Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

This article was downloaded by: [University of Auckland Library]

On: 09 October 2014, At: 15:20


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer House,
37-41 Mortimer Street, London W1T 3JH, UK

Journal of Dispersion Science and Technology


Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/ldis20

Synthesis and Surface Active Properties of Ethoxyalted


4,5-Dihydroimidazoline Surfactants
a a b a
A. M. Al-Sabagh , R. A. El-Ghazawy , H. M. Abdel Bary & M. Abd Raouf
a
Egyptian Petroleum Research Institute , Cairo, Egypt
b
Chemistry Department , Faculty of Science, Al Azhar University , Cairo, Egypt
Published online: 09 Dec 2009.

To cite this article: A. M. Al-Sabagh , R. A. El-Ghazawy , H. M. Abdel Bary & M. Abd Raouf (2009) Synthesis and Surface Active
Properties of Ethoxyalted 4,5-Dihydroimidazoline Surfactants, Journal of Dispersion Science and Technology, 31:1, 75-83, DOI:
10.1080/01932690903107232

To link to this article: http://dx.doi.org/10.1080/01932690903107232

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information (the “Content”) contained
in the publications on our platform. However, Taylor & Francis, our agents, and our licensors make no
representations or warranties whatsoever as to the accuracy, completeness, or suitability for any purpose of the
Content. Any opinions and views expressed in this publication are the opinions and views of the authors, and
are not the views of or endorsed by Taylor & Francis. The accuracy of the Content should not be relied upon and
should be independently verified with primary sources of information. Taylor and Francis shall not be liable for
any losses, actions, claims, proceedings, demands, costs, expenses, damages, and other liabilities whatsoever
or howsoever caused arising directly or indirectly in connection with, in relation to or arising out of the use of
the Content.
This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any
form to anyone is expressly forbidden. Terms & Conditions of access and use can be found at http://
www.tandfonline.com/page/terms-and-conditions
Journal of Dispersion Science and Technology, 31:75–83, 2010
Copyright # Taylor & Francis Group, LLC
ISSN: 0193-2691 print=1532-2351 online
DOI: 10.1080/01932690903107232

Synthesis and Surface Active Properties of Ethoxyalted


4,5-Dihydroimidazoline Surfactants
A. M. Al-Sabagh,1 R. A. El-Ghazawy,1 H. M. Abdel Bary and
M. Abd Raouf1
1
Egyptian Petroleum Research Institute, Cairo, Egypt
2
Chemistry Department, Faculty of Science, Al Azhar University, Cairo, Egypt

Pure oleic acid and four hydrolyzed products of locally fatty oils (namely, coconut, soybean,
linseed, and castor oils) were monoesterified individually with two different polyalkylene poly-
amines. The produced monoesters were used as precursors for imidazoline derivatives. These
Downloaded by [University of Auckland Library] at 15:20 09 October 2014

oil soluble imidazolines were ethoxylated by ethylene oxide gas with different ethylene oxide
contents (5, 10, and 20) to prepare water soluble stable imidazoline ethoxylates. The chemical
structure of the synthesized imidazoline ethoxylates were justified through spectroscopy and
surface active properties of the compounds were investigated. The obtained data were discussed
on the basis of variation chemical structure.
Keywords 4.5-dihydroimidazoline, ethoxylated imidazoline, hydrolyzed fatty acid, imidazo-
line surfactants, surfactants

INTRODUCTION emulsifiers (surfactants), viscosity modifiers, and suspend-


Imidazoline was first synthesized in 1858[1] from ammo- ing agents. The term ‘‘imidazolines’’ for these applications
nia and glyoxal whereas more recent syntheses have been is used to describe products that contain mixture of ami-
developed from para aldehyde or 2-bromoethyl-1,3- doamine and imidazoline.[8] Quaternized imidazolines with
dioxalane, which, on heating with formamide and an amido moiety are suitable formulations for general oil
ammonia, yields imidazoline. Desulphurization of thione and gas field applications.
intermediate with nitric acid yields imidazoline. Imidazo- The main object of this work is to synthesize water soluble
line has a pKa of 7.2 and a pKb of 14.5, and can function ethoxylated 4,5-dihydroimidazoline surfactans. These
as both weak acid and weak base. Imidazoline forms many products were synthesized by reacting the hydrolyzed pro-
stable salts and is thought to have intermediate properties ducts of some fatty oils with diethylene triamine (DETA)
between pyrrole and pyridine. Also, imidazoline forms or triethylene tetramine (TETA) to prepare the monoamide
strong intermolecular hydrogen bonds and associates of form. The monoamide were cyclized to obtain the corre-
20 to 50 molecules in benzene solution can be formed sponding 4,5-dihydroimidazoline derivatives. The imidazo-
giving linear hydrogen-bonded chains.[2] The applications line products were ethoxylated using ethylene oxide gas to
of imidazoline derivatives are widely used in personal care get stable ethoxylated derivatives (surfactants). Our attention
industry, antifungal formulations, and in petroleum indus- was extended to investigate their surface active properties.
try as corrosion inhibitors.
The imidazoline tends to be unstable and care has to be
EXPERIMENTAL
exercised during manufacture and storage to minimize
color degradation and hydrolysis. For this reason, imida- Hydrolysis of Fatty Oils
zoline is frequently derivitized by quaternization or carbox- Four different fatty oils (namely, coconut, soybean,
ymethylation.[3] Imidazoline provides excellent corrosion linseed, and castor oils) were separately hydrolyzed to obtain
inhibition and is widely used in the petroleum industry fatty acid mixture from each. Into a (500-ml) two neck
for corrosion protection in production and refining appli- round bottom flask attached to a reflux condenser and a
cations.[4–7] Other oilfield applications include its use as mechanical stirrer, add 10 g of fatty oil, 5 g of KOH, and
40 ml of ethanol. The reaction mixture was refluxed until
Received 6 August 2008; accepted 15 September 2008. no oil globules were observed. The reaction mixture was
Address correspondence to A. M. Al-Sabagh, Egyptian distilled to recover the alcohol. The residue was dissolved
Petroleum Research Institute, 1 Ahmed El-Zomor St., Nasr City, in (150 ml) hot water and stirred for (90 minutes) until the
11727, Cairo, Egypt. E-mail: alsabaghh@gmail.com mixture became almost clear. A diluted solution of sulfuric

75
76 A. M. AL-SABAGH ET AL.

TABLE 1 The general chemical structures, IUPAC (International


Percentage of fatty acids in hydrolyzed fatty oils[2] Union of Pure and Applied Chemistry) name, and abbre-
viation of representative compounds (amide, imidazoline,
Fatty acids % and ethoxylate) are shown in Table 2. Fourier transform
Oil Saturated Oleic Linoleic Linolenic Ricinoleic infrared spectroscopy (FTIR) was used to confirm their
chemical structures (Figures 1 and 2).
Coconut 91 7 2 — —
Soybean 15 25 51 9 — Evaluation of Some Surface Properties of the
Linseed 10 22 16 52 — Prepared Surfactant
Castor Oil 3 7 3 — 87 Surface Tension Measurements (c)
Different molar concentrations of the synthesized sur-
factants were prepared using double distilled water and
acid (300 ml, 98% concentrated in 700 ml water) was their surface tension were determined at 25 C using
added with continuous stirring for (5 minutes). The oily Lecomte De Nouy tentiometer with platinum ring-iridium,
layer was separated and filtrated in steam-jacketed funnel Krüss-K6 (Germany). The instrument was daily regulated
and finally allowed to cool until solidify. The hydrolyzed by bi-distilled water (conductivity 1.1  106 ohm1 cm1
Downloaded by [University of Auckland Library] at 15:20 09 October 2014

products of fatty oils contained mixtures of fatty acids[9] at 25 C).[15]


(Table 1).
Critical Micelle Concentration
The critical micelle concentration (CMC) of the pre-
Synthesis of the 4,5-Dihydro Imidazoline Derivatives
pared surfactants was determined by the method adopted
In this step, two series of 4,5-dihydro imidazoline surfac-
by Larinov et al.[16] The surface tension –ln molar con-
tants were prepared. Each hydrolyzed fatty oil and oleic
centration isotherms (SVC) were plotted for the prepared
acid was separately condensed with either DETA or
surfactants. The CMC values were determined from the
TETA. The first series (series I) was based on DETA,
abrupt change in the slope of SVC curves.[17] The SVC
and the second one (series II) was based on TETA. First,
curves for the ethoxylated samples of Series II surfactants
the reaction was carried out at 140 C for 1 hour using a
are represented in Figure 3.
Dean and Stark trap to produce the corresponding amides
(see Scheme 1). Then, continuous heating in the range of Surface Excess Concentration (CMax)
180 to 200 C for 3.5 hours led to cyclization of each amide
CMax is a measure of the effectiveness of adsorption of
to the corresponding 4,5-dihydro imidazolines. This is
surfactant at the liquid=air or liquid=liquid interface since
represented by the second step in Scheme 1.[10]
it is the maximum value to which adsorption can be
obtained.[18,19]
Synthesis of the Ethoxylated Amino Alkyl Imidazolines CMax can be calculated from Gibbs equation:
A high pressure stainless steel autoclave (Parr model
4530, USA) of 1-L capacity, 400 psi maximum pressure Cmax ¼ ð1=RTÞ ðdc=d ln CÞ;
and 450 C maximum temperatures was utilized through
ethoxylation reactions. The autoclave was equipped with where: CMax is surface excess concentration (mol=cm2), T is
a magnetic drive stirrer, an electric heating mantle with absolute temperature (273 þ  C), R is universal gas
thermocouple inserted in the reactor body, a cooling coil, constant (R ¼ 8.314 Jmol1 deg1), and dc=d Ln C is
a pressure gauge, and a drain valve. Each amino alkyl imi- surface activity (slope of linear portion of SVC curves).
dazoline (synthesized in the second step) was charged into
the reaction vessel with triethyl amine as a catalyst Minimum Surface Area Per Molecule (Amin)
(0.3 wt%)[11] and heated to 80 C with continuous stirring Amin is the minimum area per molecule in nm2=molecule
while passing a stream of nitrogen gas through the system at the interface.[20] The area occupied by each adsorbed
for 10 minutes to flush out air. The nitrogen stream was molecule is given by equation:
then replaced by ethylene oxide, the first series based on
Amin ¼ 1016 =ðNA  CMax Þ;
diethylene triamine was ethoxylated by only one amount
of EO, namely, n ¼ 20. For the second series based on where: NA is Avogadro’s number.
triethylene tetramine, different EO molar ratios were intro-
duced, namely, n ¼ 5, 10, or 20 EO moles. The difference in Effectiveness pCMC
sample weight was used as an indication of the amount of The effectiveness of adsorption or surface pressure pCMC
condensed EO. The ethoxylation was represented by the of the surfactants was also calculated using the following
third step in Scheme 1.[12–14] equation.[21]
ETHOXYALTED 4,5-DIHYDROIMIDAZOLINE SURFACTANTS 77
Downloaded by [University of Auckland Library] at 15:20 09 October 2014

SCH. 1. Synthesis of 4,5-dihydro imidazoline derivatives.

pCMC ¼ c0  cCMC ; 1669.45 cm1 together with a broad band in the range of
3320 to 3559 cm1 region for N-H stretching. The comple-
where: c0 is the surface tension of water at 25 C; and cCMC tion of imidazole derivative formation (I3Z) was confirmed
is the surface tension value of the measured solution at by the sharp absorption at 1625.7 cm1 (Figure 1b) corre-
CMC. sponding to C=N stretching of the imidazoline ring.
However, the characteristic absorption of the iminic
RESULTS AND DISCUSSION C=N group for imidazoline ring in the usual range
(1603–1635 cm1) rather than in the region of amidic
Spectroscopic Analysis of the Prepared Surfactants carbonyl (1640–1670 cm1) confirms the cyclization of the
FTIR of I2, I2Z, and I2ZE20 are illustrated in amides upon continuous heating.[22]
Figures 1a–1c, respectively. Figure 1a, for I3, is a typical Figure 1c, for I3ZE20, is a sharp band at 1119.84 cm1
amide carbonyl absorption band was observed at for CO stretching vibration of ethylene oxide moiety,
78 A. M. AL-SABAGH ET AL.

TABLE 2
IUPAC name and abbreviation of some representative compounds
Product
Structures and IUPAC name code Product nature

CH3 ---ðCH2 Þ7 ---CH ¼ CH---ðCH2 Þ7 CO---NH---CH2 --- I2 Amide


N-diethylene
CH2 ---NH---CH2 ---CH2 ---NH2
triamine cis-9-octadecenamide (main amide derived from pure oleic
acid and DETA)
CH3 ---ðCH2 Þ7 ---CH ¼ CH---ðCH2 Þ7 CO---
N-triethylene II2 Amide
NH---ðCH2 ---CH2 ---NHÞ2 ---CH2 ---CH2 ---NH2
tetraamine cis-9-octadecenamide (main amide derived from pure
oleic acid and TETA)
I 2Z Imidazoline
Downloaded by [University of Auckland Library] at 15:20 09 October 2014

2-(heptadeca-8-enyl)-4,5-dihydroimidazole-1-ethylamine
II2Z Imidazoline

2-(heptadeca-8-enyl)-4,5-dihydroimidazoline-1-diethyle-nediamine
I2ZE20 Ethoxylated imidazoline
(surfactant) EO ¼ 20

2-(heptadeca-8-enyl)-4,5-dihydroimidazoline-1-ethyl–N,N-dipolyoxyethylene
amine
II2ZE5 Ethoxylated imidazoline
(surfactant) EO ¼ 5
II2ZE10 Ethoxylated imidazoline
(surfactant) EO ¼ 10
II1ZE20 Ethoxylated imidazoline
2-(heptadeca-8-enyl)-4,5-dihydroimidazoline-1- (surfactant) EO ¼ 20
(N,N,N-tripolyoxyethylene)diethylene diamine
Where n ¼ x þ y, n ¼ x þ y þ z (n ¼ number of ethylene oxide units 5, 10, or 20).

an intense broad band at 3120–3610 cm1 for ethylene Further elucidation for the chemical structure of the
oxide terminal hydroxyl group and an amide carbonyl prepared compounds was performed using 1H NMR spec-
absorption at 1624 cm1 were observed. troscopy. 1H NMR spectra for I3, I3Z and I2ZE20 are
shown in Figures 2a–2c.
1
H NMR Spectroscopic Analysis Figure 2a for I2 shows two overlapped signals at d ¼ 5.15
It is worth noting that all the prepared compounds even and 5.33 ppm for the amidic proton and the double bond
those depending on oleic acid are mainly based on com- protons, respectively. In Figure 2b for I2Z, two overlapped
mercial fatty acid products. In other words, their analysis triplets appeared in the range of d ¼ 3.74–3.86 ppm
is quite difficult due to the presence of the multi- corresponding to both methylenes of the imidazoline ring.
component systems and other impurities. Besides, the intensity of the signal appeared at d ¼ 3.14 ppm
ETHOXYALTED 4,5-DIHYDROIMIDAZOLINE SURFACTANTS 79

interface and micelles in the bulk of the continuous phase


is a phenomenon playing an important role through
studying the corrosion inhibition capacity of inhibitors.
Thus, the determination of some thermodynamic para-
meters such as Gibbs free energy of both micellization
and adsorption will give an idea of their performance.

Critical Micelle Concentration


Typical curves of surface tension (c) versus ln the surfac-
tant molar concentration (lnC) for the prepared surfactants
were obtained (Figure 3). At CMC the adsorbed surfactant
molecules perform a monolayer covering the whole sur-
face[23] and meanwhile, the surfactant molecules aggregate
into micelles with their lyophilic groups directed toward the
interior of the micelles and their hydrophilic groups direc-
ted towards the solvent (water). Micellization is, therefore,
Downloaded by [University of Auckland Library] at 15:20 09 October 2014

a mechanism alternative to adsorption at the interface for


removing lyophobic groups from contact with solvent,
thereby reducing the free energy of the system.
The exact structure of the micelles is still disputed.
However, Hano et al.[24] referred the structure of the
micelle to be roughly spherical with an interior region con-
taining the hydrophobic groups of the surfactant mole-
cules, of radius approximately equals to the length of a
fully extended hydrophobic groups, surrounded by an
outer region containing the hydrated hydrophilic groups
and bound water. However, CMC values are affected
markedly in aqueous solution by the structure of the sur-
factant. Table 3 shows the variation of CMC as a function
of the number of EO units for the second series II. It can be
seen that, the CMC values decrease as the hydrophilic
character of the surfactant increases (i.e., increase in EO
units number). As an example the CMC values, for II2ZE5,
II2ZE10, and II2ZE20 are 1.5  104, 0.82  104, and
FIG. 1. Fingerprint for FTIR spectrum of a) amide of I2, b) imidazo- 0.67  104, respectively. This may be due to the increase
line of I2Z, and c) ethoxylated I2Z 20. of oxyethylene units leads to an increase in the total mole-
cular weight of the surfactant and consequently the area
occupied by the molecule at the aqueous=air interface
decreases due to the incorporation of a further amine increases. Hence, the number of molecules required to
group. Figure 2c for I2ZE20, shows a highly intense signal cause interface saturations is decreased and the process of
in the range of d ¼ 3.44–3.56 ppm attributed to the micellization commences earlier compared to the surfac-
methylene groups of ethylene oxide moiety and the imida- tant having lower oxyethylene content. This speculation
zoline ring appeared in the same range of d ¼ 3.7–3.9 ppm. is in accordance with a previous explanation.[25] Also, it
The FTIR and 1HNMR justified the structures of amide, was suggested that[26,27] the geminal poly oxyethylene
imidazoline and ethoxylate of prepared surfactants. chains linked to the same atom (nitrogen atom in our
study) never individually hold a random coil configuration,
Surface Active and Thermodynamic Properties of the but intertwist with each other and penetrate the aqueous
Prepared Surfactants solution. Therefore, the CMC values decrease with increas-
The performance of surfactants as corrosion inhibitors ing the number of oxyethylene units, which is opposite to
is assessed by their surface activity which is largely related the usual observation in single chain poly oxyethylenated
to their adsorption properties at interface and other surfactants. On the other hand, the surface activity and
parameters such as CMC, cCMC, Cmax and Amin. Also, the structure of the adsorbed films are greatly influenced
the existing equilibrium between the monomeric free by the chemical structure of the hydrophobic groups of
surfactant molecules diffusing or getting adsorbed on the the surfactants namely, different chain length and=or
80 A. M. AL-SABAGH ET AL.
Downloaded by [University of Auckland Library] at 15:20 09 October 2014

FIG. 2. Fingerprint for 1H NMR spectra of a) amide of I2, b) imidazoline of I2Z, and c) ethoxylated I2Z 20.

branching. Through this study, the hydrophobic chain hydrophobic group as a main component yields higher
length varied slightly being 18 carbon atoms as obvious CMC values compared with those having unsaturations.
from fatty acid composition of the selected oils. However, The CMC value for I1ZE20 is 3  104 (mol  dm3) while
Table 3 shows that the surfactants containing saturated those for I3ZE20, I4ZE20, and I5ZE20 are in the range of
ETHOXYALTED 4,5-DIHYDROIMIDAZOLINE SURFACTANTS 81

In case of the surfactants based on ricinoleic acid which


obtained from hydrolysis of castor oil (viz., I5ZE20,
II5ZE10, and II5ZE20), the presence of polar hydroxyl
branches in the hydrophobic moiety may oppose the pre-
viously mentioned repulsion force. Thus, based on the
same rationalization, CMC will again increases. This is
obvious from CMC values for I5 ZE20, II5 ZE5, II5 ZE10,
and II5 ZE20 listed in Table 3 based on ricinoleic acid as
a main hydrophobic component when compared with
other surfactants containing only one, two, or three unsa-
turation bonds through the hydrophobic moiety.
This feature is adopted through the values of effective-
ness for the surface-tension reduction (IICMC) tabulated
in Table 3. It can be observed that IICMC values are higher
for oleic acid based surfactants than others.
Maximum surface excess concentration Cmax in mol=
Downloaded by [University of Auckland Library] at 15:20 09 October 2014

cm2 calculated from Gibbs equation (above in the Experi-


mental section) is tabulated in Table 3. As a general trend,
Cmax decreases by increasing the number of EO units incor-
porated in the surfactant structure (as shown by the second
group surfactants). This may be due to the increase of the
total molecular weight of the surfactant as a result of the
increase in the number of EO units. Consequently, the area
occupied by the surfactant molecule at the water=air inter-
face increases. Hence, the number of molecules required to
cause interface saturation decreases, that is, a decrease in
the maximum surface excess concentration. This specula-
tion is in accordance with a previous explanation.[25]
The effect of increased number of double bonds in the
hydrophobic part has not been widely investigated. Here,
regarding Cmax values for I3ZE20, I4ZE20, and I5ZE20, it
is found that Cmax decreases with increasing the number
of double bonds. This is mainly controlled by geometrical
FIG. 3. c-ln C adsorption isotherm for a) II2ZE5, b) II2ZE10, and
packing considerations of the surfactant molecules. The
c) II2ZE20. previously stated increased repulsion between hydrophobic
chains with increased unsaturation suggests less molecular
packing. Hence, a decreased Cmax is logically expected by
1.3  104 to 1.5  104 (mol  dm3). This may be attributed increasing the number of double bonds from one through
to relatively higher electrostatic repulsion between the three double bonds.
hydrophobic chains containing unsaturation than in case Also, the area per molecule (Amin) at the interface can be
of saturated ones. This higher repulsion provides a force calculated from the Cmax values. The packing geometry,
tending to increase the area per molecule which in turn controlled by relative cross-sectional areas of hydrophobic
decreases the concentration required to saturate water=air chain and head groups, plays a vital part in (Amin) values.
interface. Thus, micellization in bulk starts earlier in case From the data presented in Table 3, Amin values increase
of unsaturated hydrophobes, that is, lower CMC values. with increasing the number of EO units incorporated. This
Regarding the surface tension at CMC (cCMC) listed in may be attributed to the strong affinity of the hydrophilic
Table 3, it is noticeable that their values are unaffected groups for water that there is a tendency for them to be
by the change in poly oxyethylene chain length (as obvious spaced out to allow as much water as possible to solvate
from the values of (cCMC) for the second group surfac- each head group. In other words, hydration tends to
tants). On the other hand, the hydrophobic part affects increase the area. This feature is also adopted in CMC
(cCMC) being the lowest for the surfactants based on oleic values, where CMC values decrease with increasing EO
acid (viz., I2ZE20, II2ZE5, II2ZE10, and II2ZE20). This units number incorporated. On the other hand, the increase
may be attributed to more optimum surfactant-water inter- in the number of double bonds through the hydrophobic
action compared with surfactant-surfactant interaction. chains tends to cause an increase in Amin. This behavior
82 A. M. AL-SABAGH ET AL.

TABLE 3
Surface properties of the prepared surfactants
Surfactant CMC  104 cmcP max C  1010 Amin
1
code cmcc (mNm ) (mol  dm3) 1
(mNm ) (mol  cm2) (nm2 molecule1)

I1ZE20 39 3.0 33 2.48 69.1


I2ZE20 39 3.3 33 2.03 81.7
I3ZE20 34 1.5 38 2.40 66.9
I4ZE20 40 1.5 32 2.31 71.8
I5ZE20 40 1.3 32 2.00 83.0
II2ZE5 34 1.5 38 2.85 58.2
II2ZE10 34 0.82 38 2.83 58.6
II2ZE20 33 0.67 39 2.81 59.0
II3ZE5 40 4.5 32 2.49 60.6
II3ZE10 39 2.7 33 2.47 67.2
II3ZE20 39 2.7 33 2.45 67.7
Downloaded by [University of Auckland Library] at 15:20 09 October 2014

II5ZE5 39 1.6 33 2.93 56.6


II5ZE10 39 1.2 33 2.79 59.4
II5ZE20 39 0.6 33 2.45 67.7

may be a result of increased electrostatic repulsion with through all the prepared surfactants. This geminal struc-
increased number of double bonds. ture is considered to be unfavorable to hydration.[32]
The free energy of adsorption DGad for the selected
Some Thermodynamic Functions of the Surfactants surfactants was calculated via the following equation:[35]
Standard free energy of micellization determines the
spontaneity of micellization. Direct measurements of the DGad ¼ DGmic  0:6023 PCMC Amin
excess free energy of micellar systems,[28–30] through
activity of a solute or a solvent are difficult to be measured The obtained values are listed in Table 4. Analyzing the
precisely[31] at concentrations below 0.01 mol=kg. Conse- obtained data, it was found that all DGad values are nega-
quently, most information on the standard free energy of tive and they are more negative than DGmic values. This
micellization DGmic has been obtained indirectly through
CMC determination.[32,33] DGmic may be calculated by
choosing the following expression: TABLE 4
Thermodynamic parameters of adsorption for the prepared
DGmic ¼ RTð1 þ aÞ ln CMC surfactants

where: R is the universal gas constant (¼8.314 J=mol  k; Surfactant DGmic DGad DGad  DGmic.
T is the absolute temperature (¼t þ 273)  K; and a is the code (KJ mol1) (KJ mol1) (KJ mol1)
fraction of the counter ions bound by the micelle in case
I1ZE20 20.06 21.38 1.32
of ionic surfactant (a ¼ 0 for nonionic surfactants).
I2ZE20 19.82 21.4 1.62
This expression uses a hypothetical standard initial state
I3ZE20 21.80 23.3 1.58
for nonmicellar surfactant species at unit mole fraction x—
I4ZE20 21.8 23.1 1.38
with ions or molecules behaving as at infinite dilution and
I5ZE20 22.02 23.6 1.59
a standard final state (the micelle itself). Analyzing the free
II2ZE5 21.8 23.1 1.33
energy of micellization data (tabulated in Table 4), one can
II2ZE10 23.2 24.5 1.34
get that micellization process is spontaneous (DGmic < 0).
II2ZE20 23.7 25.0 1.38
The data shows also that the negativity of DGmic
II3ZE5 19.07 20.35 1.28
increases with increasing the number of EO units incorpo-
II3ZE10 20.31 21.6 1.33
rated. This behavior indicates that increased oxyethylene
II3ZE20 20.81 22.1 1.34
chain length favors micellization process. This observation
II5ZE5 21.55 22.67 1.12
is opposite to the usual observed increase in DGmic for sin-
II5ZE10 22.3 23.4 1.18
gle chain poly oxyethylene surfactants.[34] This feature is
II5ZE20 24.04 25.3 1.34
attributed to the geminal poly oxyethylene chains found
ETHOXYALTED 4,5-DIHYDROIMIDAZOLINE SURFACTANTS 83

indicates that the adsorption at the interface is associated [15] Mohamed, R.S. (1999) M.Sc. Thesis, Cairo, Egypt: Alexandria
with a decrease in free energy of the system, that is, adsorp- University.
tion process is more spontaneous. Also, the more negative [16] Larinov, L.I.O. (1962) Surface Chemistry; New York:
values of DGad indicate favorable adsorption to micelliza- Reinhold.
[17] Al-Sabagh, A.M., Hamed, M.M., Badawi, A.M., and
tion. This may be due to the effect of steric factor on inhi-
Abdel-Moneem, N.E. (1996) Bull. Fac. Sci. Zagazig Univ.,
biting of micellization is more than its effect on adsorption.
18(2): 57.
The product of structural effect DGmic–DGad for the [18] Bhattachargya, D.N., Kelkar, R.Y., Al-Meida, M.R., Das,
selected surfactants are shown in Table 4. This positive A.K., and Chikhale, S.V. (1994) Tenside Surf. Det., 31 (4): 260.
product values reflect that these surfactants are more read- [19] Rosen, M.J. (1978) Surfactants and Interfacial Phenomena;
ily adsorbed at air=aqueous solution interface.[36] This in New York: John Wiley and Sons.
turn could account for investigating these surfactants as [20] Al-Sabagh, A.M., Kandil, N.G., Badawi, A.M., and
corrosion inhibitors in a subsequent article. El-Sharkawy, H. (2000) Colloids Surf. A, 170: 127.
[21] ASTM G31-72 (1990) Standard Practice for Laboratory
Immersion Testing of Metal.
REFERENCES [22] Beintiss, F., Lagrenee, M., and Traisnel, M. (2000) Corrosion,
[1] Debus, H. (1858) Ann., 107: 204. 56: 733.
Downloaded by [University of Auckland Library] at 15:20 09 October 2014

[2] Lawrence, S.A. (2004) Amines: Synthesis, properties and [23] Olivares-Xometl, O., Likhanova, N.V., Dominguez-Aguilr,
applications; Cambridge, UK: Cambridge University Press. M.A., Hallen, J.M., Zamudio, L.S., and Arce, E. (2005)
[3] Richardson, F.B. (1992) Industrial Applications of Surfac- Appl. Interface Sci., 252: 6.
tants; Cambridge, UK: Royal Society of Chemistry; [24] Preston, W.C. (1948) J. Phys. Chem., 52: 84.
pp. 161–183. [25] Hano, T., Ohtake, T., and Takagi, K. (1988) J. Chem. Eng.,
[4] Cruz, J., Martinez, R., Genesca, J., and Garcia-Ochoa, E. 21: 345.
(2004) J. Electro Anal. Chem., 566 (1): 111. [26] Mukerjee, P. (1967) Adv. Colloid Interf. Sci., 1: 241.
[5] Wang, D., Shuyuan, Li, Ying, Y., Wang, M., Xiao, H., and [27] Kuwamura, T. and Takahashi, H. (1972) Bull. Chem. Soc.
Chen, Z. (1999) Corrosion Sci., 41 (10): 1911. (Jpn.), 45: 617.
[6] Oppenlander, K., Oetter, G., Kroner, R., Mahr, N., Jatzek, [28] Esumi, K. and Ueno, M. (1997) Structure-Performance Rela-
N.J., and Taeger, K. (2000) US Pat. No. 6077460. tionships in Surfactants–Surfactant Science Series; New York:
[7] Eaton, P. and Kanwar, S. (1999) US Pat. No. 5961885. Marcel Dekker.
[8] Fishers, E.R. and Boyd, P.G. (1998) US Pat. No. 5759485. [29] Coudert, R., Poris, J., and Cao, A. (1994) J. Colloids Inter-
[9] Vogel, A.I. (1974) A Textbook of Practical Organic Chemis- face Sci., 163: 94.
try; London: Longman. [30] Sesta, B. (1989) J. Phys. Chem., 93: 7677.
[10] Wicks, Z.W., Jones, F.N., and Pappas, S.P. (1992) In [31] Sesta, B. and La-Mesa, C. (1992) Colloid Poly. Sci., 270: 912.
Organic Coatings: Science and Technology; New York: [32] Oda, H., Nagadome, S., Lee, S., Ohseto, F., Sasaki, Y., and
Wiley-Interscience, vol. 1. Sugihora, G. (1997) J. Oil Chem. Soc. (Jpn.), 46: 559.
[11] Ohshiro, Y., Ochiai, M., and Komori, S. (1961) Kogyo [33] Schramm, L.L. (2000) Surfactants: Fundamentals and
Kagaku Zasshi, 64: 23. Applications in the Petroleum Industry; Cambridge, UK:
[12] Hrec, Z.W. and Kozlek, K. (1996) Tenside Surf, Deterg., Cambridge University Press.
38(2): 621. [34] Molymeux, P., Rhodes, C.T., and Swarbrick, J. (1965) Trans
[13] Os, N.M.V. (1998) Nonionic Surfactants–Surfactant Science Faraday Soc., 61: 1043.
Series; New York: Marcel Dekker, vol. 72. [35] Rosen, M.J. (1989) Surfactants and Interfacial Phenomenon,
[14] Lang, R.F., Diaz, P.D., and Jocobs, D. (1999) J. Surf. Det., 2nd ed.; New York: John Wiley.
7 (4): 503. [36] Rosen, J. and Aronson, S. (1981) Colloids Surf., 3: 201.

You might also like