (2004) 105117.
[48] C. Torres-sanchez, J.R. Corney, Effect of ultrasound on polymeric foam porosity,
Ultrason. Sonochem. 15 (2008) 408409.
[49] M.H. Entezari, P. Kruus, Effect of frequency on sonochemical reactions. I:
Oxidation of iodide, Ultrason. Sonochem. 1 (1994) S7589.
[50] S. Findik, G. Gunduz, Sonolytic degradation of acetic acid in aqueous solutions,
Ultrason. Sonochem. 14 (2007) 157169.
[51] J.D. Seymour, R.B. Gupta, Oxidation of aqueous pollutants using ultrasound:
salt-induced enhancement, Ind. Eng. Chem. Res. 36 (1997) 34533469.
[52] V.S. Sutkar, P.R. Gogate, Mapping of cavitational activity distribution in high
frequency reactors at pilot scale operation, Chem. Eng. J. 158 (2010) 296304.
[53] L.A. Crum, Measurements of the growth of air bubbles by rectied diffusion, J.
Acoust. Soc. Am. 68 (1980) 203210.
[54] J. Lee, M. Ashokumar, S.E. Kentish, F. Grieser, Effect of alcohols on the initial
growth of multibubble sonoluminescence, J. Phys. Chem. B 110 (2006) 17282
17285.
[55] M. Ashokkumar, F. Grieser, The effect of surface active solutes on bubbles in an
acoustic eld, Phys. Chem. Chem. Phys. 9 (2007) 56315643.
[56] A.V. Mohod, P.R. Gogate, Ultrasonic degradation of polymers: Effect of
operating
parameters
and
intensication
using
additives
for
carboxymethylcellulose (CMC) and polyvinyl alcohol (PVA), Ultrason.
Sonochem. 18 (2011) 727739.
[57] P.R. Gogate, A.M. Wilhelm, B. Ratsimba, H. Delmas, A.B. Pandit, Destruction of
formic acid using high frequency cup horn reactor, Water Res. 40 (2006) 1697
1705.
[58] S.N. Katekhaye, P.R. Gogate, Intensication of cavitational activity in
sonochemical reactors using different additives: Efcacy assessment using a
model reaction, Chem. Eng. Proc. 50 (2011) 95103.
[59] T. Tuziuti, K. Yasui, M. Sivakumar, Y. Iida, Correlation between acoustic
cavitation noise and yield enhancement of sonochemical reaction by particle
addition, J. Phys. Chem. A 109 (2005) 48694872.
[60] Z. Guo, R. Feng, J. Li, Z. Zheng, Y. Zheng, Degradation of 2,4-dinitrophenol by
combining sonolysis and different additives, J. Hazard. Mater. 158 (2008) 164
169.
[61] I.Z. Shirgaonkar, A.B. Pandit, Sonophotochemical destruction of aqueous
solution of 2,4,6-trichlorophenol, Ultrason. Sonochem. 5 (1998) 5361.
[62] N. Shimizu, C. Ogino, M. Dadjour, K. Ninomiya, A. Fujihira, K. Sakiyama,
Sonocatalytic facilitation of hydroxyl radical generation in the presence of
TiO2, Ultrason. Sonochem. 15 (2008) 988994.
[63] T. Aarthi, M.S. Shaama, G. Madras, Degradation of water soluble polymer under
combined ultrasonic and ultraviolet radiation, Ind. Eng. Chem. Res. 46 (2007)
62046210.

Das könnte Ihnen auch gefallen