spectrochimica acta part a: molecular and biomolecular spectroscopy · 2016-07-29 · 80 c. hansda...

9
Layer-by-layer lms and colloidal dispersions of graphene oxide nanosheets for efcient control of the uorescence and aggregation properties of the cationic dye acridine orange Chaitali Hansda a,b , Utsav Chakraborty a , Syed Arshad Hussain c , Debajyoti Bhattacharjee c , Pabitra Kumar Paul a, a Department of Physics, Jadavpur University, Jadavpur, Kolkata 700032, India b Department of Physics, The University of Burdwan, Golapbag, Burdwan 713104, India c Department of Physics, Tripura University, Suryamaninagar, 799022, Tripura West, India abstract article info Article history: Received 13 August 2015 Received in revised form 9 October 2015 Accepted 5 December 2015 Available online 15 December 2015 Chemically derived graphene oxide (GO) nanosheets have received great deal of interest for technological appli- cation such as optoelectronic and biosensors. Aqueous dispersions of GO become an efcient template to induce the association of cationic dye namely Acridine Orange (AO). Interactions of AO with colloidal GO was governed by both electrostatic and ππ stacking cooperative interactions. The type of dye aggregations was found to de- pend on the concentration of GO in the mixed ensemble. Spectroscopic calculations revealed the formation of both H and J-type dimers, but H-type aggregations were predominant. Preparation of layer-by-layer (LbL) elec- trostatic self-assembled lms of AO and GO onto poly (allylamine hydrochloride) (PAH) coated quartz substrate is also reported in this article. UVVis absorption, steady state and time resolve uorescence and Raman spectro- scopic techniques have been employed to explore the detail photophysical properties of pure AO, AO/GO mixed solution and AO/GO LbL lms. Scanning electron microscopy was also used for visual evidence of the synthesized nanodimensional GO sheets. The uorescence quenching of AO in the presence of GO in aqueous solution was due to the interfacial photoinduced electron transfer (PET) from photoexcited AO to GO i.e. GO acts as an efcient quenching agent for the uorescence emission of AO. The quenching is found to be static in nature. Raman spec- troscopic results also conrmed the interaction of AO with GO and the electron transfer. The formation of AO/GO complex via very fast excited state electron transfer mechanism may be proposed as to prepare GO-based uorescence sensor for biomolecular detection without direct labeling the biomolecules by uorescent probe. © 2015 Elsevier B.V. All rights reserved. Keywords: Adsorption Charge transfer H-dimer TCSPC Exciton coupling Stacking 1. Introduction Atomically thin exfoliated graphite oxide or graphene oxide (GO) a non-stoichiometric, two dimensional lattice of sp 3 and sp 2 hybridized carbon network has attracted great attention in recent years [1] in the eld of Chemistry, Physics, Materials science and Nanotechnology [24]. The extraordinary properties of GO make it suitable for potential applications in electronic devices [5], solar cells [6], sensors [7], Li-ion batteries [8], supercapacitors [9], biosensing devices etc. [10]. GO bears covalently attached epoxide (1, 2, ether) hydroxyl functional groups on either side of basal plane and carboxyl groups on either side of edges [11]. Therefore, GO becomes strongly hydrophilic and easily exfo- liated in water due to electrostatic repulsion between their nanosheets forming a stable colloidal dispersions [12]. GO undergoes a complex in- terplay of various ionic and non-ionic interactions in solution due to the presence of aromatic moieties and specic functional groups [13]. Colloidal GO can be considered as a promising material with a net neg- ative surface charge because of ionizible carboxylic groups and it can be used for fabricating multilayer thin lms onto oppositely charged sur- face based on electrostatic interactions [14]. Borislav Angelov et al. re- ported the preparation of self-assembled structures of DNA and protein by Lipid Nanocarrier via electrostatic interaction [15,16]. Al- though LangmuirBlodgett technique is an elegant choice to precisely control the molecular arrangement in multilayered systems [17,18], layer-by-layer electrostatic self-assembly technique is also becoming promising as it includes wide diversity of materials to fabricate multilayered nanostructured lms onto solid substrate [19]. The excellent colloidal dispersibility and stability of GO can make it efcient nanotemplate to support wide range of organic, inorganic and biomate- rials on their surface via ππ stacking and/or electrostatic interactions [20]. Some authors reported enzyme immobilization onto the surface of GO for its large surface area [21]. Therefore, the signicant colloidal properties of GO sheets in aque- ous media may provide a platform for the preparation of ordered struc- tures of organic dye molecules namely dye aggregates. Spectroscopic Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 157 (2016) 7987 Corresponding author. E-mail addresses: [email protected], [email protected] (P.K. Paul). http://dx.doi.org/10.1016/j.saa.2015.12.006 1386-1425/© 2015 Elsevier B.V. All rights reserved. > Contents lists available at ScienceDirect Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy journal homepage: www.elsevier.com/locate/saa

Upload: others

Post on 03-Aug-2020

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy · 2016-07-29 · 80 C. Hansda et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 157

Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 157 (2016) 79–87

Contents lists available at ScienceDirect

Spectrochimica Acta Part A: Molecular and BiomolecularSpectroscopy

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

>

Layer-by-layer films and colloidal dispersions of graphene oxidenanosheets for efficient control of the fluorescence and aggregationproperties of the cationic dye acridine orange

Chaitali Hansda a,b, Utsav Chakraborty a, Syed Arshad Hussain c,Debajyoti Bhattacharjee c, Pabitra Kumar Paul a,⁎a Department of Physics, Jadavpur University, Jadavpur, Kolkata 700032, Indiab Department of Physics, The University of Burdwan, Golapbag, Burdwan 713104, Indiac Department of Physics, Tripura University, Suryamaninagar, 799022, Tripura West, India

⁎ Corresponding author.E-mail addresses: [email protected], pabitra_tu@

http://dx.doi.org/10.1016/j.saa.2015.12.0061386-1425/© 2015 Elsevier B.V. All rights reserved.

a b s t r a c t

a r t i c l e i n f o

Article history:Received 13 August 2015Received in revised form 9 October 2015Accepted 5 December 2015Available online 15 December 2015

Chemically derived graphene oxide (GO) nanosheets have received great deal of interest for technological appli-cation such as optoelectronic and biosensors. Aqueous dispersions of GO become an efficient template to inducethe association of cationic dye namely Acridine Orange (AO). Interactions of AO with colloidal GO was governedby both electrostatic and π–π stacking cooperative interactions. The type of dye aggregations was found to de-pend on the concentration of GO in the mixed ensemble. Spectroscopic calculations revealed the formation ofboth H and J-type dimers, but H-type aggregations were predominant. Preparation of layer-by-layer (LbL) elec-trostatic self-assembled films of AO and GO onto poly (allylamine hydrochloride) (PAH) coated quartz substrateis also reported in this article. UV–Vis absorption, steady state and time resolve fluorescence and Raman spectro-scopic techniques have been employed to explore the detail photophysical properties of pure AO, AO/GO mixedsolution and AO/GO LbL films. Scanning electronmicroscopywas also used for visual evidence of the synthesizednanodimensional GO sheets. The fluorescence quenching of AO in the presence of GO in aqueous solution wasdue to the interfacial photoinduced electron transfer (PET) fromphotoexcited AO toGO i.e. GO acts as an efficientquenching agent for the fluorescence emission of AO. The quenching is found to be static in nature. Raman spec-troscopic results also confirmed the interaction of AOwith GO and the electron transfer. The formation of AO/GOcomplex via very fast excited state electron transfer mechanism may be proposed as to prepare GO-basedfluorescence sensor for biomolecular detection without direct labeling the biomolecules by fluorescent probe.

© 2015 Elsevier B.V. All rights reserved.

Keywords:AdsorptionCharge transferH-dimerTCSPCExciton couplingStacking

1. Introduction

Atomically thin exfoliated graphite oxide or graphene oxide (GO) anon-stoichiometric, two dimensional lattice of sp3 and sp2 hybridizedcarbon network has attracted great attention in recent years [1] in thefield of Chemistry, Physics, Materials science and Nanotechnology[2–4]. The extraordinary properties of GO make it suitable for potentialapplications in electronic devices [5], solar cells [6], sensors [7], Li-ionbatteries [8], supercapacitors [9], biosensing devices etc. [10]. GO bearscovalently attached epoxide (1, 2, ether) hydroxyl functional groupson either side of basal plane and carboxyl groups on either side ofedges [11]. Therefore, GO becomes strongly hydrophilic and easily exfo-liated in water due to electrostatic repulsion between their nanosheetsforming a stable colloidal dispersions [12]. GO undergoes a complex in-terplay of various ionic and non-ionic interactions in solution due to thepresence of aromatic moieties and specific functional groups [13].

yahoo.co.in (P.K. Paul).

Colloidal GO can be considered as a promising material with a net neg-ative surface charge because of ionizible carboxylic groups and it can beused for fabricating multilayer thin films onto oppositely charged sur-face based on electrostatic interactions [14]. Borislav Angelov et al. re-ported the preparation of self-assembled structures of DNA andprotein by Lipid Nanocarrier via electrostatic interaction [15,16]. Al-though Langmuir–Blodgett technique is an elegant choice to preciselycontrol the molecular arrangement in multilayered systems [17,18],layer-by-layer electrostatic self-assembly technique is also becomingpromising as it includes wide diversity of materials to fabricatemultilayered nanostructured films onto solid substrate [19]. Theexcellent colloidal dispersibility and stability of GO canmake it efficientnanotemplate to support wide range of organic, inorganic and biomate-rials on their surface via π–π stacking and/or electrostatic interactions[20]. Some authors reported enzyme immobilization onto the surfaceof GO for its large surface area [21].

Therefore, the significant colloidal properties of GO sheets in aque-ous media may provide a platform for the preparation of ordered struc-tures of organic dye molecules namely dye aggregates. Spectroscopic

Page 2: Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy · 2016-07-29 · 80 C. Hansda et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 157

Fig. 1.Molecular structure of (a) AO and (b) PAH.

80 C. Hansda et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 157 (2016) 79–87

characteristics of organic dyes have shown a lot of fundamental impor-tance since long back [22]. The precise controls of these structures arealso very important because of the use of their aggregates in some bio-logical applications [23]. Specific dipole-dipole interactions between thedye monomer units result the formation of dimers or aggregates [24,25]. This is generally reflected as spectral splitting and spectral shiftingor generation of new band in the absorption spectra of dyes whencompared to that of their free monomers. Depending on the type of in-termolecular association of the monomers, the structure and opticalproperties, aggregates are classified into two major groups namelyH-type (face-to-face arrangement of the monomer) [24] and J-type(head to tail arrangement) [26]. By using simple UV–Vis absorptionspectroscopic technique one can easily characterize them. Whileconsidering the emission properties of many adsorbed fluorophoresonto GO sheets, the emission is certainly quenched due to highlyspecific interaction between fluorophores and GO and this phenome-non hasmany possible applications. For example, the GO-based fluores-cence quenching has already been demonstrated as an efficient sensingplatform for the quantitative detection of nucleotides [27]. Thequenching of fluorescence of cationic dye molecules can also be due tothe process of excited state electron transfer between dye and somesuitable electron acceptor molecules like GO either by static or dynamicquenching mechanism [28,29]. In static quenching, lifetime of theexcited fluorophore remains same whereas in dynamic quenching thelifetime decreases [30]. Furthermore, with the increasing attention toenergy and environmental issue, many researchers focus on solar cell,which has potential applications in our daily life. The electron transferafter photoexcitation of the electron acceptor moieties is a fundamentalprocess, which could strongly influence efficiency of solar cell. Thereforethe charge transfer from sensitizer (may be some cationic dye) to GO orG (electron acceptor) is of particular importance for future technologi-cal applications.

In the present article, we have addressed the detail aggregations be-havior and fluorescence quenching of a cationic dye namely AcridineOrange (abbreviated as AO) in presence of GO in aqueous media.We also report the preparation of AO/GO layer-by-layer (LbL) self-assembled mono and multilayered film onto a poly(allylaminehydrochloride) (abbreviated as PAH) coated quartz substrates viaelectrostatic interaction between the positively charged group of AOand negatively charged GO nanosheets from their aqueous solutionsand dispersion respectively. AO has a planar heterocyclic aromaticstructure which is suitable as electron donor in the excited state whencomeclose enough to anoppositely charged electron deficientmaterials[31]. AO remains in cationic form in pH below 10 and is completelydeprotonated at pH N 10 [32]. Owing to its high photostability, highphotoluminescence quantum yield and low cost, it has been widelyused as a fluorescent probe in numerous investigations [33]. In thispresent work, we have explored the detailed aggregation behaviors ofAO in a simple ion exchange strategy for electrostatic complexation aswell as π–π stacking cooperative interactions with GO and the fluores-cence quenching via photo induced electron transfer (PET)mechanism.UV–Vis absorption, steady state and time resolved fluorescence andRaman spectroscopic measurements have been employed to exploretheir molecular level interactions. AO is known to form H-aggregatesas reported by some authors [34]. However, our results confirm thepresence of some amount of J-aggregates in AO/GO mixed ensemblewith predominant H-aggregates as evidenced by using exciton couplingtheory. The aggregation behavior of AO is of particular importance be-cause of their use as molecular probe for intercalation with DNA [35].Also GO acts as an efficient quencher for the quenching of AO fluores-cence due to photoinduced electron transfer from excited AOmolecules(electron donor) to GO (acceptor). This electron transfer is a veryultrafast process than the intrinsic fluorescence of AO.We also observedthe aggregation of dye molecules in the AO/GO LbL self-assembledfilms onto the solid substrates as confirmed by UV–Vis absorptionspectroscopy.

2. Experimental

2.1. Materials

The cationic dye AO (MW= 265.35 g/mol, purity N99.9%) was pur-chased from Aldrich Chemical Company, USA and was used as received.The purity of the dye was also checked by UV–Vis absorption and fluo-rescence spectroscopy before use. The molecular structure of AO isshown in Fig. 1a. The solubility of AOwas about 0.7 mg/ml in the ambi-ent temperature and becomes light orange in color in the studied aque-ous phase at neutral pH. Natural Graphite powder was purchased fromAldrich chemical co, USA. H2SO4, NaNO3 and KMnO4 were purchasedfrom Merck Chemical Company, Germany. PAH was purchased fromPolyscience inc., USA and its molecular structure is shown in Fig. 1b.Tripled distilled deionized Milli-Q water (Resistivity 18.2 MΩ-cm) wasused to prepare all the solutions in the present work.

2.2. Synthesis of graphene oxide

GO powder was prepared according to modified Hummer's method[36], with natural graphite powder as a starting material. At first, 1 g ofgraphite powder and 0.5 g of sodium nitrate were mixed togetherfollowed by addition of 23 ml of conc. sulphuric acid under constantstirring by a magnetic stirrer. After 1 h, 3 g of KMnO4 was added tothe above solution gradually while keeping the temperature less than20 °C to prevent overheating and explosion. The mixture was thenstirred at 35 °C for 12 h and the resulting solutionwas diluted by adding500ml of distilledwater under vigorous stirring. In order to confirm thecompletion of reaction with KMnO4 the suspension was further treatedwith 30% H2O2 solution (5 ml). The Final mixture was washed off withHCl and H2O subsequently, followed by filtration and drying, grapheneoxide sheets were thus obtained. After preparing GO, it was character-ized by Field Emission Scanning Electron Microscopy (FESEM), UV–Visabsorption spectroscopy and Raman spectroscopy. FESEM image of GOas shown in Fig. 2 reveals that GO has layered sheet like structure andthey are not planner rather having wrinkled morphology. Such sheetsare folded or continuous at times. It is possible to distinguish theedges of individual sheets, including kinked and wrinkled areas.UV–Vis absorption and Raman Spectrum of GO are discussed later.

2.3. UV–Vis absorption spectroscopy

UV–Vis absorption spectroscopic measurements were done bya dual beam UV–Vis-NIR absorption spectrophotometer (Model:UV-3600PC, Make: Shimadzu, Japan). For measuring all solution spec-tra, quartz cuvette of 1 cm path length was used. Absorption spectraof LbL films deposited onto fluorescence grade quartz substrates wererecorded by this spectrophotometer at room temperature.

2.4. Steady state and time resolved fluorescence spectroscopy

All the steady state fluorescence measurements were done by aSpectrofluorometer (Model: LS-55, Make: Perkin Elmer, USA). Thefluorescence lifetime of the samples were recorded by Time-correlatedsingle photon counting (TCSPC) set up from Horiba Jobin-Yvon, Japan.

Page 3: Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy · 2016-07-29 · 80 C. Hansda et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 157

Fig. 2. FESEM image of synthesized graphene oxide (GO).Fig. 3. The UV–Vis absorption spectra of AO and AO/GO mixed solutions for various GOconcentration.

81C. Hansda et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 157 (2016) 79–87

The samples were excited at 480 nm using a picoseconds pulsed diodelaser in an IBH Fluorocube apparatus. The fluorescence decay datawere collected over 200 channels using a nonlinear time scale withthe time increment increasing according to the arithmetic progressionon a Hamamatsu MCP photomultiplier (R3809) and were analyzed byusing IBH DAS6 software.

2.5. Raman spectroscopic measurements

Raman spectra of as synthesized GO sample, AO/GO complex sampleand AO/GO LbL self-assembled film deposited onto glass substrate wererecorded in the region 800–3000 cm−1 at room temperature by using amicro Raman set-up consisting of a spectrometer (Lab RAM, HR Jovin-Yvon) and a Peltier cold CCD detector. A diode laser of wavelength785 nm was used as an excitation light source and a 10× objectivewith a numerical aperture (NA) of 0.9 was used to focus the laser onthe sample for collecting the scattered light during the experiment.

2.6. Field emission scanning electron microscopy

Field emission scanning electron microscopy (FESEM) image of theGO sample, was taken by a commercial Field Emission Scanning Elec-tron Microscope (Model: Inspect F50, Make: FEI, Checz Republic). Theoperating Voltage was set at 20 KV and emission current was 143 μAduring the measurements.

3. Results and discussion

3.1. UV–Vis absorption spectroscopy

UV–Vis absorption spectrum of pure GO in aqueous dispersion isshown Fig. S1 (see supplementary information). It exhibits two charac-teristic features that can be used as a means of identification: a maxi-mum at 231 nm corresponding to π−π⁎ transition of aromatic CCbonds and a shoulder at 300 nmwhich can be attributed to n−π⁎ tran-sitions of CO bonds [37], both are bathochromically shifted by conjuga-tion. Fig. 3 shows the UV–Vis absorption spectra of aqueous solutions ofpure AO (concentration of 1.0 × 10−5 M) and AO/GO mixed solution atdifferent concentrations of GO. Pure AO solution absorption spectrumshows distinct and prominent absorption bandswithin 450–540 nm re-gion having a strong peak at around 490 nmalongwith aweak shoulderat around 470 nm and is in well agreement with the reported results

[38,39]. The 490 nm peak is due to themonomer absorption of AOmol-ecules in solution and theweak shoulder at around 470 nm is due to thepresence of dimeric sites of AO in solution [40]. However, addition of GOdispersion into the AO solution exhibits the reduction in monomericband intensity and sequential increase in the absorption band at470 nm located at higher energy side of the monomer band indicatingthe progressive formation ofH-type dimer. That is on increase inGOdis-persion into AO solution, the number of free monomers decreases dueto their adsorption onto GO nanosheets in the mixed ensemble. This isreflected as the reduced absorbance of the monomeric band as shownin Fig. 3b. Interestingly when the concentration of GO was 100 μg/ml,the intensity of the absorption band at 470 nm increases but the spec-trum is sufficiently broadened and diffused. This is possibly due to theformation of higher order aggregates of AO inmixed solution as a resultof strong electrostatic and π-π stacking interactions between AO andGO [41]. The reactionmechanism between AO and GO for the formationof aggregates is shown in Eq. (1).

AOþ GO ¼ GO:AO½ � complexð Þ þ GO: AO:AO½ � dimerð Þ þ GO: AO :::::::AO½ �nð1Þ

Even at low concentration of GO, the presence of AO dimer or aggre-gates may be due to the columbic attraction between negativelycharged surface at the GO edges and cationic part (N+) of AO. Also,the π–π stacking co-operative interaction between the aromatic π-system of AO and the electron lone pair of surface O atom in GO sheetsmay influence the formation of dimers or aggregates. However, it is alsovery important to know about the geometry and extent of the dimericsites of the dye molecules in GO nanosheets.

The formation of AOdimers (M⇌D) in aqueous solution as a functionof GO concentration can be understood by the following equation [42]:

kD ¼ 1−x2Cx2

ð2Þ

Where, kD is defined as the equilibriumconstant for dimer formationand x is the molar fraction of free monomer at a concentration C. So kDcan be determined by using monomer molar fraction. On the otherhand, the absorptivity of AO at this concentration as a function of λ isgiven by [42]:

ε λð Þ ¼ εM λð Þxþ εD λð Þ 1−xð Þ ð3Þ

Page 4: Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy · 2016-07-29 · 80 C. Hansda et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 157

Fig. 4. Calculated dimer spectra of AO in aqueous solution with various concentrations ofGO.

82 C. Hansda et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 157 (2016) 79–87

where εM(λ) and εD(λ) are the absorptivities of themonomer and of themonomer units in the dimer respectively. The spectrum for diluted so-lutions (x ≃ 1) is the monomer spectrum, i.e. εðλÞ ¼ εMðλÞ.

Therefore, to evaluate the dimer spectrum of AOmolecules, we needto know the value ofmonomermolar fraction (x) using Eqs. (2) and (3).The knowledge of the absorptivity in the visible region at any given dyeconcentrationswith variousGO concentrations inmixed solution allowsthe use of the following iterative method. We can define a parameter Ras the ratio between the absorbance intensity of the main absorptionband (~490 nm) and that of the vibronic shoulder (~470 nm) i.e. R ¼A490

�A470

at different GO concentration in AO solution. This approximate

method based on the determination of R parameter and as a firstapproximation,we assume x(1)= Ro/R, where Ro is defined as the corre-sponding R parameter of the dye solutions at lower GO concentrations.However, this assumption is valid when the absorption capacity of thedimer εD(λ) at λmax and λsh is much lower and similar, respectively,with respect to that of the free monomer εM(λ) i.e. εD(λmax)≪εM(λmax)and εD(λsh)≈εM(λsh). Then at every GO concentration, the initial x(1)

using Eqs. (2) and (3) gives the values of approximate dimerization con-stant kD(1) and molar absorptivity of dimer. From the average values ofthese dimerization constantskD andmolar absorptivities εD, a more spe-cific molar fraction x(2) can be determined. This value is again intro-duced in the iterative process and the process is repeated to get anaccurate x(n) value. Taking into account of this method the value ofthe dimerization constant as well as transition dipole moment,oscillator strength are calculated and summarized in Table 1. UsingEqs. (2) and (3), absorption spectra of AO dimers can be evaluated formoderate GO concentration, some of which are shown in Fig. 4 fordifferent GO concentrations namely 24 μg/ml, 45 μg/ml, 55 μg/ml and65 μg/ml. The calculated absorption spectra show two well separatedabsorptions bands which are placed at both sides of the monomer ab-sorption spectrum (for very dilute concentration of AO) and this is con-sistent with the exciton coupling theory for twisted sandwich dimer oroblique head to tail dimer. However, amore precise analysis of the calcu-lated absorption spectra of AO for various GO concentrations reveals thatthe absorption band of AO is not independent of GO concentration in therange 0–55 μg/ml, rather the GO sheets are being suitable nanotemplatesfor self-association of dye molecules resulting their aggregations. Al-though the observed absorption spectra of AO solutions in presence ofGO indicate the H-type aggregates, our analysis also confirms the pres-ence of some J-type aggregates in themixedAO/GO solutions as illustrat-ed in Fig. 4. More precisely some fraction of the molecular dipoles of AOare arranged in oblique head-to-tail causing transition to lower energystate and is reflected as one of the absorption band of the calculateddimer is placed in the red side of the monomer absorption band. Thisis possibly due to the non-coplanar or wrinkled surface morphology ofGO templates [11]. However, the relative contribution of H-aggregatesis predominant over J-aggregates for the given concentration range ofGO along with overall blue shifting when compared to the monomericabsorption band of AO. Therefore the coexistence of H and J bands corre-sponds to both sandwich type and oblique head to tail dimers or aggre-gates. With regards to the calculated absorption spectra, the ratio of the

Table 1Spectroscopic parameters of AO dimers in GO sheets.

Systems Transition dipole moment (μ) λ

Pure AO monomer μM = 4.45843 λ

AO solution (concentration of 1 × 10−5 M)μ1 = 4.42829μ2 = 3.22846

λλ

AO dimers (for GO concentration of 24 μg/ml)μ1 = 4.27345μ2 = 8.53922

λλ

AO dimers (for GO concentration of 45 μg/ml)μ1 = 3.96733μ2 = 6.57447

λλ

AO dimers (for GO concentration of 55 μg/ml)μ1 = 5.08998μ2 = 6.77348

λλ

absorbance intensities (obtained from the area under the curve) of H andJ-band i.e. AH/AJ is an indicative parameter for the dominating type of ag-gregates (sandwich or head-to-tail geometry) [43]. These values aresummarized in Table 1. The progressive formation of new aggregatesconsisted of two absorption bands appeared in both sides of the mono-mer absorption spectrum of AO and this corresponds to the co-existence of H and J- dimer [44]. These splittings of the electronic energystate are caused by the dipole- dipole interactions between the mono-mer units of AO adsorbed to GO surface. Interestingly, it is observedthat the dimerization constant (kD) value increases with increase in GOconcentration in AO solution because of the formation of new higherorder aggregates along with AO dimers. This is also reflected as the dif-fused and broadened absorption spectra of AO at higher GO concentra-tion. Additionally the ionic behavior of the mixture might havechanged with increase in anionic GO concentration.

To obtain the general information about the spectroscopic parame-ters and aggregation geometry, the average absorption spectrum ofAO dimers for three different concentrations of GO in AO/GOmixed so-lution has been calculated using exciton coupling theory. Thedeconvolution of these spectra as shown in Fig. 5(a, b, c) into severalGaussian curves (by using Origin 6.5 Software) allows the adequate de-termination of absorption parameters which are given in Table 1. Thetwo main absorption bands of the dimer placed at λ1 ≈ 516 nm andλ2 ≈ 456 nm indicate that the monomer visible absorption spectrumcorresponds to an electronic transition along with two low intenseGaussian curves placed at higher energies. These two low intenseGaussians can be attributed to the corresponding vibronic transitions

Oscillator strength (f) Dimerization constant (kD) AH/AJ

M = 490 nm fM = 0.1907 – –1 = 490 nm2 = 463 nm

f1 = 0.1878f2 = 0.1062

2221 M−1 1.49

1 = 519 nm2 = 467 nm

f1 = 0.1663f2 = 0.7365

17,095 M−1 2.92

1 = 516 nm2 = 456 nm

f1 = 0.1431f2 = 0.4484

19,693 M−1 1.99

1 = 516 nm2 = 456 nm

f1 = 0.2362f2 = 0.4728

23,628 M−1 1.3

Page 5: Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy · 2016-07-29 · 80 C. Hansda et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 157

Fig. 5. Gaussian deconvolution of calculated dimer spectrum of AO solution in the presence of GO for concentration of (a) 24 μg/ml (b) 45 μg/ml (c) 55 μg/ml.

83C. Hansda et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 157 (2016) 79–87

of the twomain Gaussians. This is similar to theweak shoulder associat-ed withmain absorption band of themonomer. By applying the excitoncoupling theory the observed absorption spectra are consistent withtwisted sandwich type dimer of AO (because AH/AJ ≥ 1.3) in the mixedsolution at lower concentration range of GO as illustrated in Table 1.

Considering the allowed electric dipole transitions, the interactionpotential is of the dipole-dipole type and the resonance interaction(U) is half the energy difference of two absorption bands [39]:

U ¼ ν2−ν1

2ð4Þ

where the wave number ν is in cm−1. On the other hand the twistedangle α between transition dipole moments of the monomers can bedefined as [39]:

tan2 α=2ð Þ ¼ f 1ν2

f 2ν1: ð5Þ

Here, f1 and f2 denote the oscillator strengths of lower energy andhigher energy band of dimer respectively in absorption spectra. The in-termolecular distance R (in Å) can be derived bymeans of using follow-ing equation [39]:

R3 ¼ 2:14� 1010: cosα: f MνM 2Uð Þ : ð6Þ

The area under the curve of low and high energy bands of themonomer units in the dimer absorption spectra are referred to as theoscillator strengths f1 and f2 respectively and are given in Table 1.The geometrical parameters of AO dimers associated in GO templatesas derived from absorption spectra are summarized in Table 2. Consid-ering Eqs. (4), (5) and (6), the value of twisted angles lies between~53–74°, suggesting that the two monomer dye molecules stack ontop of each other, being significantly rotated and this corresponds tothe twisted sandwich type dimer. These results suggest that the changein GO concentration in aqueous media rather increase favorable

Page 6: Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy · 2016-07-29 · 80 C. Hansda et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 157

Table 2Geometrical parameters of AO dimers in GO templates.

Concentrationof GO

Interaction energybetween monomerunits in thedimer (U)

Angle between thetransition momentsof the monomersin dimer (α)

Intermoleculardistance in thedimer (R)

24 μg/ml 1072 cm−1 53.2° 4.81 Å45 μg/ml 1299 cm−1 62° 4.60 Å55 μg/ml 1275 cm−1 73.9° 4.73 Å

84 C. Hansda et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 157 (2016) 79–87

condition for dye aggregations which are reflected as the decreased in-termolecular distance of about 4.60 Å.

In the present work, we have also prepared alternate layer-by-layer(LbL) self-assembled films of AO and GO onto PAH coated quartz sub-strate. The method of preparation of films was described in some ofour earlier work [45]. Here cationic PAHwas used to eventually supportGO from its aqueous dispersion onto quartz substrate. On the top oftheGO surface in the substrate AOwas assembled via electrostatic inter-action from its aqueous solution (concentration of 1 × 10−5 M). Fig. 6ashows the UV–Vis absorption spectra of 1 bilayer AO/GO LbL film, pureAO solution and AO/GOmixed solution (GO concentration of 55 μg/ml).From the figure we observe that the monomer peak was reduced toa weak hump and shifted to 512 nm in case of LbL filmwhereas dimericband becomes stronger with an intense peak at around 475 nm.This dimer can be assigned to H-dimers because this dimeric bandwas also prominent in case of AO/GO mixed solution absorption spec-trum [38]. But this band is somewhat defused and broadened whencompared to the AO/GOmixed solution due to the more closer associa-tion of AO molecules in the solid state resulting some higher orderaggregates.

Fig. 6b shows the UV–Vis absorption spectra of different bilayered(1–10 bilayer) LbL films of alternate AO and GO deposited onto PAHcoated quartz substrate. It is interesting to note that absorption spectraof different layered LbLfilms showalmost similar band pattern irrespec-tive of layer number except an increase in absorption intensity. This in-dicates that AO and GO are alternately and successfully transferred onto

Fig. 6. (a) UV–Vis absorption spectra of AO solution (1 × 10−5M), AO/GO LbLfilms (1 bilayer) adifferent bilayered AO/GO LbL films deposited onto PAH coated quartz substrate. Inset shows t

LbL films during the process of film fabrication. To know the adsorptionbehavior with respect to the layer number, we have also plotted (Insetof Fig. 6b) the absorbance intensity of the band at 475 nm as a functionof layer number. From this figure we see that up to six bilayers, the ad-sorption of AO onto LbL films is more or less uniform and linear andthereafter it is non-uniform or zigzag pattern. In the process of film fab-rications, the substrates containing the terminal deposited layer eitherGO or AOwere carefully rinsedwith ultrapurewater in order to removeany loosely bound or surplus ions after deposition of each layer in LbLfilms and subsequently dried in nitrogen flow. The main driving forcein multilayer build up is the charge overcompensation during eachcycle and the surface charge overcompensation factor is merely con-stant in the first few layers (as the linear increase of the absorbance ofthe film at 475 nm) and then non-uniform. Therefore the distributionof AO molecules in GO templates in this way is more or less homoge-neous up to few initial bilayers (six bilayers) [46]. However, for thehigher layer numbers the adsorption of AO molecules in LBL films wasnot uniform. This is possibly due to the fact that the electrostatic aswell as π–π stacking interaction between AO and GO does not work aswell at such higher layer numbers because the roughness of the filmmight be increased [47].

In solid thin film, AO molecules are reached to a close proximity toGO sheet as AO/GO LbL films are fabricated in restricted geometry inwhich cationic AO interact with negatively charged GO via columbic in-teraction as well as also π–π co-operative interactions. The number ofAO molecules adsorbed to GO surface in LbL films increases with in-crease in bi-layer number resulting the increase in absorbance intensityat 475 nm.

3.2. Steady state fluorescence spectroscopy

Fig. 7a shows the steady state fluorescence emission spectra of pureAO solution (1.0 × 10−5 M), pure GO dispersion and AO/GO mixed so-lution for various GO concentrations. Pure AO aqueous solution showsa strong fluorescence emission at 525 nm when excited at 490 nm.However, after addition of GO aqueous dispersion to AO solution, thefluorescence intensity of AO decreases significantly with increasing

nd AO/GOmixed solution (GO concentration of 55 μg/ml) (b) UV–Vis absorption spectra ofhe plot of absorbance at 475 nm as a function of number of bi-layers.

Page 7: Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy · 2016-07-29 · 80 C. Hansda et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 157

Fig. 7. (a) Steady state fluorescence emission spectra of AO (1 × 10−5 M) for various concentrations of GO in the mixture along with the fluorescence emission spectrum of pure AO so-lution and GO. (b) Overlapping of absorption spectrum of GO and emission spectrum of AO in aqueous media.

Fig. 8. Fitted curves of fluorescence decay data of AO solution in presence and absence ofGO with different concentrations.

85C. Hansda et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 157 (2016) 79–87

concentration of GO in the mixture as shown in the figure. This meansthat GO can influence fluorescence quenching of AO. This reduction inemission intensity cannot be due to the inner filter effect (i.e. high ab-sorption of excitation light and reabsorption phenomenon) for this con-centration because of the absence of emission peak of GO in the sameexcitation wavelength (490 nm) of pure AO (Fig. 7a). Such strongquenching of fluorescence intensity of AO in presence of GO may bedue to some nonradiative transition either by energy transfer or elec-tron transfer whichmay depend on the nature of fluorophore, conjuga-tionmode and variousmicroenvironments [48]. Evidently there was nospectral overlap of absorption band of GO and emission band of AO asshown in Fig. 7b; thus fluorescence quenching of dyes in presence ofGO can be due to an excited state interfacial electron transfer processsince we did not observe any ground state complex formation [29]. Ad-ditionally, there was no appreciable metachromasy effect of dyes intheir absorption spectra in presence of GO. Moreover the monomerpeak with respect to the dimer peak is placed at lower energy side intheir excitation spectra (figure not shown). This can be attributed tothe configuration of twisted sandwich type aggregates as previouslyconcluded by absorption spectroscopy. The observed fluorescencequenching here corresponds to the processes: at first the dyemoleculesadsorbed onto the GO surface, secondly the electrons move fromthe excited state dyes to the surface of GO. Using the equationE(eV)=−4.5− ENHE (eV) [49–51], thework function of AO can be cal-culated as−2.96 eV with respect to normal hydrogen electrode (NHE).On the other hand, potential of GO conduction band is about −3.5 eVwhich is lower than the excited singlet state energy of dye molecules[52]. So, the electron could transfer from excited state dyes to the con-duction band of GO. Molecular orbital theory and experimental resultsalso suggest that closed cage carbon structure such as fullerenes andCNTs are favorable electron acceptor because of their electron deficien-cy property while partially oxidized GO is no exception [53].

3.3. Fluorescence lifetime measurements by Time-Correlated Single-PhotonCounting method (TCSPC)

Fluorescence lifetime study is useful to understand the type ofquenching in aromatic systemswhether it is dynamic or static in naturecaused by electron transfer mechanism. In the present work, to confirm

the probable fluorescence quenchingmechanismof AO after addition ofGO into their solution, florescence lifetime measurements were carriedout by Time correlated single photon counting method (TCSPC) for AOin absence and presence of GO at various concentrations namely, 12,24 and 45 μg/ml. The solutions were excited at 480 nm (as nanoledsources arewavelength specific) and corresponding emissionwasmon-itored at 525 nm. Also we examined that there is no appreciable changeof emission intensity (steady state) when excited at 490 nm. Thefluorescence decay curves for pure AO aqueous solutions along withAO/GO mixed solution for various GO concentrations are fitted asmono and biexponential respectively and are shown in Fig. 8. The de-tailed decay information are summarized in Table 3. We immediatelyobserve two lifetime values of AO in presence of GO in aqueousmediumand the process is so ultrafast that it is very hard to distinguish theirchanges in the given timescale. The shorter lifetime values typically rep-resent the charge injection from AO to GO i.e. bound states of AOwhereas longer lifetime values is attributed to the charge relaxed or unboundstates of AOmolecules [54]. More specifically the shorter lifetime values

Page 8: Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy · 2016-07-29 · 80 C. Hansda et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 157

Table 3Time correlated single photon counting (TCSPC) fluorescence decay results of AO andAO/GO mixed solution.

Systems Exponential τ1(ns)

τ2(ns)

bτ N

(ns)χ2 λex

(nm)

Pure AO solution(10−5 M) 1 1.8 – 1.8 1.055 480AO/GO(12 μg/ml) mixedsolution

2 0.949 1.827 1.87 1.094 480

AO/GO(24 μg/ml) mixedsolution

2 0.983 1.874 1.94 1.160 480

AO/GO(45 μg/ml) mixedsolution

2 0.454 1.8 2.39 1.040 480

Fig. 9. Raman spectra of synthesized GO, AO/GO complex and AO/GO LbL self-assembledfilm deposited onto glass substrate.

86 C. Hansda et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 157 (2016) 79–87

(τ1) as shown in Table 3 correspond to the quenched states of AO mol-ecules adsorbed onto GO surface due to photoinduced electron transfermechanism. This implies that AO molecules (as electron donor) areclose enough to GO (as electron acceptor) surface by electrostatic inter-action and/or π–π co-operative stacking interaction so that photoin-duced electron transfer (PET) occurred to form AO/GO charge transfercomplex and PET rate is much larger than the intrinsic fluorescenceemission rate. As a result fluorescence intensity of AOmolecules in pres-ence of GO drastically reduced. However, it is interesting that theshorter lifetime values of AO with GO for concentrations of 12 μg/mland 24 μg/ml are merely same but it is much reducedwhen GO concen-tration was 45 μg/ml. In fact, at this higher GO concentration in themixed solution, the mean distance between photoexcited AO⁎ mole-cules andGOnanosheets decreases [55] causing a very ultrafast electrontransfer process which is manifested as much reduced lifetime (τ1) ofadsorbed AO species in the excited states. Furthermore, there was suffi-cient increase in number of dimers or higher order aggregates at suchhigh GO concentration as confirmed by absorption spectroscopy andtherefore the intermolecular vibrational coupling also influence theoverall lifetime of the AO species in the mixture.

From the lifetime data we observe that the longer lifetime compo-nent (τ2) of the fluorophore in presence of GO is almost constant onlytheir population increases. So the quenching mechanism can be staticin nature. The lifetime of AO molecules in absence of GO was found tobe 1.8 ns as expected [56]. However if we look into the average lifetimeof AO fluorophore in presence of GO, we immediately see that it isslightly increased for higher GO concentrations as summarized inTable 3. This increase in average lifetime corresponds to the decreasein number of free monomers as well as some contribution of J-dimersor its higher order aggregates. The average lifetime was estimatedby the equation bτN = A1τ1 + A2τ2, where A1 and A2 are the pre-exponential factors corresponding to the lifetime associated with eachexponential decay.

3.4. Raman spectroscopy

To confirm the AO/GO complexation, Raman spectroscopy has alsobeen used in the present work. The changes in electronic structure ofGO occurred as a result of their interaction with AO (donor) molecules.Fig. 9 shows the normalized Raman spectra of pure GO, AO/GO complex(concentrations of GO and AOwere 0.05mg/ml and 1 × 10−5 M respec-tively) and one layered AO/GO complex LbL film. Raman signal of pureAO in the present experimental method was very low and noisy(Figure not shown) after excitation under normal Raman spectroscopy.From the figure, it is observed that pure GO has its two well resolvedcharacteristic bands, called G band and D band located at 1592 cm−1

and 1332 cm-l respectively and is consistent with the results reportedelsewhere [57]. The observed D band in GO basically arises due tobreathings modes K-point phonons of A1g for sp3 carbon atoms andthis indicates the formation of defects associated with vacancies andgrain boundaries [58]. On the other hand, G band corresponds to theC–C stretching mode vibrations of sp2 carbon lattice [58,59]. The ratioof the intensities of D and G band is related to the microstructure of

GO. The additional peaks in the range 2500–3000 cm−1 are due to thesecond order disorder mode (2D) of nanodimenasional GO [60]. Afteraddition of AO into GO in solution G band peak position was shifted to1617 cm−1 with an increase of their relative intensity (compared toD band), whereas there was no significant changes of peak position ofD band except the change in their intensity distribution. The shiftingin G band peak position may be attributed due to the molecular chargetransfer between aromatic molecules containing electron withdrawingand electron donating groups and GO. These groups can modify theelectronic structure of GO by stiffening or softening of phonons whichgive rise to the shift of Fermi energy level [61,62]. Here in case ofAO/GO mixed solution the G band shifted to higher frequency regioni.e. stiffening of G band is occurredwhich is basically due to thepresenceof electron donating group in AO. So the electron is transferred fromdyemolecule to GO. Again in case of AO/GO LbL films the D band andG bandpeak position observed at 1336 cm−1 and 1629 cm−1 respectivelyalong with a generation of a new band at around 2060 cm−1 becauseof formation of defects in GO. These results also indicate the complexa-tion of GOwith AO in LbL films and the change is possibly due to the dif-ferent local microenvironment than AO/GO mixed solution and alsophoto-excited charge transfer between AO and GO. Therefore Ramanspectroscopic results also successfully confirm AO/GO complexation insolution as well as in LbL film as was already evidenced by UV–Vis ab-sorption and steady state fluorescence spectroscopy.

4. Conclusions

In conclusion, the present study deals with the interactions of chem-ically synthesized GOwith cationic dye AOboth in aqueous solution andin LbL self-assembled film deposited onto solid substrate. The UV–Visabsorption spectroscopic studies reveal the formation of H-type aggre-gates of AO in AO/GOmixed solution. Butmore precise and quantitativeanalysis has confirmed that some J-type aggregates are also present inthe mixture. However, H-aggregates are predominant in the mixed so-lution as evidenced by absorption spectra. With increase in GO concen-tration, the absorption spectra of AO were diffused and broadenedindicating progressive formation of higher order aggregates and so thedimerization constant (kD) for AO molecules gradually increased. InAO/GO LbL films, the dimeric band is more intense and monomerband has been reduced to weak hump. The multilayer growth of LbL

Page 9: Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy · 2016-07-29 · 80 C. Hansda et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 157

87C. Hansda et al. / Spectrochimica Acta Part A: Molecular and Biomolecular Spectroscopy 157 (2016) 79–87

films onto quartz substrate was uniform up to six bilayers and then itwas non-uniform or zigzag pattern. Steady state and time resolved fluo-rescencemeasurements have confirmed that GO behaves as an efficientnanoquencher for the fluorescence of AO in aqueous solution andquenchingmechanism is static in nature showing almost constant fluo-rescence lifetime after addition of GO sample to AO aqueous solution.However for higher GO concentration the slight increase in fluorescencelifetime can be due to the presence of J-type aggregates in the AO/GOmixed solution. The fluorescence quenching of AO is mainly due to theformation AO/GO complex via photo-induced electron transfer (PET)from AO to the surface of GO. The electron transfer corresponds tosuch an ultrafast process that it is much faster than the intrinsic fluores-cence of AO as confirmed by fluorescence lifetime measurements.Raman spectroscopy also confirms the complexation of AO with GOnanotemplates both in solution as well as in LbL films because G bandof GO was shifted due to electron doping. Finally, GO has induced thetemplated assembly of cationic dye AO and also can strongly quenchtheir florescence intensity. The association and strong fluorescencequenching of cationic dye molecules onto GO nanotemplate may beproposed as a probe for biomolecular recognition as well as to detectany local microenvironment.

Supplementary data to this article can be found online at http://dx.doi.org/10.1016/j.saa.2015.12.006.

Acknowledgments

Authors are grateful to Department of Science and Technology, Govt.of India for DST-FIST Project (Level-II) for providing excellent infrastruc-tural and equipment facilities to the Department of Physics at JadavpurUniversity to perform this work. Dr. P.K. Paul is especially thankful toProf. Sujoy Baitalik andMr. Shrikanta Karmakar of Department of Chem-istry, Jadavpur University for providing access to fluorescence lifetimemeasurement set upwith valuable discussions related to this work. Au-thors are also thankful to Mr. Sumit Majumder of SINP, Kolkata for ex-perimental assistance of preparing graphene oxide and Mr. ShibShankar Singha, Bose Institute, Kolkata for his assistance in Raman Spec-troscopic measurements.

References

[1] C. Liu, F. Hao, X. Zhao, Q. Zhao, S. Luo, H. Lin, Nat. Sci. Rep. 4 (2014) 1–6.[2] A.M. Dimiev, J.M. Tour, ACS Nano 8 (2014) 3060–3068.[3] D.R. Dreyer, S. Park, C.W. Bielawski, R.S. Ruoff, Soc. Rev. 39 (2010) 228–240.[4] A.Y. Romanchuk, A.S. Slesarev, S.N. Kalmykov, D.V. Kosynkin, J.M. Tour, Phys. Chem.

Chem. Phys. 15 (2013) 2321–2327.[5] C.-C. Lin, H.-Y. Wu, N.-C. Lin, C.-H. Lin, Jpn. J. Appl. Phys. 53 (2014) 05FD03.[6] Y. Chen, W.-C. Lin, J. Liu, L. Dai, Nano Lett. 14 (2014) 1467–1471.[7] S. Borini, R. White, D. Wei, M. Astley, S. Haque, E. Spigone, N. Harris, J. Kivioja, T.

Ryhänen, ACS Nano 7 (2013) 11166–11173.[8] J. Chang, X. Huang, G. Zhou, S. Cui, P.B. Hallac, J. Jiang, P.T. Hurley, J. Chen, Adv.Mater.

26 (2014) 665–815.[9] K. Gopalakrishnan, A. Govindaraja, C.N.R. Rao, J. Mater. Chem. A 1 (2013)

7563–7565.[10] L. Yong, D. Yu, C. Zeng, Z. Miao, L. Dai, Langmuir 26 (2010) 6158–6160.[11] D.A. Dikin, S. Stankovich, E.J. Zimney, R.D. Piner, G.H. Dommett, B.G. Evmenenko, S.T.

Nguyen, R.S. Ruoff, Nature 448 (2007) 457–460.[12] S. Park, J. An, R.D. Piner, I. Jung, D. Yang, A. Velamakanni, S.T. Nguyen, R.S. Ruoff,

Chem. Mater. 20 (2008) 6592–6594.[13] Y.Y. Liang, D.Q. Wu, X.L. Feng, K. Mullen, Adv. Mater. 21 (2009) 1679–1683.[14] T. Cassagneau, F. Guerin, J.H. Fendler, Langmuir 16 (2000) 7318–7324.

[15] B. Angelov, A. Angelova, S.K. Filippov, T. Narayanan, M. Drechsler, P. Štepanek, P.Couvreur, S. Lesieur, J. Phys. Chem. Lett. 4 (2013) 1959–1964.

[16] B. Angelov, A. Angelova, S.K. Filippov, M. Drechsler, P. Stepanek, S. Lesieur, ACS Nano8 (2014) 5216–5226.

[17] A. Angelova, J. Reiche, R. Ionov, D. Janietz, L. Brehmer, Thin Solid Films 242 (1994)289–294.

[18] A. Angelova, F. Penacorada, B. Stiller, T. Zetzsche, R. Ionov, H. Kamusewitz, L.Brehmer, J. Phys. Chem. 98 (1994) 6790–6796.

[19] F.N. Crespilho, V. Zucolotto, O.N. Oliveira Jr., F.C. Nart, Int. J. Electrochem. Sci. 1(2006) 194–214.

[20] Y. Xu, L. Zhao, H. Bai, W. Hong, C. Li, G. Shi, J. Am. Chem. Soc. 131 (2009)13490–13497.

[21] J. Zhang, F. Zhang, H. Yang, X. Huang, H. Liu, J. Zhang, S. Guo, Langmuir 26 (2010)6083–6085.

[22] B. Angelov, A. Angelova, R. Ionov, J. Phys. Chem. B 104 (2000) 9140–9148.[23] B.A. Armitage, Top. Curr. Chem. 253 (2005) 55–76.[24] A. Eisfeld, J.S. Briggs, Chem. Phys. 324 (2006) 376–384.[25] A. Angelova, C. Fajolles, C. Hocquelet, F. Djedaïni-Pilard, S. Lesieur, V. Bonnet, B.

Perly, G. Lebas, L. Mauclaire, J. Colloid Interface Sci. 322 (2008) 304–314.[26] A. Angelova, D. Vollhardt, R. Ionov, Thin Solid Films 284 (1996) 85–89.[27] S. He, B. Song, D. Li, C. Zhu, W. Qi, Y. Wen, L. Wang, S. Song, H. Fang, C. Fan, Adv.

Funct. Mater. 20 (2010) 453–459.[28] T. Tachikawa, S.-C. Cui, M. Fujitsuka, T. Majima, Phys. Chem. Chem. Phys. 14 (2012)

4244–4249.[29] Y. Pang, Y. Cui, Y. Ma, H. Qian, X. Shen, Micro Nano Lett. 7 (2012) 608–612.[30] S. Doose, H. Neuweiler, M. Sauer, ChemPhysChem 10 (2009) 1389–1398.[31] S. De, A. Girigoswami, J. Colloid Interface Sci. 271 (2004) 485–495.[32] R.D. Falcone, N.M. Correa, M.A. Biasutti, J.J. Silber, Langmuir 18 (2002) 2039–2047.[33] Y. Moriyama, T. Takano, S. Ohkuma, J. Biochem. 92 (1982) 1333–1336.[34] C. Peyratout, E. Donath, L. Daehne, J. Photochem. Photobiol. A Chem. 142 (2001)

51–57.[35] S. Siddiquee, N.A. Yusof, A.B. Salleh, S.G. Tan, F.A. Bakar, Microchim. Acta 172 (2011)

357–363.[36] S.D. Perera, R.G. Mariano, K. Vu, N. Nour, O. Seitz, Y. Chabal, K.J. Balkus Jr., ACS Catal.

2 (2012) 949–956.[37] D. Zhang, X. Liu, X. Wang, J. Inorg. Biochem. 105 (2011) 1181–1186.[38] J. Bhattacharjee, S.A. Hussain, D. Bhattacharjee, Spectrochim. Acta A 116 (2013)

148–153.[39] L. Antonov, G. Gergov, V. Petrov, M. Kubista, J. Nygren, Talanta 49 (1999) 99–106.[40] X.-Z. Feng, Z. Lin, L.-J. Yang, C. Wang, C. Bai, Talanta 47 (1998) 1223–1229.[41] M. Şinoforoğlu, B. Gür, M. Arık, Y. Onganer, K. Meral, RSC Adv. 3 (2013)

11832–11838.[42] I.L. Arbeloa, J. Chem. Soc. Faraday Trans. 277 (1981) 1725–1733.[43] T. Fujii, H. Nishikiori, T. Tamura, Chem. Phys. Lett. 233 (1995) 424–429.[44] H. Nishikiori, T. Fujii, Phys. Chem. B 101 (1997) 3680–3687.[45] P.K. Paul, S.A. Hussain, D. Bhattacharjee, M. Pal, Chin. J. Chem. Phys. 24 (2011)

348–352.[46] Z. Li, J. Wang, X. Liu, S. Liu, J. Ou, S. Yang, J. Mater. Chem. 21 (2011) 3397–3403.[47] S.A. Hussain, P.K. Paul, D. Dey, D. Bhattacharjee, S. Sinha, Chem. Phys. Lett. 450

(2007) 49–54.[48] M. Miguel, M. Álvaro, H. García, Langmuir 28 (2012) 2849–2857.[49] Y. Liu, C. Liu, Y. Liu, Appl. Surf. Sci. 257 (2011) 5513–5518.[50] A. Fujishima, T.N. Tao, D.A. Tryk, J. Photochem. Photobiol. C 1 (2000) 1–21.[51] R. Czerw, B. Foley, D. Tekleab, A. Rubio, P.M. Ajayan, D.L. Carroll, Phys. Rev. B 66

(2002) 033408–033412.[52] T.F. Yeh, J.M. Syu, C. Cheng, T.H. Chang, H.S. Teng, Adv. Funct. Mater. 20 (2010)

2255–2262.[53] P.W. Fowler, A. Ceulemans, J. Phys. Chem. 99 (1995) 508–510.[54] K. Ray, H. Nakahara, J. Phys. Chem. B 106 (2002) 92–100.[55] K. Fan, Z. Guo, Z. Geng, J. Ge, S. Jiang, J. Hu, Q. Zhang, Chin. J. Chem. Phys. 26 (2013)

252–258.[56] M.S. Chan, J.R. Bolton, Photochem. Photobiol. 34 (2008) 537–554.[57] J. Li, X. Zeng, T. Ren, E.V.D. Heide, Lubricants 2 (2014) 137–161.[58] K.N. Kudin, B. Ozbas, H.C. Schniepp, R.K. Prudhomme, I.A. Aksay, R. Car, Nano Lett. 8

(2008) 36–41.[59] A.C. Ferrari, Solid State Commun. 143 (2007) 47–57.[60] X. D'ıez-Betriu, S. Alvarez-Garcıa, C. Botas, P. Alvarez, J. Sanchez-Marcos, C. Prieto, R.

Menendez, A. Andres, J. Mater. Chem. C 1 (2013) 6905–6912.[61] B. Das, R. Voggu, C.S. Rout, C.N.R. Rao, Chem. Commun. (2008) 5155–5157.[62] R. Voggu, C.S. Rout, A.D. Franklin, T.S. Fisher, C.N.R. Rao, J. Phys. Chem. C 112 (2008)

13053–13056.