You are on page 1of 9

b i o s y s t e m s e n g i n e e r i n g 1 1 5 ( 2 0 1 3 ) 2 7 4 e2 8 2

Available online at www.sciencedirect.com

journal homepage: www.elsevier.com/locate/issn/15375110

Research Paper

A novel slow-release urea fertiliser: Physical and


chemical analysis of its structure and study of its
release mechanism
Ni Xiaoyu, Wu Yuejin*, Wu Zhengyan, Wu Lin, Qiu Guannan, Yu Lixiang
Key Laboratory of Ion Beam Bio-engineering, Institute of Technical Biology & Agriculture Engineering of Chinese
Academy of Sciences, 350# Shushanhu Road, Hefei 230031, PR China

article info

Reducing the release rate of urea can increase its efficiency of use and reduce nitrogen

Article history:

pollution. A slow-release urea (S-urea) was produced using a new method; a bentonite and

Received 12 February 2012

organic polymer (OP) were used to form a three-dimensional lattice structure by melting urea

Received in revised form

directly. The structure affected the recrystallisation of urea and increased its stacking density.

28 September 2012

The specific surface area of S-urea was 0.046 m2 g1, much lower than that of common urea

Accepted 5 April 2013

(1.698 m2 g1). The static release experiment showed that 75% of 12 g sample of S-urea was

Published online 17 May 2013

released in 1 l water for about 14 h, much longer than that of common urea (<0.5 h). The kinetic
simulation results showed that the release of S-urea was not based on Fickian diffusion but
underwent anomalous diffusion with its release rate was mainly affected by the dissolvingeroding process of the medium which was controlled by the compactness of the lattice
structure. This process may be strengthened by increasing the amount of bentonite.
2013 IAgrE. Published by Elsevier Ltd. All rights reserved.

1.

Introduction

Nitrogen is the necessary nutrition for plant growth and it is


the most important factor commonly considered as being
yield limiting (Ni, Liu, & Lue, 2009). Urea is the most widely
used nitrogen fertiliser in world agriculture because of its high
nitrogen content (46%) (Trenkel, 1997; Zheng, Liang, Ye, & He,
2009). However, urea cannot be easily fixed by soil particles
before hydrolysation as it is a neutral organic molecule. It has
been estimated that only 30e50% of the dose of nitrogen
applied as urea can be recovered by plants (Al-Zahrani, 1999;
Prasad, Rajale, & Lacakhdive, 1971), consequently run-off occurs with serious environmental consequences. The leaching
losses from conventionally formulated urea contribute greatly
to the non-point source pollution and the eutrophication of
lakes and reservoirs (Diez et al., 1994; Li & Yang, 2004). The

various environmental and economic drawbacks associated


with the use of conventional urea have therefore become a
focus of worldwide concern (James & Sojka, 2008; Xie, Liu, Ni,
Zhang, & Wang, 2011).
An effective method of mitigating the problem is to
develop slow-release urea. Much research has reported on the
improved performance of the coated urea where a core is
encapsulated within an inert carrier (Govind & Sharma, 1979;
Han, Chen, & Hu, 2009; Liu, Wang, Qin, & Jin, 2008). The slowrelease of coated urea is controlled by diffusion through the
coating; sulphur-coated urea, polyethylene-coated urea, and
superphosphate-coated urea are all typical examples (Salman,
1989; Subrahmanyan & Dixit, 1988). Coated urea provides a
much longer release time and higher utility rate, however it is
mostly used in developed countries. It has not been popular in
developing countries because of its higher cost. However,

* Corresponding author. Tel.: 86 551 5593172; fax: 86 551 5595670.


E-mail address: yjwu@ipp.ac.cn (W. Yuejin).
1537-5110/$ e see front matter 2013 IAgrE. Published by Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.biosystemseng.2013.04.001

b i o s y s t e m s e n g i n e e r i n g 1 1 5 ( 2 0 1 3 ) 2 7 4 e2 8 2

Nomenclature
Symbols
A
Absorbency value
C
Constant incorporating the characteristics of the
adsorption of N2
Ct
Concentration of urea at time t (mg ml1)
k
Constant of the carrier-active agent system
Diffusion constant
k1
Dissolving-erosion constant
k2
m
Diffusion exponent
n
Diffusion exponent
P
Partial pressure of N2 (Pa)
Saturated vapour pressure of liquid N2 (Pa)
P0
Fraction of active agent released at time t
Qt
Coefficient of determination
R2

developing countries consume more and more nitrogen fertiliser and yet have only 20e35% efficiency of nitrogen use
(Fan & Liao, 1998; Jiang, Hu, Sun, & Huang, 2010).
In this study a novel slow-release urea (S-urea) is presented whose structure and release mechanism is quite
different from that of coated urea. The formulation forms a
three-dimensional lattice structure in the urea solution that
could influence its release process (Cai et al., 2009; Chinese
Patent Specification ZL200610040631.1, 2006). The method of
production is developed by melting urea directly and using
bentonite which is a cheap and safe material as a main
substrate (Chinese Patent Specification CN201110003090.6,
2011). This new type of urea can reduce costs greatly since
it increases cost only by about 30e50U t1 above common
urea and it is much cheaper than coated urea
(200e2000U t1). This improvement should make this new
technique popular, particularly in developing countries. In
order to investigate the slow-release mechanism of this
new type of urea, its structure was analysed using infrared
spectra (IR), scanning electron microscopy (SEM), X-ray
diffraction (XRD) techniques and a static release experiment
designed mainly according to the model of Higuchi (1963).
The affect of the proportion of additives was tested using
the release kinetics data and the results simulated using the
equation of Peppas (Lenaerts, Dumoulin, & Mateescu, 1991;
Peppas, 1985) and the double-exponent equation (Kaunisto,
Marucci, Borgquist, & Axelsson, 2011; Peppas & Sahlin,
1989).

2.

Materials and methods

2.1.

Materials

Bentonite (Zhejiang Fenghong Bentonite Co. Ltd., China)


sieved through a 200 mesh screen was washed with distilled
water, and then dried at 105  C for 8 h before use. Organic
polymer (OP) (chemically pure, Shanghai Chemical Regent
Factory, Shanghai, China) and urea (Shanghai Chemical
Regent Factory, Shanghai, China) were dried at 80  C for 8 h
before use.

t
V
Vm

275

Time (h)
Total gas volume adsorbed by sample (ml)
Gas volume adsorbed by monolayer of sample (ml)

Abbreviations
BET
Brunauer, Emmett, & Teller equation to calculate
specific surface area B-urea Urea added with
bentonite
IR
Infrared spectra
OP
Organic polymer
P-urea Urea added with organic polymer.
SEM
Scanning electron microscopy
S-urea Slow-release urea added with bentonite and
organic polymer
XRD
X-ray diffraction

2.2.
Preparation of bentonite-urea (B-urea), organic
polymer-urea (P-urea) and slow-release urea (S-urea)
A quantity of bentonite (5%) was prepared and mixed equally in
the melting urea according to the method in Chinese Patent
Specification (CN201110003090.6). The admixture was taken
into a mould and recrystallised at room temperature; the final
product (B-urea) was dried at 80  C for 8 h before use. Using a
quantity of OP (0.15%) to replace the bentonite and the final
product (P-urea) was also dried at 80  C for 8 h before use.
OP with a proportion of bentonite (from 1% to 5%) was added
to the urea and mixed according to the same method. The final
product of S-urea was dried at 80  C for 8 h before use.

2.3.

Physical and chemical analysis of structure

The common urea and S-urea samples were tested respectively


using IR, SEM and XRD. The IR spectra were obtained in the wave
number range of 400e4000 cm1 using a Fourier transform IR
spectrophotometer (Alpha-T, Bruker Company, Germany). The
SEM images were recorded using scanning electron microscope
(Sirion200, FEI Company, USA). The common urea and S-urea
samples were scanned in the angle range of 10 e60 on the instrument of X-ray diffraction (Xpert, Philips Company, Holland).
The specific surface area of the two samples was also
measured (Ommishop 100CX, Coulter Company, USA) and information on specific surface area and pore size distribution was
obtained using Brunauer, Emmett, and Teller equation (1938):
P
1
C  1P

VP0  P Vm C Vm C P0
where P is partial pressure of N2 (Pa); P0 is saturated vapour
pressure of liquid N2 (Pa); Vm is the gas volume adsorbed by
monolayer of sample (ml); V is the total gas volume adsorbed
by sample (ml); C is a constant incorporating the characteristics of the adsorption.

2.4.

Static release experiment in water

The experimental apparatus for determining the static release


of urea in water is shown in Fig. 1. Samples of about 12 g of the

276

b i o s y s t e m s e n g i n e e r i n g 1 1 5 ( 2 0 1 3 ) 2 7 4 e2 8 2

Another model double-exponent equation (Peppas &


Sahlin, 1989) was proposed:
Qt k1  tm k2  t2m

(2)

where Qt is the fraction of active agent released at time t, k1 is


the diffusion constant, k2 is the dissolving-erosion constant
and m is the diffusion exponent. The first item k1tm indicates
the cumulative release rate by the diffusion, and the second
item k2t2m indicates the cumulative release rate by dissolution
of the auxiliary frame by water.

2.6.

Some experiments were carried out to test the slow-release


effect of S-urea in soil at room temperature: 0.5 g common
urea and S-urea were placed respectively in a flowerpot with
200 g light clay soil (d  2 mm, 100% soil field capacity), urea
granules were placed about 10 mm deep and 50 mm high on the
soil, and a perforated film was placed over the flowerpot to
reduce volatilisation of water. After 24 h incubation, 50 ml of
water was sprinkled evenly on the sample and this produced a
leachate, the concentration of urea and total N in the leachate
were tested. The leaching process was repeated at intervals of
24 h and the whole experiment was carried out over two weeks.

Fig. 1 e Experimental device for testing static release. 1


Urea sample, 2 sample pipe, 3 thermometer, 4 water, 5
vessel, 6 magnetic stirring rod, 7 magnetic stirring
apparatus.

different types of urea (common urea, P-urea, B-urea and Surea) were poured into a pipe which was 100 mm long and
10 mm inside diameter with one end closed. After the urea
recrystallised at room temperature the pipe was placed horizontally with 1 l water in the apparatus.
The components of the apparatus were assembled
together according to Fig. 1. The speed of magnetic stirrer was
10 revolutions s1 that should have ensured that the concentration of the solution was uniform and not affecting
diffusion. At given time intervals the concentration of urea at
3 different positions in the centre of the vessel was determined. The water temperature was controlled at 25  C and the
pipe was turned 90 every 15 min to keep the area of the
interface constant. All the results were based on three
replicates.

2.5.

The model of urea static release rate in water

The experiment was designed according to the procedure of


Higuchi (1963). As the solubility of urea was quite large
(120.17 g urea can dissolve in 100 g water at 25  C) and there in
total about 1% urea in the water, the device could fit the hypothesis of Higuchi as an infinite-trap. In this condition,
when the interface between urea and water moves, the concentration gradient in the pipe can be ignored and the urea
concentration in the pipe can be assumed uniform and
assumed to be equal to that in the whole vessel.
Urea release data were analysed using the equation by
Peppas (1985):
Qt k  tn

Slow-release effect of S-urea in practical condition

(1)

where Qt is the fraction of active agent released at time t, k a


constant incorporating the characteristics of the carrier-active
agent system, and n the diffusion exponent, indicative of the
transport mechanism.

2.7.

Determination of urea and total N concentration

The concentration of urea in water was determined according


to method for the determination of urea residues in canned
mushrooms for export in the Specialised Standard of Peoples
Republic of China (SN/T 1004-2001). The complex compound
of urea and p-dimethylaminobenzaldehyde were detected
using a spectrophotometer (UV-2550, Shimadzu Company,
Japan) operating at a wavelength of 440 nm, and the concentration of urea was calculated according to following formula:
1 

Ct 3:7487  A 0:0171 mg ml

where Ct was the concentration of urea, A was the absorbency


value and the coefficient of determination (R2) was 0.999.
The concentration of total N was determined according to
method for the determination of total N in water in The Specialised Standard of Peoples Republic of China (GB 11894-89).
The NO
3 was also detected using a spectrophotometer (UV2550, Shimadzu Company, Japan) at wavelengths of 220 nm
and 275 nm.

2.8.

Statistical method

The statistical results and the nonlinear fit of Eq. (1) and Eq. (2)
were calculated using Origin software (Origin 8.725, Originlab
Company, USA).

3.

Results and discussion

3.1.

Morphology and physical structure

Figure 2 shows SEM images of the surface of common urea


and S-urea. The surface of common urea was even and its

b i o s y s t e m s e n g i n e e r i n g 1 1 5 ( 2 0 1 3 ) 2 7 4 e2 8 2

277

Fig. 2 e SEM images of the morphology and physical structure. (A and B) the surface of common urea, (C and D) the surface of
S-urea.

molecules ordered forming a uniform layer (Fig. 2A and B). The


bentonite molecules inside S-urea formed an irregular surface
like a membrane with disordered mesh (Fig. 2C and D). Inorganic molecules may connect with each other mainly by
electrostatic attraction using its double-electronic layer
(Bhattacharjee & Elimelech, 1997; Cai et al., 2009). If bentonite
is dispersed homogeneously by full blending it should form a
lattice structure in three-dimensional space. OP should
dissolve in melting urea and extend its long chain in the solution, in this way it cross-linked the bentonite molecules and
strengthened the lattice frame (Fig. 2D).

3.2.
Analysis of the physical and chemical character of
S-urea
The IR spectra of S-urea sample were similar to that of common urea (Fig. 3A). The peaks at 3447 and 3343 cm1 of common urea as well as of S-urea could be assigned to asymmetric
and symmetric stretching vibration of NH2. The peak at
3250 cm1 of both the two types of urea can be assigned to
OeH vibration of absorbed water. The peak at 1688 cm1 can
be assigned to carbonyl (C]O) and 1613 cm1 peak can be
assigned to NeH bending vibration and CeH stretching vibration (mainly NeH bending vibration domain) of O]CeNH2
(He et al., 2007; Xie et al., 2011). In the finger print zone of
1500e400 cm1, all the peaks were similar.
The peaks of 3430 and 1641 cm1 for bentonite can be
assigned to OeH vibration of absorbed water. The peak at
1041 cm1 can be assigned to SieO vibration (Chen, Yang, Luo,
& Lu, 2002). It showed that there might be no chemical reaction during the mixing process and the mixture of molecules
could connect with each other mainly by some physical
attraction such as the Van der Waals force, hydrogen bond
and electrostatic attraction.

The results from X-ray diffraction of common urea and Surea samples are shown in Fig. 3B. Both the samples had the
similar diffraction angle (2q). S-urea had a sharp peak at
22.28 , it was a little less intense than that of common urea at
22.33 , it showed that S-urea had a tighter arrangement of
molecules than common urea using the Bragg calculation
(Ding & Liu, 1998; Zheng, Zhang, Cai, Fu, & Wang, 2005). That
might be because the bentonite was inserted between the gap
in the urea crystals and this influenced its process of
recrystallisation.
The pore size distribution of common urea and S-urea were
obtained by N2 adsorption and desorption experiment and the
specific surface area was calculated using BET equation. As
shown in Fig. 4A, the shaded area indicates the final volume of N2
adsorbed by the sample, the larger shaded area the larger the
surface area. The surface area of common urea was much larger
than that of S-urea. With increasing N2 pressures more gas
comes into contact with the smaller pores and is adsorbed, the
shaded area under the differential relative pressure indicates the
amount of the pore sizes and their proportion. It indicates that
the pore size of S-urea is distributed mainly in a smaller size
range. The specific surface areas of the two forms of urea were
calculated and this is shown in Fig. 4B; common urea was
1.698 m2 g1 and S-urea was 0.046 m2 g1. This result shows that
S-urea has a more compact structure. This reinforces the conclusions that the two additives occupy the potential space inside
urea and form a compact lattice structure with network connections that should greatly decrease the specific surface area.

3.3.

The effect of bentonite and OP on the release of urea

The static release experiment was carried out with the samples of common urea, P-urea, B-urea and S-urea respectively
using the apparatus shown in Fig. 1. The releasing process of

278

b i o s y s t e m s e n g i n e e r i n g 1 1 5 ( 2 0 1 3 ) 2 7 4 e2 8 2

Fig. 3 e Physical and chemical analysis of the structure. (A)


IR spectra of common urea, S-urea and bentonite. (a)
Common urea; (b) S-urea; (c), bentonite. (B) XRD analysis of
common urea and S-urea. (a) Common urea; (b) S-urea.

each kind of sample was tested to investigate the affect of


each of the auxiliary materials, bentonite and OP. Figure 5A
shows the release rate of each type of urea at the specified
intervals of time. The time for common urea to be released
entirely was less than half an hour, the time for P-urea was
about 1 h and that of B-urea was near 9 h. The release rate of Purea was about half that of common urea, but it was still more
rapid than the others. OP can dissolve and extend the length of
its chain in the dissolving urea, making a physical connection
with it, but both OP and urea dissolve easily in water. Therefore, the release rate from P-urea was similar to that of common urea as the water infiltrated into its structure. B-urea had
a longer release time than P-urea, this could be because the
bentonite could not dissolve in water and the lattice framework connected by bentonite increased the path length for the
penetration of water. Furthermore, this action could be more
effective as the bentonite particles absorbed water and
swelled (Slade, Quirk, & Norrish, 1991). The final result
comparing the release times of P-urea and B-urea showed that

Fig. 4 e BET analysis of common urea and S-urea. (A) The


distribution of the pore size of common urea and S-urea. (a)
Common urea; (b) S-urea. (B) Specific surface area of
common urea and S-urea.

the structure of the bentonite was the main factor slowing


down the release of urea.
The release rate of S-urea was slower than that of B-urea, 75%
of B-urea was released for about 7 h and S-urea with the same
amount was released after almost 14 h, approximately double
time compared of B-urea and almost 28 times that of common
urea. This meant that S-urea could have a longer residence time
in soil and the nitrogen released could have more chance of being
used by plants thereby reducing the potential for water pollution.
This result also showed that OP played a very important role
when it was used with bentonite, since it cross-linked the
bentonite particles to form a firm network and strengthened its
lattice structure. These connections were strong and not easily
broken by water because OP was anchored by the bentonite (Li
et al., 2007). Therefore, OP was the auxiliary material that
greatly strengthened the slow release effect of bentonite.

3.4.

The effect of bentonite amount on the release of S-urea

Bentonite was the main additive that could influence the release
rate of S-urea, the static release experiment was carried out with

b i o s y s t e m s e n g i n e e r i n g 1 1 5 ( 2 0 1 3 ) 2 7 4 e2 8 2

Fig. 5 e The characteristic of slow release. (A) Time dependent


release of urea with different auxiliary addition. (B) Time
dependent release of urea with the amount of bentonite.

different amounts of bentonite ranging from 1% to 5%. Figure 5B


shows the results of the relationship between the release time of
urea and the different amounts of added bentonite. About 75% of
common urea was released within half an hour, the same
amount of S-urea was released for about 4 h with 1% of bentonite
added, and this time was extended to about 7.5 h and 14 h
respectively for S-urea with 3% bentonite and S-urea with 5%
bentonite. The S-urea had prominent effect on slowing the
release rate of nitrogen and this can greatly increase its efficiency
of use. The release rate of urea was remarkably reduced by
increasing the amounts of bentonite (from 1% to 5%). Bentonite
was diffused evenly within the melting urea, its particles were
unrestricted and stable and they deposited in the space of urea
molecules. The more the quantity of bentonite used, the more
compact the lattice structure which decreased the specific surface area and pore size (Ding & Liu, 1998). The release rate could
therefore be controlled to accommodate different requirements.

3.5.

Analysis of slow release kinetics

Figure 6 shows the release kinetics of S-urea (5% bentonite


and 0.15% OP) and the simulation results using the Eq. (1)

279

Fig. 6 e Release kinetics of S-urea. (A) Correlation results of


equation Peppas. (B) Correlation results of doubleexponent equation.

(Fig. 6A) and Eq. (2) (Fig. 6B). The parameters of Eq. (1) were
characterised by the values of k was 7.5939 and n was 0.8687
(with R2 0.9988), indicating that the release of S-urea did not
agree with the Higuchi model. That suggests that the release
of S-urea should not vary with the square root of time, as with
Fickian diffusion, but as with anomalous diffusion (Peppas,
1985; Peppas & Sahlin, 1989). Using Eq. (2) supported this
point and differentiated the mechanism from the Fickian
diffusion and dissolving-erosion diffusion by the values of
k1 0.5650, k2 7.1202 and m 0.4421 (with R2 0.9988). The
ratio k2t2m divided k1tm was 39.52 at 13 h (75% of the total
amount of urea), this showed that the dissolution of the
auxiliary frame eroded by water was the main factor controlling the release (Kaunisto et al., 2011; Peppas & Sahlin,
1989).
The data for all the types of urea shown in Fig. 5A were
correlated using Eq. (1) and Eq. (2). The kinetic parameters are
shown in Table. 1.
The correlation results of Eq. (1) showed that the release
mode of all the types of urea tested was not Fickian diffusion
but anomalous diffusion. It indicated that, even including the
sample of common urea, all of the release process was
controlled not only by the concentration diffusion but also by
some other mechanism that may be associated with the

280

b i o s y s t e m s e n g i n e e r i n g 1 1 5 ( 2 0 1 3 ) 2 7 4 e2 8 2

Table 1 e Kinetic parameters of Fig. 5A using Eq. (1) and


Eq. (2) respectively.
Type of
sample

Urea
P-urea
B-urea
S-urea

Kinetic parameters
for Equation (1)

Kinetic parameters
for Equation (2)

R2

Ratioa

R2

0.8245
0.9104
0.9647
0.8687

0.9472
0.9955
0.9954
0.9988

0.37b
2.62b
8.93c
39.52c

0.9968
0.9988
0.9952
0.9988

a : k2t2m divided k1tm.


b : Time for 90% urea release.
c : Time for 75% urea release.

porous structure of the urea. The results also indicated that


additives may affect the release process of urea by strengthening its original structure.
The cumulative release rate of Fickian diffusion and
dissolving-eroding diffusion were calculated using Eq. (2). The
coefficients of determination (R2) all approached 0.99, showing
that this equation may describe the release process well and
could be used to explain the mechanism. However, the true
process is probably more complex and this does not agree
with the hypothesis referred to by Lee (2011); some results
expressed that the value of the release quantity of urea were
even calculated as being negative. If using the absolute value
to denote the degree of these two release modes, namely the
ratio of k2t2m divided k1tm, it increased regularly according to
the sequence common urea, P-urea, B-urea and S-urea (shown
in Table. 1). This further confirmed the conclusion that the
dissolving-eroding effect becomes more and more preponderant with more compact structures reinforced mainly by
bentonite.
Dealing with the data shown in Fig. 5B in the same way, a
similar result can be found in Table. 2. It also showed that
increasing the proportion of bentonite could strengthen the
structure of urea, at the same time increasing the influence of
dissolving-eroding diffusion. The ratio of S-urea (1%) was less
than P-urea as the structure of the latter was not more
compact, this abnormal result might because that OP diffused
into the water and its chain structure increased the local
viscosity, and this could affect the diffusion process of urea.
On the other hand, a low concentration of bentonite (1%) did

not increase the compact of structure as much as 5%


bentonite did and it can flocculate with OP and decrease the
affect of viscosity (Xiao & Cezar, 2003).

3.6.
Slow-release effect of S-urea under practical
conditions
The result of Fig. 7 shows that during the first 4 days, the
quantity of urea and total N in leaching solution of S-urea
was much less than that of common urea; it was about only
half of the latter. Also the remaining nutrition in soil of Surea provided a steady release velocity during the next 10
days as common urea did. This result shows that S-urea
might have some slow release effect under practical conditions and this affect may last longer in soil than it does in
water. It greatly increases the feasibility of using S-urea in
agricultural production. The reason might be that OP can
connect between the soil particles during the release process
as it did with bentonite, in this way the local circumstances
in the soil around the fertiliser might be changed and it could
affect the diffusion of water and change the process of
nutrition.

Table 2 e Kinetic parameters of Fig. 5B using Eq. (1) and


Eq. (2) respectively.
Type of
sample

Urea
SeU(1%)
SeU(3%)
SeU(5%)

Kinetic parameters
for Equation (1)

Kinetic parameters
for Equation (2)

R2

Ratioa

R2

0.8245
0.9979
0.7548
0.8687

0.9472
0.9928
0.9976
0.9988

0.37b
0.60c
7.41c
39.52c

0.9968
0.9936
0.9976
0.9988

a k2t2m divided k1tm.


b Time for 90% urea release.
c Time for 75% urea release.

Fig. 7 e Slow-release effect of S-urea in soil. (A) The


leaching loss of urea under practical conditions. (B) The
leaching loss of total N under practical conditions.

b i o s y s t e m s e n g i n e e r i n g 1 1 5 ( 2 0 1 3 ) 2 7 4 e2 8 2

4.

Conclusions

A new type of slow-release urea was produced and tested. IR


analysis showed that the two additives used, bentonite and OP,
did not react and no new chemical bond appeared. The additives connected mainly by the Van der Waals force, hydrogen
bond and electrostatic attraction. A lattice structure came into
being after the urea recrystallised at room temperature.
Bentonite accumulated in the space of urea, or inside its crystals, and linked together as the SEM image showed. OP can
strengthen this connection by setting up a bridge between the
bentonite aggregates forming a network. The new type of urea
had a lower specific surface area and its pore size was distributed in a smaller range and its larger stacking density could
increase the path length for water. Kinetic simulation of the
results using the Peppas and the double-exponent equations
showed that the release rate of this type of urea was mainly
affected by dissolving-eroding process which may be controlled
by compactness of the lattice structure and this trend may be
strengthened by increasing the amount of bentonite.

Acknowledgement
We thank Dr. Fanghua Li for his valuable discussions and careful
revisions. This work was supported by the National Agriculture
Transformation Fund of China (No. 2010GB2C300185) and the
Directional Project of Chinese Academy of Sciences.

references

Al-Zahrani, S. M. (1999). Controlled-release of fertilizers:


modelling and simulation. International Journal of Engineering
Science, 37, 1299e1307.
Bhattacharjee, S., & Elimelech, M. (1997). Surface element
integration: a novel technique for evaluation of DLVO
interaction between a particle and a flat plate. Journal of Colloid
and Interface Science, 193, 273e285.
Brunauer, S., Emmett, P. H., & Teller, E. (1938). Adsorption of gases
in multimolecular layers. Journal of the American Chemical
Society, 60, 309e316.
Cai, D. Q., Wu, Z. Y., Jiang, J., Ding, K. J., Tong, L. P., Chu, P. K., et al.
(2009). A unique technology to transform inorganic nanorods
into nano-networks. Nanotechnology, 20, 255e302.
Chen, M., Yang, L. M., Luo, Z. G., & Lu, Q. M. (2002). Preparation
and characterization of polyacrylamide/montmorillonite
intercalated composite. Journal of South China Agricultural
University (Natural Science Edition), 23, 84e86.
Chinese Patent Specification ZL200610040631.1 Loss-control
fertilizer made by active clay, flocculant and sorbent.
Chinese Patent Specification CN201110003090.6 A new technique
used in producing a novel slow-release urea with a network
inside and its application.
Diez, J. A., Roman, R., Cartagena, M. C., Vallejo, A., Bustos, A., &
Caballero, R. (1994). Controlling nitrate pollution of aquifers by
using different nitrogenous controlled release fertilizers in
maize crop. Agriculture, Ecosystems and Environment, 48, 49e56.
Ding, S. L., & Liu, Q. F. (1998). Experimental study on using
montmorillonite as slow releasing matrix for urea. ACTA
Mineralogica Sinica, 18, 67e72.

281

Fan, X. L., & Liao, Z. W. (1998). Increasing fertilizer use efficiency


by means of controlled release fertilizer (CRF) production
according to theory and techniques of balanced fertilization.
Plant Nutrition and Fertilizer Science, 4, 219e223.
Govind, C., & Sharma. (1979). Controlled-release fertilizers and
horticultural applications. Scientia Horticulturae, 11,
107e129.
Han, X. Z., Chen, S. S., & Hu, X. G. (2009). Controlled-release
fertilizer encapsulated by starch/polyvinyl alcohol coating.
Desalination, 240, 21e26.
He, X. S., Liao, Z. W., Huang, P. Z., Duan, J. X., Ge, R. S., Li, H. B.,
et al. (2007). Characteristics and performance of novel waterabsorbent slow release nitrogen fertilizers. Agriculture Science
in China, 6, 338e346.
Higuchi, T. (1963). Mechanism of sustained-action medicationtheoretical analysis of rate of release of solid drugs dispersed
in solid matrices. Journal of Pharmaceutical Sciences, 52,
1145e1149.
James, A. E., & Sojka, R. E. (2008). Matrix based fertilizers reduce
nitrogen and phosphorus leaching in three soils. Journal of
Environmental Management, 87, 364e372.
Jiang, J. Y., Hu, Z. H., Sun, W. J., & Huang, Y. (2010). Nitrous oxide
emissions from Chinese cropland fertilized with a range of
slow-release nitrogen compounds. Agriculture, Ecosystems and
Environment, 135, 216e225.
Kaunisto, E., Marucci, M., Borgquist, P., & Axelsson, A. (2011).
Mechanistic modelling of drug release from polymer-coated
and swelling and dissolving polymer matrix systems.
International Journal of Pharmaceutics, 418, 54e77.
Lee, P. I. (2011). Modeling of drug release from matrix systems
involving moving boundaries: approximate analytical
solutions. International Journal of Pharmaceutics, 418, 18e27.
Lenaerts, V., Dumoulin, Y., & Mateescu, M. A. (1991). Controlled
release of theophylline from cross-linked amylose tablets.
Journal of Controlled Release, 15, 39e46.
Li, S., Sun, Y. B., Feng, J. W., Tian, Y. C., Yang, H. F., & Xu, Y. W.
(2007). Research on adsorption of high concentration PAM
solution to natural Na-bentonnite. Chinese Journal of
Environmental Engineering, 1, 47e50.
Li, Z. F., & Yang, G. S. (2004). Research on non-point source
pollution in Taihu Lake region. Journal of Lake Sciences,
16(Suppl.), 83e88.
Liu, Y. H., Wang, T. J., Qin, L., & Jin, Y. (2008). Urea particle coating
for controlled release by using DCPD modified sulfur. Powder
Technology, 183, 88e93.
Ni, B. L., Liu, M. Z., & Lue, S. Y. (2009). Multifunctional slowrelease urea fertilizer from ethylcellulose and superabsorbent
coated formulations. Chemical Engineering Journal, 155,
892e898.
Peppas, N. A. (1985). Analysis of fickian and non-fickian drug
release from polymers. Pharmaceutica Acta Helvetiae, 60,
110e111.
Peppas, N. A., & Sahlin, J. J. (1989). A simple equation for the
description of solute release III. Coupling of diffusion and
relaxation. International Journal of Pharmaceutics, 57, 169e172.
Prasad, R., Rajale, G., & Lacakhdive, B. (1971). Nitrification
retarders and slow-release nitrogen fertilizers. Advances in
Agronomy, 23, 337e383.
Salman, O. A. (1989). Polyethylene-coated urea: 1. Improved
storage and handling properties. Industrial & Engineering
Chemistry Research, 28, 630e632.
Slade, P. G., Quirk, J. P., & Norrish, K. (1991). Crystalline swelling of
smectite samples in concentrated NaCl solution in relation to
layer charge. Clays and Clay Minerals, 39, 234e238.
Subrahmanyan, K., & Dixit, L. A. (1988). Effect of different coating
materials on the pattern of phosphorous release from
superphosphate. Journal of the Indian Society of Soil Science, 36,
461e465.

282

b i o s y s t e m s e n g i n e e r i n g 1 1 5 ( 2 0 1 3 ) 2 7 4 e2 8 2

Trenkel, M. E. (1997). Controlled-release and stabilized fertilizers in


agriculture (pp. 53e102). Paris: France International Fertilizer
Industry Association.
Xiao, H. N., & Cezar, N. (2003). Organo-modified cationic silica
nanoparticles/anionic polymer as flocculants. Colloid and
Interface Science, 267, 343e351.
Xie, L. H., Liu, M. Z., Ni, B. L., Zhang, X., & Wang, Y. F. (2011).
Slow-release nitrogen and boron fertilizer from a
functional superabsorbent formulation based on wheat

straw and attapulgite. Chemical Engineering Journal, 167,


342e348.
Zheng, T., Liang, Y. H., Ye, S. H., & He, Z. Y. (2009). Superabsorbent
hydrogels as carriers for the controlled-release of urea:
experiments and a mathematical model describing the release
rate. Biosystems Engineering, 102, 44e50.
Zheng, Y. Y., Zhang, H. H., Cai, W. B., Fu, M. L., & Wang, L. E.
(2005). Preparation and properties of organobentonites.
Spectroscopy and Spectral Analysis, 25, 62e64.

You might also like