Incorporation of nano-clay saponite layers in the organo-clay hybrid films using anionic amphiphile...

8
Incorporation of nano-clay saponite layers in the organo-clay hybrid lms using anionic amphiphile stearic acid by LangmuirBlodgett technique Syed Arshad Hussain a, , S. Chakraborty a , D. Bhattacharjee a , R.A. Schoonheydt b a Department of Physics, Tripura University, Suryamaninagar-799022, India b Centres for Surface Chemistry and Catalysis, K.U. Leuven, Kasteelpark Arenberg 23, 3001 Leuven, Belgium abstract article info Article history: Received 23 March 2012 Received in revised form 5 March 2013 Accepted 6 March 2013 Available online xxxx Keywords: Pressure-area isotherm Langmuir lms LangmuirBlodgett lms Organo-clay Saponite Atomic Force Microscopy Infrared reection absorption spectroscopy Attenuated total reectance Fourier transform infrared spectroscopy In general cationic amphiphiles are used to prepare organo-clay hybrid lm in LangmuirBlodgett (LB) tech- nique. In this present communication we demonstrated a unique technique to prepare the organoclay hybrid lms using an anionic amphiphile. The TOT type clay saponite was incorporated onto a oating stearic acid monolayer via a divalent cation Mg 2+ . Salt MgCl 2 was mixed along with the clay dispersion in the LB trough and amphiphile solution was spread onto the subphase in order to make the organo-clay hybrid lms. It was observed that salt (MgCl 2 ) concentration on the subphase affects the organization of nano-dimensional clay platelet (saponite) in organo-clay hybrid lms at airwater interface as well as in LB lms. Noticeable changes in area per molecule and shape of the isotherms were observed and measured at subphases with different salt concentrations. Infrared reection absorption spectroscopy studies reveal that only an in-plane (996 cm -1 ) vibration of ν (Si\O) band occurred when the salt concentration was 10 mM. However, both in-plane (996 cm -1 ) and out-of-plane (1063 cm -1 ) vibrations of the ν (Si\O) band of saponite occurred when the subphase salt concentration was 100 mM. Also the out-of-plane vibra- tion of ν (OH) of saponite was prominent at higher salt concentration. This is because at lower salt concen- tration clay sheets remain at on the surface whereas; at higher MgCl 2 concentration they aggregated and form stacks of saponite layers. Also they may be slightly tilted with a very small tilt angle at higher salt concentration making a favorable condition for both in-plane and out-of-plane vibrations of ν (Si\O) in the hybrid lms. Observed decrease in starting area per molecule in the pressure area isotherm measured at higher salt concentration also supports the tilting of clay layers at airclay dispersion interface. Attentuated total reectance Fourier transform infrared spectroscopy and Atomic Force Microscopy investigations of the hybrid lms prepared with subphase using different salt concentrations also support this thesis. © 2013 Elsevier B.V. All rights reserved. 1. Introduction There is much current interest in the nano-order organo-clay hybrid lms due to their potential use in various technological appli- cations [13]. The organic part in the hybrid material imparts exibil- ity and versatility, whereas, the inorganic clay particles provide structural stability. These types of hybrid material can have unique conducting, semiconducting, dielectric and non-linear properties [3]. Construction of such organo-clay nanostructured hybrid material is an important target of material research. This research has been mo- tivated by the purpose of developing functional materials, which have vast technological applications towards molecular and optoelectronic devices such as sensors, electrode modiers, pyroelectric and non- linear optical devices etc. [15]. Clay minerals are inorganic sheet-like natural mineral. The main interest in clay minerals results essentially from the colloidal size and the permanent structural charge of the clay particles. Through the increased role of surface chemistry and surface physics, properties show up in the materials with dimensions of nanometers that do not exist in larger-sized materials. Smectite clay materials are of valuable mineral class for industrial applications because of their high cation exchange capacities, surface areas, surface reactivities, adsorptive properties, high viscosity and transparency in solution [6]. Thin lm preparing techniques such as LangmuirBlodgett (LB) [710], Layer-by-Layer self assembly [1113], and spin coating [1416] are being used to construct ultrathin hybrid clay lms. The LB technique is preferred because it leads to uniform lms with a good organization of elementary clay mineral layers and surfactants. Also the organization of molecules as well as clay sheets in thin lms can be controlled by changing various LB parameters [710]. These lms can have interesting properties, such as second harmonic genera- tion, photo-induced magnetization, electron transfer and chiral sensing [17]. These properties are determined by the specic organization of the adsorbed molecules and of the elementary clay mineral layers in the LB lms. Evidence of the formation of organo-clay hybrid monolayer Thin Solid Films xxx (2013) xxxxxx Corresponding author at: Department of Physics, Tripura University, Suryamaninagar- 799022, Tripura, India. Tel.: +91 9862804849; fax: +91 381 237 4802. E-mail addresses: [email protected], [email protected] (S.A. Hussain). TSF-31781; No of Pages 8 0040-6090/$ see front matter © 2013 Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.tsf.2013.03.015 Contents lists available at SciVerse ScienceDirect Thin Solid Films journal homepage: www.elsevier.com/locate/tsf Please cite this article as: S.A. Hussain, et al., Thin Solid Films (2013), http://dx.doi.org/10.1016/j.tsf.2013.03.015

Transcript of Incorporation of nano-clay saponite layers in the organo-clay hybrid films using anionic amphiphile...

Thin Solid Films xxx (2013) xxx–xxx

TSF-31781; No of Pages 8

Contents lists available at SciVerse ScienceDirect

Thin Solid Films

j ourna l homepage: www.e lsev ie r .com/ locate / ts f

Incorporation of nano-clay saponite layers in the organo-clay hybrid films usinganionic amphiphile stearic acid by Langmuir–Blodgett technique

Syed Arshad Hussain a,⁎, S. Chakraborty a, D. Bhattacharjee a, R.A. Schoonheydt b

a Department of Physics, Tripura University, Suryamaninagar-799022, Indiab Centres for Surface Chemistry and Catalysis, K.U. Leuven, Kasteelpark Arenberg 23, 3001 Leuven, Belgium

⁎ Corresponding author at: Department of Physics, Tripu799022, Tripura, India. Tel.: +91 9862804849; fax: +91

E-mail addresses: [email protected], tuphusic@

0040-6090/$ – see front matter © 2013 Elsevier B.V. Allhttp://dx.doi.org/10.1016/j.tsf.2013.03.015

Please cite this article as: S.A. Hussain, et al.,

a b s t r a c t

a r t i c l e i n f o

Article history:Received 23 March 2012Received in revised form 5 March 2013Accepted 6 March 2013Available online xxxx

Keywords:Pressure-area isothermLangmuir filmsLangmuir–Blodgett filmsOrgano-claySaponiteAtomic Force MicroscopyInfrared reflection absorption spectroscopyAttenuated total reflectance Fourier transforminfrared spectroscopy

In general cationic amphiphiles are used to prepare organo-clay hybrid film in Langmuir–Blodgett (LB) tech-nique. In this present communication we demonstrated a unique technique to prepare the organo–clayhybrid films using an anionic amphiphile. The T–O–T type clay saponite was incorporated onto a floatingstearic acid monolayer via a divalent cation Mg2+. Salt MgCl2 was mixed along with the clay dispersion inthe LB trough and amphiphile solution was spread onto the subphase in order to make the organo-clayhybrid films. It was observed that salt (MgCl2) concentration on the subphase affects the organization ofnano-dimensional clay platelet (saponite) in organo-clay hybrid films at air–water interface as well as inLB films. Noticeable changes in area per molecule and shape of the isotherms were observed and measuredat subphases with different salt concentrations. Infrared reflection absorption spectroscopy studies revealthat only an in-plane (996 cm−1) vibration of ν (Si\O) band occurred when the salt concentration was10 mM. However, both in-plane (996 cm−1) and out-of-plane (1063 cm−1) vibrations of the ν (Si\O)band of saponite occurred when the subphase salt concentration was 100 mM. Also the out-of-plane vibra-tion of ν (OH) of saponite was prominent at higher salt concentration. This is because at lower salt concen-tration clay sheets remain flat on the surface whereas; at higher MgCl2 concentration they aggregated andform stacks of saponite layers. Also they may be slightly tilted with a very small tilt angle at higher saltconcentration making a favorable condition for both in-plane and out-of-plane vibrations of ν (Si\O) inthe hybrid films. Observed decrease in starting area per molecule in the pressure area isotherm measuredat higher salt concentration also supports the tilting of clay layers at air–clay dispersion interface. Attentuatedtotal reflectance Fourier transform infrared spectroscopy and Atomic Force Microscopy investigations of thehybrid films prepared with subphase using different salt concentrations also support this thesis.

© 2013 Elsevier B.V. All rights reserved.

1. Introduction

There is much current interest in the nano-order organo-clayhybrid films due to their potential use in various technological appli-cations [1–3]. The organic part in the hybrid material imparts flexibil-ity and versatility, whereas, the inorganic clay particles providestructural stability. These types of hybrid material can have uniqueconducting, semiconducting, dielectric and non-linear properties [3].Construction of such organo-clay nanostructured hybrid material isan important target of material research. This research has been mo-tivated by the purpose of developing functional materials, which havevast technological applications towards molecular and optoelectronicdevices such as sensors, electrode modifiers, pyroelectric and non-linear optical devices etc. [1–5].

Clay minerals are inorganic sheet-like natural mineral. The maininterest in clay minerals results essentially from the colloidal size

raUniversity, Suryamaninagar-381 237 4802.sancharnet.in (S.A. Hussain).

rights reserved.

Thin Solid Films (2013), http

and the permanent structural charge of the clay particles. Throughthe increased role of surface chemistry and surface physics, propertiesshow up in the materials with dimensions of nanometers that do notexist in larger-sized materials.

Smectite clay materials are of valuable mineral class for industrialapplications because of their high cation exchange capacities, surfaceareas, surface reactivities, adsorptive properties, high viscosity andtransparency in solution [6].

Thin film preparing techniques such as Langmuir–Blodgett (LB)[7–10], Layer-by-Layer self assembly [11–13], and spin coating[14–16] are being used to construct ultrathin hybrid clay films. TheLB technique is preferred because it leads to uniform films with agood organization of elementary clay mineral layers and surfactants.Also the organization of molecules as well as clay sheets in thin filmscan be controlled by changing various LB parameters [7–10]. Thesefilms can have interesting properties, such as second harmonic genera-tion, photo-induced magnetization, electron transfer and chiral sensing[17]. These properties are determined by the specific organization ofthe adsorbed molecules and of the elementary clay mineral layers inthe LBfilms. Evidence of the formation of organo-clay hybridmonolayer

://dx.doi.org/10.1016/j.tsf.2013.03.015

2 S.A. Hussain et al. / Thin Solid Films xxx (2013) xxx–xxx

includes surface pressure vs. area per molecule (π − A) isothermmea-surement [8,17], Atomic Force Microscopy (AFM) [8,9,17], Fouriertransform infrared spectroscopy (FTIR) [7], X-ray diffraction [18], etc.The π − A isotherm measurements are strongly influenced by thepresence of clay minerals. The lift off area of the surfactant is shiftedtowards larger area in the presence of clays.

In order to form organo-clay hybrid LB films usually amphiphiliccations are being used [7,9,10,17]. When suitable amphiphilic cationsare spread onto an aqueous clay dispersion interface in a LB trough,negatively charged clay platelets in the dispersion are adsorbed ontothe bottom of the floating amphiphile monolayer through ion exchangereaction between the clay layers in the subphase and the amphiphile atair–water interface. Thus a hybrid monolayer is formed which consistsof inorganic clay platelets covered on one side by one layer of adsorbedamphiphile at air–clay dispersion interface. Accordingly, this hybridmonolayer is transferred onto solid substrate to form LB films.

Now-a-days organo-clay nanocomposites as well as organo-clayhybrid films are of intense interest for researchers and are being ex-tensively investigated in material science [19,20]. There are two levelsof organizations which actually exist in organo-clay hybrid films:(i) the organization of organic molecules on the clay mineral surfaceas well as on the substrate and (ii) the organization of clay mineral onsolid substrate. If the orientation of the incorporated molecules can becontrolled, then the hybrid clay composite materials would be applica-ble to devices for current rectifying, nonlinear optics, one-way energytransfer, Fluorescence Resonance Energy Transfer (FRET), etc. [1,2,4,5].

Depending on the nature of the amphiphilic cations and theiraggregation in the films, films with various properties have beenobtained. Up to now the functionalities of the films, that were inves-tigated, were limited to optics: nonlinear optics, second harmonicgeneration, FRET, etc. It would be interesting to extend the range offunctionalities of the hybrid films, based on clay mineral layers. Asthese functionalities depend on the amphiphilic molecules and theirorganization in the films, it would be interesting to extend the rangeof molecules from amphiphilic cations, investigated until now, to neu-tral amphiphiles and amphiphilic anions. However, there is hardly anyreport of using anionic amphiphile in order to form organo-clay hy-brid LB films. We tried to form organo-clay hybrid film using anionicamphiphile stearic acid (SA) and clay platelet saponite. As both SAand saponite are anionic, normally there was no interaction betweenSA and saponite. However, when salt MgCl2, was mixed in the claydispersion subphase clay platelet saponite was incorporated in thefloating SAmonolayer. Here divalent cationMg2+ acts as themediatorto incorporate the negatively charged inorganic clay layers onto theanionic amphiphile SA layer. We also tried with monovalent cationsuch as Na1+ but failed to prepare hybrid film. Being divalent Mg2+

acts as a mediator in between anionic clay and anionic amphiphile.So in this present paper we report a unique method to incorporateanionic clay saponite onto a floating anionic amphiphile monolayer,which can be extended to other similar systems.

In the present paper we reported the organization of a well knownclay mineral saponite in Langmuir films on the air–water interface.Stearic acid (SA) was used as amphiphilic surfactant and MgCl2 wasused as the mediator to adsorb the saponite sheets onto floating SAmonolayer via divalent cation Mg2+. It was observed that the concen-tration of MgCl2 affects the organization of saponite sheets at the air–water interface aswell as onto solid substrate. At lower salt concentra-tion clay sheets remain flat on the surface whereas, at higher MgCl2concentration they were tilted to some extent and get aggregated.

2. Experimental details

2.1. Materials

Stearic acid (SA, 99%, Sigma-Aldrich Chemical Co.) was used as re-ceived without further purification. SA was dissolved in HPLC grade

Please cite this article as: S.A. Hussain, et al., Thin Solid Films (2013), http

chloroform (99.9%, Sigma-Aldrich Chemical Co.). The clay mineralsaponite used in this studywas obtained from the Source Clays Repos-itory of the Clay Minerals Society. The clay minerals were stored asfreeze-dried powders. Saponite is a 2:1 or T–O–T type of clay withthree sheets, silica–alumina–silica in a sandwich like formation. Ingeneral saponite clay platelets consist of stacks of such 2:1 layers.However it is possible to make highly dispersed clay suspensionconsisting of single clay layers [9,10]. Such clay dispersions were pre-pared by stirring the aqueous clay dispersion for more than 24 h witha magnetic stirrer and followed by ultrasonication for about half anhour before use [9,10].

2.2. Isotherm measurement and film formation

A commercially available Langmuir–Blodgett (LB) film depositioninstrument (APEX 2000C, India) was used for isotherm measurementand hybrid monolayer film preparation. Either triple distilled deionisedwater (resistivity 18.2 MΩ cm), or clay dispersions stirred for 24 h(by a magnetic stirrer) in triple distilled deionised water was usedas subphase. The clay concentration was fixed at 10 mg/L (10 ppm).In few cases MgCl2 with different concentrations was added in thesubphase. 50 μl of chloroform solution of SA (1.75 × 10−3 mol/l) wasspread on the subphase of the LB trough by a micro syringe. Allowing30 min to evaporate the solvent and to complete the hybrid film forma-tion, the floating monolayer was compressed at a rate of 5 mm/minto monitor the pressure-area isotherm. Surface pressure was recordedusing a Wilhelmy arrangement [21]. Each isotherm was repeatedseveral times and consistent results were obtained. The monolayerswere then transferred very carefully by Y type (vertical) deposition at15 mN/m surface pressure (lifting speed 5 mm min−1) onto smoothsilicon wafer for AFM measurement and onto a zinc sellenide singlecrystal substrate for attenuated total reflectance Fourier transform in-frared spectroscopy (ATR-FTIR) measurement. The transfer ratio wasfound to be 0.96 ± 0.02. The room temperature was 24 °C.

2.3. Infrared reflection absorption spectroscopy and ATR-FTIRspectroscopy

Infrared reflection absorption spectroscopy (IRRAS) of the Lang-muir monolayer at the air–water interface was done by a specialaccessory attached with a Bruker IFS66v/s FTIR spectrometer. ForIRRASmeasurement themonolayer film on the liquid surfacewas pre-pared in a polytetrafluoroethylene trough (area = 17.6 cm2) with amanually controlled barrier. Details of the IRRAS measurement werereported elsewhere [9,22]. The trough was filled with aqueous disper-sion of saponite. MgCl2 was added in the subphase at different concen-trations. A microsyringe was used to spread 3 μl chloroform solutions(1.75 × 10−3 mol/l) of SA on the air–clay dispersion subphase of thetrough. Then the spectra were recorded with regular time interval. Atotal of 512 scans were signal averaged using an optical resolution of2 cm−1. The films on the liquid surface were measured at 40° inci-dence to have maximum signal strength. The background spectrawere the subphase surface before spreading the SA solution. The spec-trometer used in this study was equipped with a liquid nitrogencooled Mercury cadmium telluride detector and a KBr beam splitter.

ATR-FTIR measurement of the SA–Mg2+–saponite hybrid film wasdone by a FTIR spectrophotometer (Spectrum 100, PerkinElmer). Aclean zinc–sellenide substrate was used for background measurement.

2.4. Atomic Force Microscope

The AFM images of hybrid monolayer films were taken in air with acommercial AFM systemAutoprobeM5 (Veeco Instr.) using silicon can-tileverswith a sharp, high apex ratio tip (UltraLevers™, Veeco Instr.). Allthe AFM images presented here were obtained in intermittent-contact

://dx.doi.org/10.1016/j.tsf.2013.03.015

3S.A. Hussain et al. / Thin Solid Films xxx (2013) xxx–xxx

(“tapping”) mode. Typical scan areas were 2 × 2 μm2. The monolayerson Si wafer substrates were used for the AFMmeasurements.

3. Results and discussions

3.1. Monolayer characteristics at the air–water interface

Surface pressure vs. area per molecule (π − A) isotherms of SA on10 mMMgCl2 subphase and 100 mMMgCl2 subphase with and with-out clay are shown in Fig. 1a and b respectively. The isotherms of SAon pure water and on clay subphase are also shown for comparison.

SA isotherm on pure water subphase indicates that surface pres-sure starts to increase at the area per molecule of ~0.32 nm2 uponcompression. The mean molecular area estimated by extrapolatingthe steeply rising part of the isotherm curve to zero pressure is~0.23 nm2, which is consistent with the area occupied by denselypacked alkyl chains [21,23]. Also the area permolecule for SA isothermis 0.21 nm2 at a surface pressure of 25 mN/m,which is consistentwiththe reported results [21,23].

π − A isotherm of pure SA shows a gaseous state between surfacepressures of 0 to 1.8 mN/m, where the area per molecule drops from0.32 nm2 to 0.28 nm2. Then a liquid phase is observed up to the sur-face pressure of ~13.6 mN/m. After which the isotherm shows steeprising up to the collapse pressure is reached at about 53.4 mN/m.This is an indication of attainment of solid condensed phase of theSA monolayer.

The SA isotherm in presence of MgCl2, is shifted to a slightly lowerarea (curve 3 of Fig. 1a and b), revealing the complexation betweenMg2+ ion of MgCl2 and SA monolayer at air–water interface resulting

0.18 0.24 0.30 0.360

10

20

30

40

50

60

70

b

3

2 1 4

pure water (1) 10 ppm saponite (2)

pure water (1) 10 ppm saponite (2) 10 mM MgCl2 (3)

10 mM MgCl2 (3)

10 mM MgCl2 + 10 ppm saponite (4)

10 mM MgCl2 + 10 ppm saponite (4)

Area per molecule (nm2)

0.18 0.24 0.30 0.360

10

20

30

40

50

60

70

80

a

4

3

21

Su

rfac

e p

ress

ure

(m

N/m

)

Fig. 1. Surface pressure-area per molecule isotherms of SA on different subphases[a] purewater (1), 10 ppm saponite (2), 10 mMMgCl2 (3), and 10 mMMgCl2 + 10 ppmsaponite (4) and [b] pure water (1), 10 ppm saponite (2), 100 mM MgCl2 (3), and100 mM MgCl2 + 10 ppm saponite (4).

Please cite this article as: S.A. Hussain, et al., Thin Solid Films (2013), http

in the straightening up of alkyl chain. Here for both the salt concen-trations (10 mM and 100 mM) the liquid phase of the isothermsexpanded up to the surface pressure of 26.2 mN/m and the gaseousphase becomes almost indistinguishable after zero surface pressure.The collapse pressure in the presence of salt is larger compared topure SAmonolayer, suggesting that Mg2+ ion stabilizes themonolayer.There are several literatures available on this topic [21].

Marked changes in the isotherms are observed when the SA iso-therms are measured on the clay (saponite) dispersed subphasecontaining the salt MgCl2. Here the isotherms are shifted to a compar-atively larger area than that in the absence of clay. This is an indicationthat clay sheets are incorporated into the monolayer at air–waterinterface. However no change in SA isotherm is observed when theisotherm is measured only on the clay dispersion subphase (curve 2of Fig. 1a and b). The isotherm almost overlaps the pure SA isothermmeasured on pure water. This is because both clay and SA are anionicand no interaction occurred in between them. In the presence ofMgCl2, the divalent cation Mg2+ interacts with the SA molecule offloating monolayer and the clay platelets present at the subphase.Thus the clay platelets are incorporated onto the monolayer via diva-lent cation Mg2+. When the subphase is 10 ppm saponite + 10 mMMgCl2 solution (curve 4 of Fig. 1a), the isotherm starts rising at a sur-face area of 0.4 nm2. The gaseous phase exists up to ~0.5 mN/m andliquid phase extends up to 20 mN/m. After which the solid condensedphase is achieved and the surface pressure rises up to 56.5 mN/m,where, the collapse of monolayer occurred. On the other hand the SAisotherm on 10 ppm saponite + 100 mM MgCl2 solution (curve 4 ofFig. 1b) starts rising at an area per molecule of 0.34 nm2. Here the iso-therm shows two liquid phases: (i) liquid expanded phase (0.2–7.4 mN/m) and (ii) liquid condensed phase (7.4–23.3 mN/m). Afterwhich the solid condensed state is reached and the surface pressurerises steeply up to the attainment of collapse pressure at ~61 mN/m.The observed decrease of starting area per molecule and the existenceof two types of liquid phases at higher MgCl2 concentration indicatethat some orientational changes of clay sheets occurred in the mono-layer. Our investigations ofmonolayer using IRRAS, ATR-FTIR and AFMrevealed that at lower salt concentration the clay sheets remain flat inthe monolayer, whereas, the clay sheets are somewhat tilted in themonolayer with higher MgCl2 concentration due to aggregation andovercrowding of saponite layers at air–clay dispersion interface.

3.2. Infrared reflection absorption spectroscopy (IRRAS) and ATR-FTIRstudies

The IRRAS technique is very attractive because it can be applied toobtain spectra from monolayer directly on the air–water interface,which provides information about themolecular orientation [9,22,24].

Fig. 2a and b shows the IRRAS spectra of SA monolayer on 10 ppmsaponite + 10 mMMgCl2 subphase and 10 ppm saponite + 100 mMMgCl2 subphase respectively measured at different time intervalsafter spreading of SA solution. The figures are presented in reflectancemode, as the absorption band of IRRAS curves in dielectric medium isnegative [24].

The IRRAS spectra of SA monolayer on 10 ppm saponite + 10 mMMgCl2 subphase possess strong prominent bands at 2851 cm−1

and 2920 cm−1. These two bands are the diagnostic bands of SAand identified as the stretching vibrations of CH2 group of SA [25,26].The prominent band at 996 cm−1 corresponds to the in-plane Si\Ostretching vibration, ν (Si\O), of saponite [27]. Also almost indistin-guishable weak hump at 1063 cm−1 is observed which is assignedas due to out-of-plane vibration of ν (Si\O) band. The out-of-planeν (OH) vibration at 3680 cm−1 is almost indistinguishable. It may bementioned in this context that this band gives information about theorientation of the saponite clay. Ras et al. [7] performed polarizedATR-FTIR measurement of a hybrid LB film of saponite clay sheets.They found prominent ν (OH) band at 3680 cm−1 using p-polarized

://dx.doi.org/10.1016/j.tsf.2013.03.015

4 S.A. Hussain et al. / Thin Solid Films xxx (2013) xxx–xxx

light and a very low intense band using s-polarized light with dichroicratio 0.33 which is consistent with the orientation of OH group normalto the clay plane. With increasing time the intensity of IRRAS band in-creases and reaches its maximum with waiting time of ~20 min. Thishas been shown in the inset of Fig. 2a. Inset of Fig. 2a was drawn con-sidering the baseline correction.

It is worthwhile to mention in this context that when the IRRASspectra were measured for SA monolayer on clay subphase in the ab-sence of MgCl2, the ν (Si\O) band at 996 cm−1 was not observed[figure not shown]. However, when the IRRAS spectra were recorded

Fig. 2. a. IRRAS spectra of SA monolayer on 10 mMMgCl2 and 10 ppm saponite subphase takregion and (ii) intensity of 996 cm−1 band as a function of time. b: IRRAS spectra of SA mtervals after spreading of SA solution. Inset: (i) enlarged ν (Si\O) region and (ii) intensity ohybrid films for two different salt concentrations: (i) ν (OH) region and ν (Si\O) region. Thtime) and 30 min (maximum waiting time) for 10 mM and 100 mM MgCl2 respectively.

Please cite this article as: S.A. Hussain, et al., Thin Solid Films (2013), http

in presence of MgCl2 and clay, the ν (Si\O) band at 996 cm−1 wasobserved. This indicates that when negatively charged clay plateletscome in contact with the floating SA monolayer no interaction oc-curred between SA and clay being both anionic. However, in the pres-ence of MgCl2, the divalent Mg2+ ion interacts with the SA moleculeson the floating monolayer along with the clay platelets in subphasethrough electrostatic interaction. Thus MgCl2 acts as the mediator toincorporate the clay sheets into the floating monolayer at the air–water interface. Number of clay sheets adsorbed onto the floatinglayer increases with time resulting in an increase in surface coverage

en at different time intervals after spreading of SA solution. Inset: (i) enlarged ν (Si\O)onolayer on 100 mM MgCl2 and 10 ppm saponite subphase taken at different time in-f 996 cm−1 band as a function of time. c: Enlarged IRRAS spectra of SA–Mg2+–saponitee spectra were taken from Fig. 2a and 2b for waiting time of 26 min (maximum waiting

://dx.doi.org/10.1016/j.tsf.2013.03.015

3200 3400 3600 3800 40001.02

1.04

1.06

1.08

1.10

1.12(i)

3681

ν(OH) region

10 mM

100 mM

Inte

nsi

ty

Wavenumbers (cm-1) Wavenumbers (cm-1)

800 1000 1200 14001.02

1.04

1.06

1.08

1.10

1.12 (ii)

ν(Si-O) region

1063

998

10 mM

100 mM

Inte

nsi

ty

c

Fig. 2 (continued).

5S.A. Hussain et al. / Thin Solid Films xxx (2013) xxx–xxx

as well as the thickness of the floating organo-clay hybrid film. Thisincrease in thickness as well as surface coverage of the floating hybridfilm causes an increase in reflectance of 996 cm−1 band. Howeverafter a certain time interval the reaction is completed. Consequentlythe reflectance also reaches itsmaximumand becomes almost constant.

It may be mentioned in this context that we tried to perform thisexperiment using a mono-valent salt NaCl. However, we did not ob-serve the characteristic bands of clay in the IRRAS spectra indicatingthat clay-SA hybrid films were not formed at the air–water interface.

The IRRAS spectra of SA monolayer taken on 10 ppm saponite +100 mM MgCl2 subphase are quite interesting. Here in addition tothe in-plane ν (Si\O) band at 996 cm−1 the out-of-plane ν (Si\O)band at 1063 cm−1 is also present. Also an intense out-of-plane ν(OH) stretching band is observed. This indicates that the organizationof saponite platelets onto the floating monolayer is different for lowerand higher MgCl2 concentrations. The absence of 1063 cm−1 band atlower MgCl2 concentration suggests that the out-of-plane vibration ofν (Si\O) band of saponite is restricted. However, both in-plane andout-of-plane vibrations of ν (Si\O) band occurred in the case ofhigher MgCl2 concentration. This is possible if the organizations ofclay layers are different in the monolayer. At lower salt concentrationthe clay sheets remain flat in the monolayer, which in turn prohibitthe out-of-pane vibration of ν (Si\O) band. However, at higher saltconcentration some reorientation of clay layers occurred. Due to thepresence of larger number of Mg2+ ion there may be overcrowdingof saponite layers resulting in the formation of stacks and aggregates.Also there may be a possibility of formation of preformed aggregatesof clay platelets in the subphase at higher MgCl2 concentration. Ac-cordingly these aggregates are adsorbed onto the SA monolayer atthe interface in the presence of a large number of Mg2+ ions at ahigher salt concentration. Consequently the clay layers may be tiltedwith a very small tilt angle in the monolayer, which makes favorableconditions for both in-plane and out-of-plane vibrations of ν (Si\O)band in the hybrid monolayer. This is also supported by the observeddecrease in starting area per molecule of the pressure area isothermat higher salt concentration. However, to know about the exact orien-tation of the clay layers in Langmuir and LB films further study isgoing on in our laboratory.

Please cite this article as: S.A. Hussain, et al., Thin Solid Films (2013), http

Fig. 2c shows the enlarged form of the bands at lower and higherMgCl2 concentrations for comparison. A clear difference in intensityof 1063 cm−1 out-of-plane ν (Si\O) band and 3681 cm−1 out-of-plane ν (OH) band for two different salt concentrations gives a clearevidence of the different orientations of saponite layers for two dif-ferent MgCl2 concentrations. It was observed that the intensity of ν(Si\O) band gets smaller at 100 mM concentration of MgCl2. Thismay be due to the change in orientation of the clay platelets whichare supposed to occur for overcrowding and aggregation of the clayplatelets at higher salt concentration. An explanation has been givenlater using a schematic diagram.

The ATR-FTIR spectra of hybrid LB films (Fig. 3) are also in goodagreement with the IRRAS spectra. The ATR-FTIR spectra of SA LBfilm prepared on 10 ppm saponite + 10 mMMgCl2 subphase possessonly the in-plane ν (Si\O) band at 996 cm−1 and the out-of-planeband is absent. On the other hand both in-plane and out-of-planebands of ν (Si\O) bond are present in the spectrum of SA LB film pre-pared using 10 ppm saponite + 100 mM MgCl2 subphase. This alsosupports the change in orientation of clay sheets at two different saltconcentrations.

Our AFM investigations give compelling visual evidence of changein orientation of clay sheets in organo-clay hybrid films with changein salt concentration.

3.3. Atomic Force Microscopy

In order to have an idea about the organization of clay sheets wehave taken the AFM images of LB films prepared at lower as well ashigher MgCl2 concentrations of clay dispersed subphase. The repre-sentative images are shown in Figs. 4a and b.

Here the monolayer films are transferred onto smooth silicon sub-strates at 15 mN/m surface pressurewith waiting time of 15 min afterreaching this surface pressure. This time is allowed in order to stabilizethe film.

It is interesting to mention in this context that the elementary clayminerals in the subphase are attracted to the interface (against grav-itational forces) to form hybrid monolayer, which is metastable atthe air–water interface. The waiting time for adsorption and before

://dx.doi.org/10.1016/j.tsf.2013.03.015

Fig. 4. a. AFM image of LB film of SA prepared on 10 mM MgCl2 + 10 ppm saponite subphafilm of SA prepared on 100 mM MgCl2 + 10 ppm saponite subphase. The height profile alo

1000 1500 2000 2500 3000

0.6

0.8

1.0

1.2

1.4b

ν(Si-O) 1063998

1462

28482922

%T

Wavenumbers (cm-1)

1000 1500 2000 2500 3000

0.6

0.8

1.0

1.2

1.4a

29212849

1461997

ν(Si-O)

%T

Fig. 3. ATR-FTIR spectra of SA–Mg2+–saponite hybrid LB monolayer for two differentsalt concentrations: (a) 10 mM MgCl2 and (b) 100 mM MgCl2.

6 S.A. Hussain et al. / Thin Solid Films xxx (2013) xxx–xxx

Please cite this article as: S.A. Hussain, et al., Thin Solid Films (2013), http

compression was 30 min. This time was given to complete the ad-sorption process. After this time the barrier was compressed to recordthe isotherm or to reach a particular surface pressure for film deposi-tion. Here the film was deposited at 15 mN/m surface pressure. Afterreaching 15 mN/m surface pressure 15 min was allowed to stabilizethe surface pressure before film deposition on to the substrate. Upto 30 min the surface pressure remains almost fixed. It was observedthat with increase in waiting time (>30 min) the individual hybrid-ized clay layers diffuse slowly into the subphase indicating a decreasein surface pressure. So the waiting time is kept fixed at 15 min beforedeposition. T. Szabo et al. [28] reported in one of their papers that forsaponite particles with increasing waiting time between the mono-layer formation and transfer to the substrate, the particle density inthe monolayer decreased. They prepare the LB film at waiting timeof 15 min and 60 min respectively and found that the particle densityis much less in the film lifted at 60 min We also observe that withwaiting time greater than 30 min before deposition the surface pres-sure decreases due to diffusion of hybridized clay layers into thesubphase. So we optimize the waiting time 15 min after reaching15 mN/m surface pressures and before deposition.

From the AFM images one can note marked changes with respectto textures of the films. Fig. 4a shows the AFM image of hybrid filmprepared using 10 mM MgCl2 concentration. The saponite layerswith various size and shapes are clearly visible. It is observed thatthe saponite sheets are lying flat on the substrate and individual sap-onite layers are distinguishable. The thickness of the clay sheets are ofthe order of ~2–3 nm. The lateral sizes of the clay platelets are closeto 0.5–1 μm. This information is realistic for a single saponite layerlying flat on the substrate surface [7,28]. The surface coverage is al-most ~70%. However the AFM image of hybrid LB monolayer pre-pared on 100 mM MgCl2 concentration is quite different than that

se. The height profile along lines I and II in the image is also shown. b: AFM image of LBng lines I and II in the image is also shown.

://dx.doi.org/10.1016/j.tsf.2013.03.015

7S.A. Hussain et al. / Thin Solid Films xxx (2013) xxx–xxx

prepared on 10 mMMgCl2 concentration. Here it appears that almostall the surfaces are covered with the clay layers. Also the individualclay layer is indistinguishable. From the height profile analysis it isobserved that here the height of films across the line is bigger thanthe previous case. The highest thickness is ~16–17 nm across theline shown. This thickness is equal to the stacking of several claylayers. At higher salt concentration the Mg2+ concentration in thesubphase is large. This may trigger overcrowding of clay platelets atthe air–clay dispersion. This overcrowding may result in a small tiltof the organization of the saponite layers. Such tilting of saponitelayer may also contribute to the observed increase in height. However,the tilt angle should be extremely small because if the tilt angle is larger,then the surface roughness should be extremely large, since theplateletsshould be viewed edge-on or oblique.

It is worthwhile tomention in this context that IRRAS studies showthat the out-of-plane vibration is restricted on the films at air–waterinterface for lowerMgCl2 concentration. The flat organization of sapo-nite sheet prohibits the out-of-plane vibration. On the other hand thetilted organization of clay sheets in the films with higher MgCl2 con-centration allows both the in-plane and out-of-plane vibrations of ν(Si\O) band of saponite. So it can be concluded that concentrationof MgCl2 in the subphase plays a vital role in controlling the orienta-tion and organization of clay particles at the air–water interface.

Fig. 5. Schematic representation of SA–Mg2+–clay film at air–water interface f

Please cite this article as: S.A. Hussain, et al., Thin Solid Films (2013), http

3.4. Schematic diagram

In order to explain the flat and tilted organization of clay plateletsat different salt concentrations schematic diagrams are shown inFig. 5a and b. Mg2+ is a divalent cation whereas clay and SA bothare anionic. Mg2+ divalent cation acts as a mediator to both SA andclay as shown in Fig. 5a and b.

It is worthwhile to mention in this context that, the elementaryclay minerals in the subphase are attracted to the interface (againstgravitational forces) to form hybrid monolayer at the air–water inter-face. At lower salt concentration most of the Mg2+ ions take part toimmobilize the clay on the SA monolayer. Also the tendency to formpreformed aggregate on the subphase is very less due to the presenceof a lower number of Mg2+ ions. The tendency to adsorb preformedaggregates at air–water interface against gravity is also less due tothe presence of a lower number of Mg2+ cation. As a result the SA–Mg2+–saponite hybrid film consists of mainly single saponite sheetslying flat at air–water interface and in LB film. This has been shownschematically in Fig. 5a.

Fig. 5b shows the schematic diagram of SA–Mg2+–saponite hybridfilm at higher salt concentration. The presence of a large number ofMg2+ cations at higher salt concentration creates a favorable environ-ment for preformed aggregate formation (Fig. 4b). Due to the presence

or (a) 10 mM MgCl2 concentration and (b) 100 mM MgCl2 concentration.

://dx.doi.org/10.1016/j.tsf.2013.03.015

8 S.A. Hussain et al. / Thin Solid Films xxx (2013) xxx–xxx

of a large number of Mg2+ cations these preformed clay aggregatesare easily adsorbed onto floating SA monolayer. Due to the presenceof a larger number of Mg2+ ions the electrostatic interaction betweensaponite aggregates, Mg2+ ion and amphiphilic SA molecules is suffi-cient to overcome the gravitational forces. Also there is a possibility toadsorb successive saponite sheets at air–water interface due to thepresence of a large number of Mg2+ ions, resulting in an overcrowdingaswell as aggregate formation. The height profile analysis of AFM imageat higher salt concentration also reveals that aggregates of three to fourlayers of saponite are formed at air–water interface and no single sapo-nite sheet is available. There may be a possibility of slight tilting of thesaponite sheets due to the presence of saponite aggregates as well asoverlapping of several saponite sheets. Also at higher salt concentrationmanyMg2+ ions lie within the subphase and may attach to the bottomof the floating clay platelets or clay aggregates. This will also contributeto the tilting of clay sheets at the interface.

Saponite particles are very soft [29]. It was observed that whenthey are transferred onto any rough surfaces, they take the textureof the surface and lost their actual shape [29]. In the present casedue to the overcrowding and aggregation as well as tilting the sapo-nite particle lost their actual shape. Accordingly, in the AFM image ofhybrid film prepared at a higher salt concentration, no definite shapeor size of the individual saponite particles is observed.

4. Conclusion

In conclusion in the present work we have demonstrated the prep-aration of organo–clay hybrid films using the LB technique. Generallyin LB technique amphiphilic cations are used to incorporate anionicclay layers onto hybrid films. However, here we have used an anionicamphiphile SA. MgCl2 was used to incorporate the clay layers onto thefloating SA layers through the divalent cation Mg2+. T–O–T type claysaponite was used in this study. This has been investigated using sur-face pressure vs. area per molecule isotherm, IRRAS, and AFM studies.It was found that the organization of saponite layers was largely af-fected by the subphase MgCl2 concentration. Marked changes in iso-therms measured at different salt concentrations were observed. Thestarting area per molecule for isotherm measured at 10 mM MgCl2concentration was larger than that measured at 100 mM MgCl2concentration. This is an indication of reorientation of clay layersat the interface. In-situ IRRAS investigations of organo-clay hybridmonolayer at 10 mM salt concentration reveal that only the in-plane(996 cm−1) vibration of the ν (Si\O) band of saponite exists. Where-as, for the films prepared at 100 mM salt concentration, both thein-plane (996 cm−1) and out-of-plane (1063 cm−1) vibrations oc-curred. Also a clear difference in out-of-plane ν (OH) band intensityfor lower and higher MgCl2 concentration was observed. These resultsindicate that the orientation of saponite layers in the organo-clay hy-brid films at air–water interface as well as on to solid substrate is dif-ferent for different salt concentrations. At lower salt concentration theclay sheets remain flat in the films, consequently the out-of-planevibration is restricted. Whereas, at higher salt concentration the claysheets are somehow tilted such that the two types of vibration can

Please cite this article as: S.A. Hussain, et al., Thin Solid Films (2013), http

occur. This may be due to overcrowding of saponite layers in thepresence of a large number of Mg2+ cations. ATR-FTIR and AFM inves-tigations of the organo-clay hybrid films prepared at different salt con-centrations also support this thesis.

Acknowledgments

The author SAH is grateful to DST, CSIR and DAE for financial sup-port to carry out this research work through DST Fast-Track projectRef. No. SE/FTP/PS-54/2007, CSIR project Ref. 03(1146)/09/EMR-II andDAE Young Scientist Research Award (No. 2009/20/37/8/BRNS/3328).DB is grateful to DST for financial support through DST project Ref No:SR/S2/LOP-19/07. The author SC is grateful to CSIR for financial supportto carry out this research work through SRF award (No. 09/714(0014)/2012-EMR-I).

References

[1] Y. Lvov, Y.K. Ariga, I. Ichinose, T. Kunitake, Langmuir 12 (1996) 3038.[2] N.A. Kotov, T. Haraszi, L. Turi, G. Zavala, R.E. Geer, I. Dekany, J.H. Fendler, J. Am.

Chem. Soc. 119 (1997) 6821.[3] F. Bergaya, B.K.G. Theng, G. Lagaly, Handbook of Clay Science, Elsevier, Amsterdam,

2006.[4] I. Ichinose, N. Kimizuka, T. Kunitake, J. Phys. Chem. 99 (1995) 3736.[5] N. Tsutsumi, K. Sakata, T. Kunitake, Chem. Lett. 21 (1992) 1465.[6] K.A. Carrado, A. Awaluddin, Prepr. Am. Chem. Soc. Div. Petrol. Chem. 38 (1993) 518.[7] R.H.A. Ras, C.T. Johnston, E.I. Franses, R. Ramaekers, G. Maes, P. Foubert, F.C. De

Schryver, R.A. Schoonheydt, Langmuir 19 (2003) 4295.[8] Y. Umemura, A. Yamagishi, R.A. Schoonheydt, A. Persoons, F.C. De Schryver, J. Am.

Chem. Soc. 124 (2002) 992.[9] S.A. Hussain, N. Islam, D. Bhattacharjee, J. Phys. Chem. Solids 71 (2010) 323.

[10] S.A. Hussain, R.A. Schoonheydt, Langmuir 26 (2010) 11870.[11] N.A. Kotov, Mater. Res. Bull. 26 (2001) 992.[12] K. Glinel, A. Laschewski, A.M. Jonas, J. Phys. Chem. B 106 (2002) 11246.[13] B. van Duffel, R.A. Schoonheydt, C.P.M. Grim, F.C. De Schryver, Langmuir 15 (1999)

7520.[14] B.V. Lotsch, G.A. Ozin, Adv. Mater. 20 (2008) 4079.[15] B.V. Lotsch, G.A. Ozin, ACS Nanosci. 2 (2008) 2065.[16] F. Lopez Arbeloa, V. Martinez Martinez, Chem. Mater. 18 (2006) 1407.[17] R.H.A. Ras, Y. Umemura, A. Yamagishi, R.A. Schoonheydt, Phys. Chem. Chem. Phys.

9 (2007) 918.[18] S. Takahashi, M. Taniguchi, K. Omote, N. Wakabayashi, R. Tanaka, A. Yamagashi,

Chem. Phys. Lett. 352 (2002) 213.[19] T.J. Pinnavaia, J.R. Butruille, in: G. Alberti, T. Bein (Eds.), Comprehensive Supramo-

lecular Chemistry, vol. 7, Elsevier, Oxford, 1996, (Chapter 7).[20] In: T.J. Pinnavaia, G.W. Beall (Eds.), Polymer–Clay Nanocomposites, Wiley, New

York, 2000.[21] A. Ulman, An Introduction to Ultrathin Organic Films: From Langmuir–Blodgett

Films of Self Assemblies, Academic Press, New York, 1991.[22] R.H.A. Ras, J. Nemeth, C.T. Johnston, I. Dekany, R.A. Schoonheydt, Thin Solid Films

466 (2004) 291.[23] S. Deb, S. Biswas, S.A. Hussain, D. Bhattacharjee, Chem. Phys. Lett. 405 (2005) 323.[24] R.A. Dluhy, S.M. Stephens, S. Widayati, A.D. Williams, Spectrochim. Acta A 51

(1995) 1413.[25] G. Xiong, Z. Zhi, X. Yang, L. Lu, X. Wang, J. Mater. Sci. Lett. 16 (1997) 1064.[26] V.E. Berkheiser, Cáliz Cáliz Miner, 30, 1982, p. 91.[27] V.C. Farmer, The Infrared Spectra ofMinerals, Mineralogical Society, London, 1974.

331.[28] T. Szabo, J. Wang, A. Volodin, C. van Haesendonck, I. Dekany, R.A. Schoonheydt,

Clays Clay Miner. 57 (2009) 706.[29] R.D. Piner, T.T. Xu, F.T. Fisher, Y. Qiao, R.S. Ruoff, Langmuir 19 (2003) 7995.

://dx.doi.org/10.1016/j.tsf.2013.03.015