Download as pdf or txt
Download as pdf or txt
You are on page 1of 8

View Online / Journal Homepage / Table of Contents for this issue

PAPER

www.rsc.org/softmatter | Soft Matter

Aerogels from nanobrillated cellulose with tunable oleophobicity


Christian Aulin,ab Julia Netrval,b Lars W agberg*b and Tom Lindstr omc
Received 28th January 2010, Accepted 19th April 2010 First published as an Advance Article on the web 14th May 2010 DOI: 10.1039/c001939a The formation of structured porous aerogels of nanobrillated cellulose (NFC) by freeze-drying has been demonstrated. The aerogels have a high porosity, as shown by FE-SEM and nitrogen adsorption/ desorption measurements, and a very low density (<0.03 g cm3). The density and surface texture of the aerogels can be tuned by selecting the concentration of the NFC dispersions before freeze-drying. Chemical vapor deposition (CVD) of 1H,1H,2H,2H-peruorodecyltrichlorosilane (PFOTS) was used to uniformly coat the aerogel to tune their wetting properties towards non-polar liquids. An XPS analysis of the chemical composition of the PFOTS-modied aerogels demonstrated the reproducibility of the PFOTS-coating and the high atomic uorine concentration (ca. 51%) in the surfaces. The modied aerogels formed a robust composite interface with high apparent contact angles (q* [ 90 ) for castor oil (glv 35.8 mN m1) and hexadecane (glv 27.5 mN m1).

Downloaded by Nanyang Technological University on 31 May 2012 Published on 14 May 2010 on http://pubs.rsc.org | doi:10.1039/C001939A

Introduction
Aerogels are materials prepared by replacing the liquid solvent in a gel by air without substantially altering the network structure or the volume of the gel body.1 The rst aerogels were reported by Kistler in 19311932,2,3 but active research in this area did not start until about 40 years later. Some of the unique properties of aerogels are their low density, high specic surface area, low thermal conductivity and low dielectric permittivity. The preparation, physical properties and applications of aerogels are described elsewhere.4,5 The liquid in a wet gel is usually replaced with air by using supercritical drying, but ambient-pressure drying has also been attempted.4,6 From a practical point of view, the challenge has been to prepare aerogels without supercritical drying in order to reduce the cost. An alternative can be freezedrying79 where the solvent in the gel is rst frozen and then sublimated without entering the liquid state. The most common aerogels are inorganic, prepared by solgel polymerization of inorganic metal oxides.1014 They are usually very brittle, but have a high compressive strength. Various types of organic aerogels have also been presented.1517 Resorcinol/formaldehyde (RF) and melamine/formaldehyde (MF) are two of the most common precursor mixtures for forming the organic network.1719 In the context of a sustainable society, there is strong motivation to replace petroleum-based polymers with polymers from renewable resources. Porous materials with nano- and microsized pores made from natural polymers are of special interest for medical, cosmetic, pharmaceutical, and other applications where biocompatibility and biodegradability are required.2022 Threedimensional scaffolds for tissue engineering, delivery matrices, green packaging, and environment-friendly insulating materials are examples of such applications. Among the polysaccharides, cellulose has a special potential as one of the most
BIM Kemi AB, Box 3102, SE-443 03 Stenkullen, Sweden Department of Fibre and Polymer Technology, School of Chemical Science and Engineering, The Royal Institute of Technology, SE-100 44 Stockholm, Sweden. E-mail: wagberg@kth.se c Innventia AB, Box 5604, SE-114 86 Stockholm, Sweden
b a

abundant renewable natural polymers, widely used in industry. In plant cellulose, the polysaccharide chains with b-(14)-D-glucopyranose repeating units pack into long brils with crosssectional dimension of ca. 530 nm, depending on the plant source.23 The high modulus and high strength of native cellulose I are results of the organization of the cellulose chains in a crystal structure in which the long parallel polysaccharide chains are physically bonded together by a large number of hydrogen bonds and are organized in sheets packed in a parallel-up fashion.24 It is interesting and challenging to consider the long nanobers as construction units for nanoscale material engineering. The term nanobrillated cellulose (NFC) refers to cellulosic I brils disintegrated from the plant cell walls. The preparation of NFC derived from wood, and specication of the term, were rst described by Turbak et al.25 and Herrick et al.26 more than two decades ago. Through a homogenization process, wood pulp is disintegrated, to give a material in which the bres are degraded and opened into their sub-structural units. New methods for the manufacture of smaller and more homogeneous nanobrillated cellulose (NFC) have recently been developed and offer new attractive concepts for material science and signicantly enhance the applicability of cellulose in novel applications.2730 Recently, NFC was prepared by a combination of mechanical and enzymatic pretreatment followed by high-pressure homogenization31 or by high-pressure homogenization of carboxymethylated cellulose bers followed by ultrasonication and centrifugation.32 The latter carboxymethylation pretreatment makes the brils highly charged and easier to liberate, and this results in slightly smaller and more uniform bril dimensions than in the enzymetreated NFC.33 Since the brils are 520 nm thick and have a length of up to several mm, they can be regarded as nanobres. Cellulose nanobrils show very interesting properties as reinforcement elements in polymer nanocomposites,3436 but a surface modication of the nanobrils might further widen the applications since they can be incorporated into new bio-based materials with a tuned interaction potential. In this respect, a hydrophobation of the aerogels is a very interesting research area.
This journal is The Royal Society of Chemistry 2010

3298 | Soft Matter, 2010, 6, 32983305

View Online

The need for surface hydrophobicity has become a hot research topic which has stimulated a rich variety of studies describing different approaches for material engineering.3741 Several studies have also been conducted on the chemical modication/hydrophobation of nanocellulose.30,42,43 The most effective treatments use chemical or physical means to append uorinated moieties which induce both hydrophobic and oleophobic properties to the resulting low-energy surfaces. The presence of these moieties also tends to impart a high thermal stability and reduce chemical and biological fragility. It is highly desirable for superhydrophobic surfaces also to be oil-repellent in order to maintain their superhydrophobicity. For instance, in an industrial or household environment, a superhydrophobic surface with poor oil repellency can easily be contaminated by oily substances, and this will compromise the superhydrophobicity of the surface. Therefore, superoleophobic surfaces combining both superhydrophobic and superoleophobic properties are desirable for many practical applications.44,45 Modifying the cellulose aerogel network with uorinated organic compounds could widen their applications towards selfcleaning surfaces in a vast array of products, including green constructions, packaging materials, protection against environmental fouling, sports and outdoor clothing, and microuidic systems. The use of cellulose-based materials could thus be extended to new areas by the introduction of functional moieties onto the ber surface. Despite extensive investigations on superhydrophobic surfaces, studies on superoleophobic surfaces with high repellency against liquids with low surface tension (<35 mN m1) have so far been rather limited. Oil repellency has been examined on various peruorinated, superhydrophobic surfaces.4648 Very recently, truly superoleophobic surfaces have been achieved on the basis of etched silicon surfaces45 with nanonail49 structures, exemplied by low contact angle hysteresis for probe liquids of low surface tension (<30 mN m1), such as octane. In both cases, the key to obtaining true superoleophobicity is the re-entrant or overhang surface structure, which ensures the entrapment of air beneath the top solid surface and prevents the transition from the Cassie Baxter state to the Wenzel state.45,49,50 However, the fabrication of most superoleophobic surfaces involves lithography and etching steps, and this may limit their practical applications. In the present work, a novel route for the production of superoleophobic NFC aerogels is used based on aerogel skeletons formed by long and entangled cellulose I containing nanoscopic brils. It gives an aerogel with sufcient strength to oppose its tendency to collapse during the solvent extraction. The aerogel network was modied with the aid of chemical vapor deposition of a uorinated silane. By adjusting the concentration of the NFC dispersion before freeze-drying, the surface texture of the aerogels was altered, making it possible to optimize the degree of oleophobicity.

dissolving plus) was rst dispersed in dissolving pulp (Domsjo deionized water at 10 000 revolutions in an ordinary laboratory reslusher. The bers were then solvent-exchanged to ethanol by washing the bers in ethanol four times with an intermediate ltration step. The bers were then impregnated for 30 min with a solution of 10 grams of monochloroacetic acid in 500 mL of isopropanol. This carboxymethylation reaction was allowed to continue for 1 h. Following the carboxymethylation step, the bers were ltered and washed in three steps: rst with deionized water, then with acetic acid (0.1 M), and nally with deionized water. The bers were then impregnated with a NaHCO3 solution (4 wt% solution) in order to convert the carboxyl groups to their sodium form to further enhance the delamination of the bers into nanobrils. Finally, the bers chner were washed with deionized water and drained on a Bu funnel. After this treatment, the bers were passed through a high-pressure homogenizer (Microuidizer M-110EH, Mircouidics Corp). Cellulose slurries containing a 2 wt% pulp bre suspension in deionized water were processed through the homogenizer. Such a procedure lead to the liberation of cellulose I nanobers, mostly with cross-sectional diameters of 520 nm and lengths of a few micrometres, although some larger entities were formed. The so prepared NFC is hence a dispersion of nanobers. NFC dispersions with concentrations of ca. 0.0032.6 wt% were prepared by diluting the 3.13 wt% NFC dispersion with deionized water followed by mixing (8000 rpm) using an Ultra Turrax mixer (IKA D125 Basic, Germany) for 5 min. The total charge density of the highly carboxymethylated NFC dispersions was measured to be 627 meq. g1 (ref. 32) by conductometric titration.51 The degree of substitution was measured to be 0.1. The surface charge density, measured by polyelectrolyte titration52 using poly-DADMAC (Ciba, Yorkshire, UK, Mw 440 000 g mol1 and 3 6.19 meq. g1), was measured to be 426 meq. g1. Cylindrical PDMS cups (48 mm in diameter, 16 mm in height) were used as moulds for the preparation of the aerogels. The aqueous gel was placed in the mould and the mould was plunged into liquid nitrogen. Thereafter, the frozen sample in the mould was transferred to a vacuum oven at 52  C (Labconco FreeZone 6, US), and the sample was kept frozen during the drying at a pressure of ca. 0.016 mbar. The drying was typically nished within 24 h. The density of the aerogels was obtained as the mass divided by the volume (4.9 cm3) of the samples. A free-standing lm of NFC was prepared by pouring a 0.1 wt% NFC dispersion onto polystyrene Petri dishes with a diameter of 14 cm. The lm was allowed to form upon drying at a temperature of 23  C and a relative humidity (RH) of 50%. The dried free-standing lm was stored in a desiccator prior to analysis.

Downloaded by Nanyang Technological University on 31 May 2012 Published on 14 May 2010 on http://pubs.rsc.org | doi:10.1039/C001939A

Modication by peruoroalkylsilane

Experimental
Preparation of NFC aerogels and lms The anionic NFC used in this study was prepared by a procedure similar to that previously described31 but using a carboxymethylation32 pretreatment of the bers. In brief, the
This journal is The Royal Society of Chemistry 2010

The silane treatment was carried out by chemical vapor deposition of 1H,1H,2H,2H-peruorodecyltrichlorosilane (PFOTS) (97%, Sigma Aldrich). Samples were placed on a copper grid located 5 cm above a beaker containing the uorinated silane, which was heated at 140  C for one hour. The samples were stored in a desiccator prior to analysis.
Soft Matter, 2010, 6, 32983305 | 3299

View Online

Scanning electron microscopy To study the micro-structure of the NFC aerogels, the specimens were studied with a Hitachi S-4800 eld emission scanning electron microscope (FE-SEM) to obtain secondary electron images. The specimens were xed on a metal stub with colloidal graphite paint and coated with a 6 nm thick gold/palladium layer using a Cressington 208HR High Resolution Sputter Coater. Nitrogen adsorption/desorption measurements The specic surface areas were determined by N2 adsorption/ desorption measurements at the temperature of liquid nitrogen (ASAP 2020, Micromeritics, US). Before measurement, the samples were dried at a temperature of 117  C until a vacuum of <105 mmHg was reached. Both adsorption and desorption isotherms were measured and the surface area was determined from the adsorption results using the BrunauerEmmetTeller (BET) method. Contact angle measurements A CAM 200 (KSV Instruments Ltd, Helsinki, Finland) contact angle goniometer was used for advancing contact angle measurements. The software delivered by the instrument manufacturer calculates the contact angle on the basis of a numerical solution of the full YoungLaplace equation. Measurements were performed at room temperature with two non-polar, low surface tension probe liquids: castor oil (Sigma Aldrich) and hexadecane (>99%, anhydrous, Sigma Aldrich). The contact angle was determined at three different positions on each sample. The values reported were taken after the contact angle had reached a stable value, typically less than 10 s after deposition of the droplet. Typical uncertainties in the experiments were 4 . X-Ray photoelectron spectroscopy (XPS) The XPS spectra were collected with a Kratos Axis Ultra DLD electron spectrometer (UK) using a monochromated Al Ka source operated at 150 W, with a pass energy of 160 eV for wide spectra and a pass energy of 20 eV for individual photoelectron lines. The surface potential was stabilized by the spectrometer charge neutralization system. Photoelectrons were collected at a take-off angle of 90 relative to the sample surface and the depth of analysis was ca. 10 nm. The binding energy (BE) scale was referenced to the C 1s line of aliphatic carbon, set at 285.0 eV. The spectra were processed with the Kratos software and the CasaXPS program package including experimental values for the atomic sensitivity factors, and peak intensities were determined by integrating the areas under the peaks.
Fig. 1 Example of cylindrically shaped aerogels obtained from freezedrying of (a) 0.7 and (b) 1.1 wt% NFC dispersions.

Downloaded by Nanyang Technological University on 31 May 2012 Published on 14 May 2010 on http://pubs.rsc.org | doi:10.1039/C001939A

demonstrates the macroscopic integrity achieved, as specimens with well-dened shapes were prepared. The densities of the samples were very low. Fig. 2 shows the density of the aerogels as a function of the initial NFC dispersion concentration. The aerogel density was almost linearly proportional to the dispersion concentrations. For example, aerogels prepared from 0.5 and 3.13 wt% dispersions resulted in material densities of 0.0053 and 0.030 g cm3, respectively. The aerogel prepared from the 0.5 wt% dispersion had a very high porosity, ca. 99.7%, where the porosity F is dened as F 1 (r/rs), where r and rs (1.63 g cm3)53 are the densities of the aerogel and the crystalline Ib cellulose bril, respectively. For comparison, the aerogel prepared from the 3.13 wt% dispersion had a porosity of ca. 98.2%. The specic surface area was analyzed by N2 adsorption/ desorption at 77 K. Aerogels with densities of 0.030 and 0.020 g cm3 showed BET specic surface areas of 11 and 15 m2 g1, respectively. Evidently, the BET-area increased with decreasing density of the aerogels as expected. These values were lower than et al.,20 who reported a value of 66 in a previous study by P a akko 2 1 m g for low-charged NFC aerogels prepared from a 2 wt% NFC-dispersion. As previously shown in detail under aqueous conditions, the initial wet gel, i.e. NFC in water, consists of long and entangled brils with diameters of ca. 5 nm with occasional thicker bril bundles.31 Fig. 3 presents FE-SEM micrographs showing the surface texture of aerogels with densities of 0.00027, 0.0018, 0.0053, 0.0070, 0.011 and 0.030 g cm3, respectively. As the

Results
Structural characteristics of NFC aerogels Aerogels from NFC dispersions of various concentrations (ranging from 0.0031 wt% to 3.13 wt%) were prepared and their structural properties were compared. After complete water removal through freeze-drying, lightweight sponge-like aerogel was produced which had not signicantly collapsed. Fig. 1
3300 | Soft Matter, 2010, 6, 32983305

Fig. 2 Aerogel density (g cm3) as a function of NFC dispersion concentrations during freeze-drying (g l1).

This journal is The Royal Society of Chemistry 2010

View Online

Downloaded by Nanyang Technological University on 31 May 2012 Published on 14 May 2010 on http://pubs.rsc.org | doi:10.1039/C001939A

0.011 g cm3, the number of pores and the pore size decreased signicantly accompanied with the formation of more sheet-like structures. At a density of 0.030 g cm3, no macroscopic pores were observed (Fig. 3f). Instead, the aerogel had a much more closed surface texture with thin sheets exhibiting random formations of wave-like roughness. The pore and sheet structures of the 0.0070 g cm3 aerogel were further studied by FE-SEM at higher magnications, and some micrographs are presented in Fig. 4. A typical sheet thickness of about 48 mm formed from aggregated nanobrils is revealed in Fig. 4a. Nanobril-aggregates or nanobril bundles with diameters in the order of 500 nm, are protruding from the surface of the sheets (Fig. 4b), some of them forming very open networks (Fig. 4c). These surface features were especially apparent in the aerogels with densities <0.01 g cm3. Fig. 4d shows an interconnected brillar structure of long and entangled nanobril aggregates with diameters of 100 nm with occasional thicker bundles. The image reveals that the bril network has a randomin-plane orientation. Surface wettability It is well established that in order to obtain a thin lm coverage of trichlorosilanes on any substrates, reactive OH-groups on the surface are generally required.47,54 These OH-groups can react with functionalized silanes (usually trichlorosilane) in the gaseous or solvent phase to yield preferably monolayers with the desired functionality depending on the chemical composition of the selected silane. In the work, this type of coating was applied to the structured and at cellulose surfaces (aerogels and solvent-cast lm, respectively) by simply reacting the hydroxylic groups of the cellulose with trichlorosilane. The unique structure of the coating was found to signicantly decrease the wetting properties evaluated by contact angle measurements using two probe liquids: castor oil and hexadecane with surface tensions of 35.8 and 27.5 mN m1, respectively. However, pure non-modied NFC aerogels and lms were completely wettable (q z 0 ) by castor oil. When 1H,1H,2H,2H-peruorooctyltrichlorosilane (PFOTS) was applied in the refunctionalization step, the structured coatings displayed clear oleophobic properties (q [ 90 ). Fig. 5 shows the advancing contact angle for castor oil as a function of the density of the PFOTS-coated aerogels.

Fig. 3 Low-magnication (100) FE-SEM micrographs of aerogels fabricated by the freeze-drying of aqueous NFC dispersions. The densities of the aerogels are (a) 0.00027 (b) 0.0018 (c) 0.0053 (d) 0.0070 (e) 0.011 and (f) 0.030 g cm3. All images are top-view images and the scale bars are 500 mm.

density of the aerogels increased, the amount of pores and protruding sheets decreased. This was accompanied by a gradual closure of the surface texture. At densities <0.0070 g cm3, the aerogel typically consisted of an open network of extended thin sheets forming macroscopic open channels and large, several micrometre wide pores that connect the different cells and thin sheets of the aerogel. When the density was increased to

Fig. 4 FE-SEM micrographs showing (a) a typical cross-section of a sheet, (b and c) bril-aggregates protruding from the cell walls forming network structures and (d) the brillar aerogel skeleton. Scale bars are (a) 40 mm (b) 10 mm (c) 50 mm and (d) 10 mm.

Fig. 5 Advancing contact angle for castor oil (glv 35.8 mN m1) as a function of the density of the PFOTS-coated aerogels.

This journal is The Royal Society of Chemistry 2010

Soft Matter, 2010, 6, 32983305 | 3301

View Online

Fig. 6 Droplets of (a) castor oil and (b) hexadecane on top of a PFOTScoated aerogel with a density r 0.0070 g cm3.

The most oleophobic aerogel (r 0.0070 g cm3) exhibited contact angles of roughly 166 and 144 for castor oil and hexadecane respectively (Fig. 6), which, to the knowledge of the authors, are among the highest contact angles reported in the literature for any substrate against liquids or liquids with similar surface tensions.45,47,50,55 At densities >0.020 g cm3, the advancing contact angle gradually decreased down to ca. 100 . At densities below 0.0003 g cm3, the advancing contact angle decreased down to 0 . Between the aerogels with densities of 0.00035 and 0.00027 g cm3 there is a rapid transition from a perfectly non-wetting (q [ 90 ) to a completely wetting surface (q z 0 ). A contact angle denoted 0 indicates that the surface was not sufciently robust to support a 10 ml droplet of castor oil, and that the liquid droplet was rapidly (<10 s) imbibed into the surface. Transitions from perfectly non-wetting (q [ 90 ) to completely wetting surfaces (q z 0 ) between aerogels with densities of 0.00035 and 0.00027 g cm3 were also observed for hexadecane and water. As expected, there was a large difference between the contact angle values on the structured aerogels and on the smooth solvent-casted NFC lm. For comparison, the contact angles for castor oil and hexadecane on PFOTS-modied smooth NFC lms were 96 and 71 , respectively, demonstrating the necessity of having a highly textured surface to achieve the high contact angles against low surface tension liquids. The XPS analysis revealed the surface chemical composition of the NFC aerogels and lm before and after PFOTS-coating. The atomic concentrations of the surfaces of pure NFC aerogel and lm and their modied analogues are given in Table 1. Carbon, oxygen and very small amounts of Na as counterions of the carboxymethylated cellulose brils were detected on the surface of the cellulose samples, whereas the modied samples clearly showed also the presence of uorine and silicon, which supported that PFOTS had reacted with the cellulose. The total oxygen concentration decreased from 38.2% to 6.6%, and the uorine concentration was found to be 51%. The atomic surface concentrations of C, O, Si and F were similar for all the
Table 1 Atomic surface concentration on PFOTS-coated and noncoated NFC aerogels and lms Surface concentration (at%) C Pure NFC aerogels/lm PFOTS-coated NFC aerogels/lm 61.3 37.6 O 38.2 6.6 Na 0.6 4.8 51.0 Si F

Downloaded by Nanyang Technological University on 31 May 2012 Published on 14 May 2010 on http://pubs.rsc.org | doi:10.1039/C001939A

structured PFOTS-coated aerogels and lms. As in the case of the uorinated NFC aerogels, the uorine signal from the PFOTS-coated NFC lm was determined to be 51%, indicating reproducible results from the PFOTS-coating. The pure NFC aerogel and lm showed a lower oxygen to carbon ratio (O/C-ratio), 0.62, compared with the theoretical value of pure cellulose, 0.83. Hydrocarbons tend to be the most common surface contaminant, and it is quite common to see CH peaks in cellulose samples.56 Although the aerogels and lms were handled carefully, it is possible that some surface contamination occurred. The high-surface area of the cellulose nanobrils makes it even more difcult to avoid contamination during e.g. lm formation and storage. The uncertainty of the XPS determinations was within 1%.

Discussion
Preparation of NFC aerogels by freeze-drying In the present work, highly porous NFC aerogels were prepared by direct water removal by freeze-drying. The primary aim has been to clarify the factors controlling the surface properties of these aerogels and to further modify and tune the oil-wetting/ resistance properties of these materials. The surface texture rather then the bulk properties of the aerogels was the aspect of major interest in this study. Nevertheless, as indicated by the FE-SEM and nitrogen adsorption/desorption-measurements, the porosity, density and surface morphology can be tuned by switchable adjustments of the concentration of the NFC dispersions before freeze-drying. A higher NFC concentration leads to a lower specic surface area of the aerogel. The same phenomenon was reported for freeze-dried cellulose/calcium thiocyanate solutions of different cellulose concentrations57 and for low-charged NFC aerogels.20 The inuence of freeze-drying conditions on the morphology and porosity of aerogels has been studied before, e.g. for NFC-based foams,20,22 regenerated cellulose II57 and for starch-based foams reinforced with NFC.36 The observed specic surface areas of the present aerogels were lower than those reported for freeze-dried brillar cellulose aerogels.20,21 This is probably due to a slower cooling of the NFC dispersions by the liquid nitrogen than the cooling by liquid et al.20 Rapid cooling is propane used in the work of P a ako known to be accompanied by the formation of amorphous ice which may lead to a more homogeneous brillar structure of the aerogel with smaller pores and less pronounced sheet-like structure of the aerogel.58 In contrast, cooling by liquid nitrogen enhances the formation of non-amorphous ice (crystals) which contributes to the sheet-formation. The aerogel structure is therefore directly related to the size and distribution of the ice crystals in the frozen system. In addition, the thickness of the aerogels plays an important role for the cooling rate. Thicker samples cool more slowly. Samples with a thickness of 10 mm were used in this study, whereas samples with thicknesses et al.20 between 1 and 5 mm were prepared by P a ako The fact that the aerogels have a structured porosity and morphology can be an advantage in various applications, for example, as in the present work, for modifying the surface wetting properties of the aerogel. The native cellulose nanobers are composed of cellulose I crystal domains as well as less
This journal is The Royal Society of Chemistry 2010

3302 | Soft Matter, 2010, 6, 32983305

View Online

ordered fractions,33 and this promotes high length, exibility, entanglements, and robust networks of the so prepared aerogels, even from very low NFC dispersion concentrations.20 As a result, the present nanobers make it possible to produce sponge-like aerogels with a lower density than previous cellulose aerogels.57 This means that lower gel concentrations are needed for the formation of robust aerogels that oppose collapse.20 In order to decrease the density and to tune the porosity by creating large pores, a surfactant has previously been used.21 In this study, no additional surfactant has been used, but, as previously discussed, a low density is due to brillar entanglement, and the porosity is tuned by adjusting the concentration of the NFC dispersions. Inuence of the aerogel surface chemistry and surface texture on the wettability by oils The apparent contact angle q* for a composite interface beneath a strongly non-wetting droplet is typically computed using the CassieBaxter relation:59 cos q* f1cos q f2 (1)

cos q* rffcos q + f 1

(6)

where f1 is the surface area of the liquid in contact with the solid divided by the projected area and f2 is the surface area of the liquid in contact with air trapped in the pores of the rough surface divided by the projected area, i.e.: f1 and f2 area in contact with air projected area (2) area in contact with liquid projected area

The Lotus leaf is an example of a natural surface that is forming a composite interface where water (glv 72.1 mN m1) droplets form beads on the surface. This natural surface comprises randomly distributed almost hemispherically topped papillae with sizes 510 mm.60 The formation of the Cassie Baxter state enhances water super-repellency by promoting a high apparent contact angle (q*) when f1  1.59 On the other hand, if the liquid fully penetrates into the surface texture, the apparent contact angle q* is determined by the Wenzel relation:61 cos q* rcos q (3)

where r is the surface roughness, dened as the ratio between the actual surface area and the projected area. Since r is necessarily greater than unity, roughness amplies both the wetting and non-wetting behavior of materials in the Wenzel regime, i.e., cos q* [ 0 if cos q > 0 and cos q*  0 if cos q < 0 . A consequence of this dependence on the roughness of the texture is that, once initiated, the imbibition of a liquid drop into a roughened texture can rapidly lead to super-wetting, because the apparent contact angle q* / 0 when r [ 1 and q < 90 . Eqn (1) has recently been rewritten as follows:62 f1 rff f2 1 f
This journal is The Royal Society of Chemistry 2010

(4) (5)

where f is the fraction of the projected area of the solid surface in contact with the liquid and rf is the roughness of the portion of the solid in contact with liquid. When f 1, rf r in the Wenzel model. It is important to note that rf in eqn (6) is not the roughness ratio of the total surface, but only of that in contact with the liquid. In this form of the CassieBaxter equation, the contributions of surface roughness and of entrapped air are clearer than in the other forms of the equation.62 The conditions for highly non-wettability (q* [ 90 ) can be realized only in the case of a composite interface where the solidliquid contact area is low. However, for low surface tension liquids with q < 90 , the fully wetted or Wenzel state represents the thermodynamic equilibrium state, whereas the composite interface or the Cassie Baxter state is metastable,45,50,6365 representing a local minimum in the overall Gibbs free energy. Thus, for low surface tension liquids, the transition from a composite interface to a fully wetted interface is irreversible, and typically this transition leads to a loss of non-wettability. Therefore the ability to preserve this metastable composite interface is crucial for engineering nonwettable surfaces.66 In our previous work, we demonstrated how the incorporation of a re-entrant surface texture (i.e., a multiscale surface topography) in conjunction with surface chemistry can be used to fabricate highly oleophobic surfaces, i.e., surfaces that can support a robust composite (solidliquidair) interface and display contact angles greater than 150 with various lowsurface-tension liquids such as castor oil and hexadecane.67 In the present work, the development of NFC aerogels with different surface textures has made it possible to ne-tune the surface wettability, including the capacity to switch the surface wetting properties between super-repellent and super-wetting against i.e. castor oil. The use of 1H,1H,2H,2H-peruorodecyltrichlorosilane (PFOTS) to generate oleophobic cellulose surfaces has been discussed earlier.67 The high concentration of peruorinated carbon atoms in the alkyl chains leads to a very low solid-surface energy for these molecules in the form of a thin layer structure (gsv z 13.5 mN m1).67 As a comparison, the surface energy of Teon is gsv 18 mN m1.68 To provide a conformal and exible coating of PFOTS molecules on the NFC aerogels and lms possessing re-entrant and at surface textures, respectively, a simple chemical vapor deposition (CVD) procedure was used. After the CVD, the equilibrium contact angle for castor oil on a smooth NFC lm increased to q z 96 compared to q z 0 on an uncoated lm. As previously discussed, a PFOTS-modied aerogel is able to support a composite interface even with hexadecane (glv 27.5 mN m1), as shown in Fig. 6b. FE-SEM micrographs showing the surface textures of aerogels with various densities are shown in Fig. 3. A comparison with the surface morphology of the PFOTS-coated aerogels (not shown) shows that all the surface details, even features in the submicrometre range, are preserved after modication. As previously discussed, for a given surface texture, the nonwetting properties can be most readily enhanced by markedly lowering the surface energy of the solid, leading to increased values of the equilibrium contact angle q (based on the Youngs equation69). According to XPS measurements, the uorine
Soft Matter, 2010, 6, 32983305 | 3303

Downloaded by Nanyang Technological University on 31 May 2012 Published on 14 May 2010 on http://pubs.rsc.org | doi:10.1039/C001939A

View Online

concentration of the PFOTS-coated aerogels was 51 1%. The variation in contact angle for castor oil on the aerogels is therefore attributed to the variation in surface texture and should be taken as the primary cause of the large difference in oleophobicity. The tunable oleophobic properties of the PFOTScoated aerogels result from the rough and porous surface textures and can be interpreted in terms of the Wenzel and CassieBaxter models. It is clear in Fig. 5 that, for very lowdensity aerogels (<0.0003 g cm3), the liquid is in contact with the entire solid surface and completely penetrates the surface texture generated by the multiple scale of roughness and pores, i.e. Wenzel mode of wetting. The number of pores and the pore size are most probably too large to enable these surfaces to support a composite interface (Fig. 3a). More robust surfaces are obtained on higher density aerogels with a maximum in oleophobicity (q z 166 ) at a density of 0.0070 g cm3 (Fig. 3d). These properties are probably related to a ne balance between the number of pores, the pore size and the extending sheet-like structures protruding from the surface. Multiple scales of roughness generated by protruding threads of bril aggregates probably also enhance the oleophobic properties. It is assumed that there is no penetration of oil into the gaps and that the liquid rests on the rough features of the protruding solid material. The air bridging these features then acts as further support for the oil droplet. The oil droplet can thus be considered to rest on only a part of the solid surface exposing a large fraction of its surface towards air and this situation can hence be modeled by the CassieBaxter equation. However, as the number of pores decreases for higher-density aerogels (>0.02 g cm3), a much more closed and non-porous surface structure is obtained (Fig. 3e and f). This simply results in a decreased surface area which greatly resembles that of a smooth NFC lm. A consequence of this dependence on roughness is a transition from the highly non-wetting state to the Wenzel state with q* z 96 . The development of highly oil-repellent surfaces requires the design of substrates that promote the formation of a composite interface with almost any liquid. The two important characteristics for arriving at a CassieBaxter state of wetting on a textured surface with a given liquid are: (i) the magnitude of the apparent contact angle q* on the composite interface and (ii) the robustness of the composite interface against external perturbation.66 Tuteja et al.45 developed a dimensionless design parameter A* to predict the robustness of a composite interface. This robustness factor represents the ratio of the breakthrough pressure, Pbreakthrough, required to cause sufcient sagging and disruption of the liquidvapor interface, to a characteristic reference pressure Pref, given as Pref 2glv/lcap where lcap p glv =rg (here r is the uid density and g is the acceleration due to gravity). If the sagging becomes severe enough for the interface to touch the underlying level of the solid texture, then the composite interface collapses, and the liquid droplet rapidly switches to a fully wetted state. The threshold pressure difference that triggers the transition, i.e. the breakthrough pressure, Pbreakthrough, can be computed as Pbreakthrough z A* Pref. Thus, large values of the robustness factor (A* [ 1) are associated with the formation of a robust composite interface with a very high breakthrough pressure.70 As previously discussed, the aerogels exhibit a random 3D structural feature consisting of brils sheets and protruding bril
3304 | Soft Matter, 2010, 6, 32983305

Downloaded by Nanyang Technological University on 31 May 2012 Published on 14 May 2010 on http://pubs.rsc.org | doi:10.1039/C001939A

bundles, and it is difcult to systematically relate or model the surface texture of the aerogels to the corresponding wetting properties. Nevertheless, attempts have been made to systemically relate certain surface geometries to a robustness factor and resulting wetting properties. Choi et al.66 designed geometrical parameters for a texture dominated by periodical cylindrical features; duck feathers, while Tuteja and co-workers45 analyzed structural parameters governing the properties of super-oleophobic electrospun bers. They showed that for a given liquid (given lcap), the robustness factor (A*) can be varied systematically, either by tuning the geometrical parameters describing the surface (such as ber radius and inter-ber gap) or by changing the equilibrium contact angle (q) through modication of the surface chemical composition. It should, however, be stressed that the authors used surfaces with well-dened periodically arranged structures and ber dimensions and that a similar treatment of the interfaces in the present work demands a new surface characterization procedure that is left for future studies on these materials.

Conclusions
Native nanobrillar cellulose aerogels have been prepared by vacuum freeze-drying of aqueous dispersions of carboxymethylated cellulose I nanobers. The morphology and porosity of the aerogels, as indicated by FE-SEM microscopy, can be tuned simply by adjusting the concentration of the NFC dispersions. A simple chemical vapor deposition process was developed to achieve a conformal coating of low-surface-energy PFOTS molecules on the aerogel surfaces. An XPS study of the chemical composition of the PFOTS-modied aerogels conrmed the reproducibility of the PFOTS-coating and the high atomic uorine concentration (ca. 51%) in the surface. The synergistic effect of roughness, re-entrant topography of the aerogels, and the low surface energy of the PFOTS molecules enables the CVD-coated surfaces to support a composite interface with low-surfacetension liquids such as castor oil and hexadecane. Advancing contact angle measurements on the PFOTS-coated aerogels were made, using castor oil as probe liquid. The results demonstrated the very high oleophobic nature (q* [ 90 ) of the aerogels compared to the corresponding contact angle (q z 90 ) for a smooth NFC lm. Aerogels with suitable surface textures were developed allowing the systematical adjustment of their surfacewettability characteristics. By combining this understanding with a CVD process that provides a conformal uorinated coating, we can switch the wettability behavior of the cellulose surfaces between super-wetting and super-repellent, using different scales of roughness and porosity created by the freeze-drying technique and change of concentration of the NFC dispersion.

Acknowledgements
The authors thank BIM Kemi Sweden AB and the Knowledge Foundation through its graduate school YPK for nancial support. Profs. Lars Odberg and Lars Berglund are greatly acknowledged for valuable discussions. M. Sc. Mikael Ankerfors is gratefully acknowledged for supplying the NFC dispersions. Dr Andrei Shchukarev at Ume a University is acknowledged for performing the XPS experiments.
This journal is The Royal Society of Chemistry 2010

View Online

References
1 C. Tan, B. M. Fung, J. K. Newman and C. Vu, Adv. Mater., 2001, 13, 644646. 2 S. S. Kistler, J. Phys. Chem., 1932, 36, 5264. 3 S. S. Kistler, Nature, 1931, 127, 741. 4 N. Husing and U. Schubert, Angew. Chem., Int. Ed., 1998, 37, 2245. 5 H. D. Gesser and P. C. Goswami, Chem. Rev., 1989, 89, 765788. 6 S. Dai, Y. H. Ju, H. J. Gao, J. S. Lin, S. J. Pennycook and C. E. Barnes, Chem. Commun., 2000, 243244. 7 H. Tamon, H. Ishizaka, T. Yamamoto and T. Suzuki, Carbon, 2000, 38, 10991105. 8 H. Tamon, H. Ishizaka, T. Yamamoto and T. Suzuki, Carbon, 1999, 37, 20492055. 9 S. Kuga, D.-Y. Kim, Y. Nishiyama and R. M. Brown, Mol. Cryst. Liq. Cryst. Sci. Technol., Sect. A, 2002, 387, 1319. 10 Z. Zhu, L. Y. Tsung and M. Tomkiewicz, J. Phys. Chem., 1995, 99, 1594515949. 11 D. J. Suh, T.-J. Park, J.-H. Kim and K.-L. Kim, Chem. Mater., 1997, 9, 19031905. 12 A. Pierre, R. Begag and G. Pajonk, J. Mater. Sci., 1999, 34, 4937 4944. 13 F. Meng, M. R. Schlup and L. T. Fan, Chem. Mater., 1997, 9, 2459 2463. 14 M. R. Ayers, X. Y. Song and A. J. Hunt, J. Mater. Sci., 1996, 31, 62516257. 15 R. W. Pekala and D. W. Schaefer, Macromolecules, 1993, 26, 5487 5493. 16 R. W. Pekala, J. Mater. Sci., 1989, 24, 32213227. 17 X. Lu, M. C. Arduini-Schuster, J. Kuhn, O. Nilsson, J. Fricke and R. W. Pekala, Science, 1992, 255, 971972. 18 J. Gross and J. Fricke, J. Non-Cryst. Solids, 1992, 145, 217222. 19 M. H. Nguyen and L. H. Dao, J. Non-Cryst. Solids, 1998, 225, 5157. , J. Vapaavuori, R. Silvennoinen, H. Kosonen, 20 M. P a akko m, L. A. Berglund and O. Ikkala, Soft M. Ankerfors, T. Lindstro Matter, 2008, 4, 24922499. 21 R. Gavillon and T. Budtova, Biomacromolecules, 2008, 9, 269277. 22 H. Sehaqui, M. Salajkov a, Q. Zhou and L. A. Berglund, Soft Matter, 2010, 6, 18241832. 23 D. Klemm, B. Heublein, H.-P. Fink and A. Bohn, Angew. Chem., Int. Ed., 2005, 44, 33583393. 24 Y. Nishiyama, P. Langan and H. Chanzy, J. Am. Chem. Soc., 2002, 124, 90749082. 25 A. F. Turbak, F. W. Snyder and K. R. Sandberg, J. Appl. Polym. Sci., 1983, 37, 815827. 26 F. W. Herrick, R. L. Casebier, J. K. Hamilton and K. R. Sandberg, J. Appl. Polym. Sci., 1983, 37, 797813. 27 T. Saito, S. Kimura, Y. Nishiyama and A. Isogai, Biomacromolecules, 2007, 8, 24852491. 28 H. Yano, J. Sugiyama, A. N. Nakagaito, M. Nogi, T. Matsuura, M. Hikita and K. Handa, Adv. Mater., 2005, 17, 153155. 29 A. N. Nakagaito and H. Yano, Appl. Phys. A: Mater. Sci. Process., 2004, 78, 547552. 30 G. Siqueira, J. Bras and A. Dufresne, Biomacromolecules, 2009, 10, 425432. , M. Ankerfors, H. Kosonen, A. Nyk 31 M. P a akko anen, S. Ahola, M. Osterberg, J. Ruokolainen, J. Laine, P. T. Larsson, O. Ikkala m, Biomacromolecules, 2007, 8, 19341941. and T. Lindstro m, M. Ankerfors 32 L. W agberg, G. Decher, M. Norgren, T. Lindstro s, Langmuir, 2008, 24, 784795. and K. Axna 33 C. Aulin, S. Ahola, P. Josefsson, T. Nishino, Y. Hirose, M. Osterberg and L. W agberg, Langmuir, 2009, 25, 76757685. 34 A. Dufresne, Can. J. Chem., 2008, 86, 484494.

35 A. N. Nakagaito and H. Yano, Appl. Phys. A: Mater. Sci. Process., 2004, 80, 155159. 36 A. J. Svagan, M. A. S. A. Samir and L. A. Berglund, Adv. Mater., 2008, 20, 12631269. 37 B. Bhushan, Y. C. Jung and K. Koch, Langmuir, 2009, 25, 32403248. 38 L. Gao and T. J. McCarthy, J. Am. Chem. Soc., 2006, 128, 90529053. 39 H. Y. Erbil, A. L. Demirel, Y. Avci and O. Mert, Science, 2003, 299, 13771380. 40 L. Zhang and D. E. Resasco, Langmuir, 2009, 25, 47924798. 41 L. Feng, S. Li, Y. Li, H. Li, L. Zhang, J. Zhai, Y. Song, B. Liu, L. Jiang and D. Zhu, Adv. Mater., 2002, 14, 18571860. m, A. Carlmark and L. W 42 S. Utsel, E. Malmstro agberg, Soft Matter, 2010, 6, 342352. nnberg, L. Fogelstro m, M. A. S. A. Samir, L. Berglund, 43 H. Lo m and A. Hult, Eur. Polym. J., 2008, 44, 29912997. E. Malmstro 44 A. Marmur, Langmuir, 2008, 24, 75737579. 45 A. Tuteja, W. Choi, M. Ma, J. M. Mabry, S. A. Mazzella, G. C. Rutledge, G. H. McKinley and R. E. Cohen, Science, 2007, 318, 16181622. 46 A. Nakajima, M. Hoshino, J.-H. Song, Y. Kameshima and K. Okada, Chem. Lett., 2005, 908909. 47 J. Zimmermann, M. Rabe, G. R. J. Artus and S. Seeger, Soft Matter, 2008, 4, 450452. 48 H. Yabu, M. Takebayashi, M. Tanaka and M. Shimomura, Langmuir, 2005, 21, 32353237. 49 A. Ahuja, J. A. Taylor, V. Lifton, A. A. Sidorenko, T. R. Salamon, E. J. Lobaton, P. Kolodner and T. N. Krupenkin, Langmuir, 2008, 24, 914. 50 L. Cao, T. P. Price, M. Weiss and D. Gao, Langmuir, 2008, 24, 1640 1643. 51 S. Katz, R. P. Beatson and A. M. Scallan, Sven. Papperstidn., 1984, 87, 4853. m, J. Colloid 52 L. Winter, L. W agberg, L. Odberg and T. Lindstro Interface Sci., 1986, 111, 537543. 53 Y. Nishiyama, J. Sugiyama, H. Chanzy and P. Langan, J. Am. Chem. Soc., 2003, 125, 1430014306. 54 D. C. Duffy, J. C. McDonald, O. J. A. Schueller and G. M. Whitesides, Anal. Chem., 1998, 70, 49744984. 55 T. Onda, S. Shibuichi, N. Satoh and K. Tsujii, Langmuir, 1996, 12, 21252127. 56 C. D. Edgar and D. G. Gray, Cellulose, 2003, 10, 299306. 57 H. Jin, Y. Nishiyama, M. Wada and S. Kuga, Colloids Surf., A, 2004, 240, 6367. 58 T. A. Jennings, Lyophilization: Introduction and Basic Principles, Interpharm Press, USA, 1999. 59 A. B. D. Cassie and S. Baxter, Trans. Faraday Soc., 1944, 40, 546551. 60 V. Zorba, E. Stratakis, M. Barberoglou, E. Spanakis, P. Tzanetakis, S. H. Anastasiadis and C. Fotakis, Adv. Mater., 2008, 20, 40494054. 61 R. N. Wenzel, Ind. Eng. Chem., 1936, 28, 988994. 62 A. Marmur, Langmuir, 2003, 19, 83438348. m, Langmuir, 2005, 21, 63 O. Werner, L. W agberg and T. Lindstro 1223512243. 64 L. Cao, H.-H. Hu and D. Gao, Langmuir, 2007, 23, 43104314. 65 M. Nosonovsky, Langmuir, 2007, 23, 31573161. 66 W. Choi, A. Tuteja, S. Chhatre, J. M. Mabry, R. E. Cohen and G. H. McKinley, Adv. Mater., 2009, 21, 21902195. m, ACS Appl. 67 C. Aulin, S. H. Yun, L. W agberg and T. Lindstro Mater. Interfaces, 2009, 1, 24432452. 68 E. Kissa, Fluorinated Surfactants and Repellents, Revised and Expanded, Marcel Dekker, New York, 2nd edn, 2001. 69 T. Young, Philos. Trans. R. Soc. London, 1805, 95, 65. 70 S. S. Chhatre, A. Tuteja, W. Choi, A. Revaux, D. Smith, J. M. Mabry, G. H. McKinley and R. E. Cohen, Langmuir, 2009, 25, 1362513632.

Downloaded by Nanyang Technological University on 31 May 2012 Published on 14 May 2010 on http://pubs.rsc.org | doi:10.1039/C001939A

This journal is The Royal Society of Chemistry 2010

Soft Matter, 2010, 6, 32983305 | 3305

You might also like