Skip to content
BY-NC-ND 3.0 license Open Access Published by De Gruyter September 19, 2017

Selective modification of inner surface of halloysite nanotubes: a review

  • Hailei Zhang

    Hailei Zhang received his BS and PhD degrees from Hebei University (P.R. China) in 2010 and 2017, respectively. Thereafter, he worked in the College of Chemistry and Environmental Science in Hebei University as a lectorate. Now, his interests are in developing chemical modification methods of nano-materials for biomedical applications, particularly on halloysite-based materials.

    ORCID logo EMAIL logo
From the journal Nanotechnology Reviews

Abstract

In this paper, we review the chemical strategies used for the modification of the inner surface of halloysite nanotubes (HNTs). The HNTs are nanotubular materials formed by rolling up the 1:1 aluminosilicate clays, where the composition is similar with kaolin. Owing to many virtues, including the high ratio of length to diameter, large cavity volume, desirable biocompatibility, and low cost, the HNTs have been applied to numerous promising domains. The modification of the outer surface is usually intended to decrease the HNT dispersal in aqueous media. Considering that the selective modification for the inner surface gives excellent prospects for hybrid HNT-based materials, herein, we explore the advances in the selective modification of the inner surface that expanded the applications of the HNTs.

1 Introduction

The breadth of research into halloysite nanotubes (HNTs) has expanded greatly over the past several years [1], [2]. The HNTs (Al2Si2O5(OH)4·nH2O) are natural hollow like aluminosilicate clay, which spread in many continents including Asia, North America, Europe, Oceania, and South America. The formation of the multilayer may be owing to the hydration effect of neighboring silica and alumina layers (shown in Figure 1A) [3], [4]. Generally, the multilayer tubule walls consist of 15~20 aluminosilicate layers with a layer spacing of ca. 10 Å or 7 Å for the hydrated or dehydrated HNTs, respectively [5], [6]. The external diameter of the HNTs usually ranges from 40 to 70 nm, and the length is within the scope from 500 to 1000 nm (shown in Figure 1B). Duarte et al. [7] presented a cross-section view of an ideal single-walled halloysite. The outer surface of the HNTs consists of siloxane groups (Si–O–Si), whereas the gibbsite octahedral array groups (Al–OH) overspread on the inner surface [7], [8].

Figure 1: The structure of HNTs: (A) the curly morphology of HNTs; (B) the micromorphology of HNTs measured by TEM, which was conducted by our group.
Figure 1:

The structure of HNTs: (A) the curly morphology of HNTs; (B) the micromorphology of HNTs measured by TEM, which was conducted by our group.

Owing to plenty of attractive virtues, including the high ratio of length to diameter, nanoscale hollow shape, lower density of hydroxyl groups on the outer surface, etc., the HNTs have been applied to numerous promising domains. The desirable biocompatibility and nontoxicity make the HNTs better candidates for bio-medical materials compared with other tubular nanomaterials, such as carbon nanotubes and TiO2 nanotubes [9], [10], [11], [12], [13], [14]. In Price’ review, they point out that the HNTs can be used to load metal and plastic anticorrosion, which can be utilized for biocide protection [8]. It has been reported that dozens of classical drugs can be loaded into the cavities of the HNTs to achieve sustained, pH-/thermo-responsive and targeted release systems [15], [16], [17], [18], [19], [20], [21], [22], [23], [24]. Lvov et al. pointed out that the relatively large inner diameter and the electropositive characteristic enable the HNTs to be DNA carriers. Afterward, some other biomacromolecules, especially for enzymes, were also reported to be loaded into the HNT lumen. The hydroxy groups make the HNT nano-catalysts in catalyzing esterification reactions, and the catalytic activity of metal nanoparticles can be significantly enhanced by the immobilization of the HNTs [25], [26], [27], [28], [29], [30], [31]. Moreover, the HNTs have been considered as attractive candidates for a number of diverse applications, ranging from biosensors [32], [33], hydrogen storage materials [34], adsorbing agents [35], [36], [37], to flame-resistant materials [38], [39].

In order to expand the application fields or their corresponding performances, the development of chemical modification methods toward the HNTs has gathered increasing concerns. Until now, the silane coupling agent remains the main strategy in the modification of the HNTs [40], [41]. However, the utilization of organosilane agents usually results in the modification of both inner and outer surfaces. With the strong demand to precisely design HNT-based materials, the selective covalent and supramolecular modification of the halloysite external surface has been achieved by Amorati, Riela, Lazzara, Massaro, Cavallaro, and others [42], [43], [44], [45]. Compared to the outer surface modifications, which are largely investigated, the modification of the alumina inner surface is much less present in the literature. Selective modification for the inner surface also gave excellent prospects for the hybrid HNT-based materials, which, however, remains a difficult task. Therefore, the main aim of this review is to summarize the selective modification methods upon the inner surface of the HNTs in order to provide some guiding principles on developing selectively modified HNTs. Moreover, some special changes about the ζ potential, morphologies, and some other physicochemical properties should be observed after halloysite lumen modification, which can serve as the confirmation of successful lumen modification and will also be briefly introduced in this review.

2 Alkyl phosphate

Alkyl phosphates are reported to be capable of self-assembling on some metallic oxide surfaces in aqueous solutions [46]. It has also been reported that organophosphorus molecules usually exhibit high affinity toward metal oxide surfaces. Michel et al. [47] used a photolithography method to create patterns consisting of TiO2 and silicon dioxide (SiO2). The XPS results show that alkane phosphates can only self-assemble on TiO2, whereas the they cannot self-assemble on SiO2. Lvov et al. [48] present that there is hope that aluminosilicate clays will exhibit a similar selectivity response to TiO2. The HNTs were treated with octadecylphosphonic acid in a weak acid environment (shown in Figure 2). It was evidenced by XPS, NMR, and FTIR that the Al−O−P bonds were formed on the inner alumina surface of the HNTs. No octadeylphosphonic acid bonding was detected on the outer surface of the HNTs, which demonstrated that the inner surface of the HNTs can be selectively modified by alkyl phosphates. In return, an inorganic micelle-like construction was created, which contains a hydrophobic core and a hydrophilic shell. Moreover, the bifunctionalization of halloysite can be achieved by treating the outermost silica surface through silanization after the selective reaction of alkyl phosphate on the inner surface.

Figure 2: Selective modification of the inner surface by octadecylphosphonic acid proposed by Lvov et al. [48].
Figure 2:

Selective modification of the inner surface by octadecylphosphonic acid proposed by Lvov et al. [48].

Sahnoune et al. [49] employed this method to graft polystyrene derivatives onto the inner surface of the HNTs (shown in Figure 3). In their study, a styrene/(methacryloyloxy)methyl phosphonic acid (P(S-co-MAPC1(OH)2)) copolymer was synthetized by typical radical polymerization. The prepared P(S-co-MAPC1(OH)2 was dissolved in toluene and, then, heated at solvent reflux to react with the inner surface of the HNTs. An amphiphilic property of the modified HNTs was achieved.

Figure 3: Preparation of amphiphilic HNTs by coupling P(S-co-MAPC1(OH)2 onto the inner surface of the HNTs [49].
Figure 3:

Preparation of amphiphilic HNTs by coupling P(S-co-MAPC1(OH)2 onto the inner surface of the HNTs [49].

On the other hand, it has been reported that aryl phosphates can also be used in the modification of HNTs. In Tang’s study, phenylphosphonic acid was employed to intercalate and unfold the sheet of the HNTs, leading to a significant improvement in basal spacing. A longer treatment time can transfer the nanotubes to nano-platelets [50]. However, no literature investigates whether aryl phosphates also show a selective modification capacity upon the inner surface of the HNTs.

3 Dopamine derivatives

Dopamine is a significant neurotransmitter in the brain, which can be released by nerve cells to send signals to other nerve cells [51], [52], [53]. As a catechol derivative, dopamine and its derivatives were reported to be utilized in the modification of metallic oxide surfaces for biomedical applications [54], [55], [56]. For example, Wang et al. [57] reported that a dopamine derivative bearing catechol units can be incorporated into magnetic Fe3O4 nanoparticles via a Schiff base bond.

Takahara et al. [58] pointed out that the binding of catechol groups to the outer surface is noncovalent, where the binding energy is calculated as 14 kcal/mol. They demonstrated that dopa, a dopamine derivative bearing catechol units, can also be coupled to the inner surface but not to the outer surface of the HNTs (shown in Figure 4). The surface modification of the HNTs was achieved by mixing the HNTs with a dopamine derivative in THF:H2O (4:1) and stirring at room temperature for 2 days. It should be noted that Takahara et al. [58] proposed a model experiment to verify the selective adsorption nature of dopamine. In this experiment, alumina and silicon nanoparticles were treated with dopamine derivative, respectively, under the same condition with the modification reaction of the HNTs. The dopamine derivative can only bond to the alumina nanoparticles but not to the silicon nanoparticles. These results are also appropriate to the inner and outer surfaces of the HNTs. Furthermore, the atom transfer radical polymerization (ATRP) was then introduced in the dopa-modified HNTs to give PMMA-filled HNTs.

Figure 4: Dopamine derivative for selective modification of the inner surface of HNTs [58].
Figure 4:

Dopamine derivative for selective modification of the inner surface of HNTs [58].

Lin et al. [59] used this method to prepare sulfonated HNTs. It significantly improved the compatibility and interfacial adhesion and gave the sulfonated HNT-based membrane materials void-free and uniform morphology.

4 Ionic liquid

An ionic liquid (IL) is an ionic salt whose fusing points is less than some arbitrary temperature, resulting in a liquid state [60], [61], [62]. ILs exhibit many potential applications, especially in the field of electric battery. They are powerful solvents with good electrical conductivity electrically, which are also known as liquid electrolytes.

Dedzo et al. [63] used 1-(2-hydroxyethyl)-3-methylimidazolium, an imidazole-based IL, to functionalize the tubes’ inner surfaces through the condensation reaction between the aluminol and alcohol to yield stable Al−O−C bonds, evidenced by the 13C solid-state NMR (shown in Figure 5). The introduction of ILs in the inner surface provides a greater electric density, ensuring a better accumulation of the nanoparticle precursors at the inner surface of the HNTs. In this way, palladium nanoparticle (PdNP) precursors selectively deposited inside the lumen of the HNTs. In contrast, the unmodified HNTs displayed nanoparticles both inside and outside the tubes. The PdNP-loaded HNTs showed excellent catalytic activity for the reduction of 4-nitrophenol and can be recycled up to three times without any significant reduction of the catalytic activities. The modification strategy developed in Dedzo’s study could be applied to a broad range of metal nanoparticles by selecting the chemical nature of the modifiers.

Figure 5: Selective modification of the inner surface by an ionic liquid (1-(2-hydroxyethyl)-3-methylimidazolium) [63].
Figure 5:

Selective modification of the inner surface by an ionic liquid (1-(2-hydroxyethyl)-3-methylimidazolium) [63].

5 Arylboronic acid

Arylboronic acid belongs to the class of organoboranes, which contains an aryl-substituted boric acid and at least a carbon-boron (B–C) bond. Early studies of alcohol-affinitive molecules revealed that arylboronic acid can rapidly react with diols via dehydration condensation; the reaction between arylboronic acid and pinacol is one of the most typical cases in organic chemistry. Yildirim et al. [64] used the reaction between vicinal diols and boronic acids to enlarge the interlamellar spacing of graphene oxide, where the gas adsorption capacity was dramatically increased.

Our group demonstrated that the arlyboronic acid can also be covalently linked to the inner surface but not to the outer surface of the HNTs [65]. In this way, 1-pyrenylboronic acid (PBA) was coupled onto the inner surface of the HNTs, which functionalized the modified HNTs with fluorescence properties (shown in Figure 6). Moreover, the established Al–O–B linkage gave a highly specific and sensitive H2O2 sensitivity to PBA-modified HNTs, which made them desirable fluorescence probes for the detection of hyperoxide.

Figure 6: Selective modification of the inner surface by 1-pyrenylboronic acid [65].
Figure 6:

Selective modification of the inner surface by 1-pyrenylboronic acid [65].

6 Electrostatic lumen coating

There is an attractive character for the HNTs that they possess a positively charged inner lumen (ζ=+24 mV) and a negatively charged silicondioxide outer surface (ζ=−35 mV) [1], [2]. Therefore, the selective modification can be achieved by utilizing the difference in electric potentials. Some electronegative molecules, e.g. negatively charged proteins and anionic surfactants, may be immobilized onto the inner surface of the HNTs [66].

In Lvov’s study, negatively charged proteins were selectively immobilized onto the lumen surface of the HNTs after typical vacuum cycles. The immobilized negatively charged proteins showed enhanced thermal stability and temporal biocatalytic abilities compared to free enzymes in solution [67]. In this way, Yan et al. [68] immobilized lipase, a negatively charged protein, onto the lumen surface of the HNTs, and then, the resulting enzyme-nanotube complex was mixed with chitosan to give an enzymatic membrane, which is capable of hydrolyzing lipids without the loss of enzymes.

Cavallaro et al. [69], [70], [71] demonstrated that an anionic surfactant can be an alternative to the use of the selective modification of the alumina inner surface on the HNTs. In their study, the inner surface of the HNTs can be selectively modified by sodium dodecanoate (NaL), decyltrimethylammonium bromide (DeTAB), and perfluorinated anionic surfactants through magnetic stirring. The use of anionic surfactant modification has been considered as a good strategy for oil and gas entrapment as well as a way to stabilize the HNT dispersion in water or controlling its partition between an aqueous and an oil phase [42], [69], [70]. The inner cavity modification with fluorinated anionic surfactant was evidenced to be a good strategy to generate flame-retardant additives [71]. Moreover, after halloysite lumen modification, its ζ potential has to increase making its stable aqueous colloids [35], [69], [72]. It happens because any internal tube modification “kills” internal positive charges, and a total ζ potential of the tubes is an algebraic sum of external minuses and inner pluses. This may be a simple confirmation of successful lumen modification.

7 Etching

It is well accepted that HNTs possess better biocompatibility than other nanotubular materials, especially for carbon nanotubes. HNTs may have a much broader application in the bio-medical fields when getting a larger cavity volume. Abdullayev et al. demonstrated a feasible process to enlarge the lumen diameter of the HNTs by etching the alumina sheets on the inner surface of the HNTs. In their study, sulfuric acid was employed to selectively dissolve the alumina sheets on the inner surface of the HNTs, whereas the exterior surface was preserved, which enhances two~ to -threefold the loading capacity as well as the surface area of the tubes [73].

PVA is a water-soluble polymer, which has the idealized formula [CH2CH(OH)]n [74], [75]. Because of its excellent film forming, emulsifying, adhesive, and nontoxic properties, PVA has been widely used in the field of bio-medicine, gas storage, etc. [76], [77], [78]. Our group utilized a unidirectional freezing method upon PVA/HNTs aqueous dispersions to achieve an ordered LC mesophase with parallel-banded structures in the formed PVA/HNTs aerogel [79]. The HNTs showed a desirable dispersibility in PVA matrix, resulting in serious PVA/HNTs composite materials in recent studies [80].

Cheng et al. [81] utilized PVA to develop a selective modification method upon the inner surface of the HNTs. In their work, the HNTs were initially calcined at 550°C and then added into the PVA aqueous solution. After drying, the PVA-treated HNTs were rinsed under sonication aiming to remove those PVA residues on the outer surface of the HNTs, evidenced by XPS patterns. Interestingly, the PVA-modified HNTs were etched to denote carbon nanotubes. The as-mentioned method provided a low-cost, versatile, and facile approach to obtain carbon nanotubes.

8 Click chemistry

The classic click reaction is a copper-catalyzed reaction of an azide(−N3) with an alkyne (−C≡H) to form a triazole ring [82], [83]. Owing to the desirable specificity and high efficiency, click reaction has been widely used to prepare organic-inorganic hybrid materials [84], [85]. In a previous study, we functionalized the azide groups onto the surface of the HNTs [86]. Polyfluorenes with terminal alkyne groups were grafted onto the azide-modified HNTs via a copper-catalyzed click reaction, where the grafting degree was significantly enhanced. However, the dispersibility of the HNTs was found to be worse mainly due to the fact that the treatment of organosilanes was for aluminosilicate surfaces usually resulting in nonselective modification.

Lazzara et al. [87] demonstrated a selective modification of the cavity in the HNTs by recourse to the click reaction (shown in Figure 7). To functionalize the internal surface of the HNTs, they facilitated two compounds with terminal azide or alkyne groups to be encapsulated into the HNT lumens, respectively. Two kinds of modified HNTs were catalyzed by Cu2+ to process the click reaction between −N3 and −C≡H terminations to generate the connected HNTs and then to give the HNT-based microfibers with a significant enhancement in mechanical properties. The FTIR and TGA results indicated that the modification occurred only in the cavity.

Figure 7: Selective modification of the inner surface of HNTs by click reaction [87].
Figure 7:

Selective modification of the inner surface of HNTs by click reaction [87].

9 Poly(ethylene oxide) (PEO)

Previously, it has been demonstrated that the selective modification of the inner surface of the HNTs by covalent bonds can be feasibly achieved, while the selective modification of the inner surface may also be achieved through hydrogen bonds.

In Liu’s study, the water natively living in the lumen of the HNTs was used as the medium to form the hydrogen bond between the HNTs and the PEO by extracting the PEO/HNT nanocomposite with a Soxhlet extractor. The hydrogen bonding interactions were evaluated, which confirmed that the PEO only interacts with the inner surface within the lumen of the HNTs [88].

10 Conclusion

We focus on the development of selective modification of the inner surface of the HNTs, including the employment of alkyl phosphate, dopamine derivative, IL, arylboronic acid, PVA, PEO, and click chemistry, which will be able to provide multiple choices to prepare the selectively modified HNT-based organic-inorganic hybrid materials. These advances may expand the application fields, as well as enhance the performance of the HNTs. Otherwise, the methods discussed in this paper may also be capable of providing some guiding lights on the selective modification of other kaolin-based inorganic materials, such as imogolite nanotubes. Furthermore, a further understanding of how structure and composition are precisely affected by the selective modifications, or the detailed applicability of these methods, cannot be overemphasized.

About the author

Hailei Zhang

Hailei Zhang received his BS and PhD degrees from Hebei University (P.R. China) in 2010 and 2017, respectively. Thereafter, he worked in the College of Chemistry and Environmental Science in Hebei University as a lectorate. Now, his interests are in developing chemical modification methods of nano-materials for biomedical applications, particularly on halloysite-based materials.

References

[1] Lvov Y, Wang W, Zhang L, Fakhrullin R. Halloysite clay nanotubes for loading and sustained release of functional compounds. Adv. Mater. 2016, 28, 1227–1250.10.1002/adma.201502341Search in Google Scholar PubMed

[2] Lvov Y, Aerov A, Fakhrullin R. Clay nanotube encapsulation for functional biocomposites. Adv. Colloid Interface 2014, 207, 189–198.10.1016/j.cis.2013.10.006Search in Google Scholar PubMed

[3] Liu M, Jia Z, Jia D, Zhou C. Recent advance in research on halloysite nanotubes-polymer nanocomposite. Prog. Polym. Sci. 2014, 39, 1498–1525.10.1016/j.progpolymsci.2014.04.004Search in Google Scholar

[4] Rawtani D, Agrawal YK. Multifarious applications of halloysite nano tubes: a review. Rev. Adv. Mater. Sci. 2012, 30, 282–295.Search in Google Scholar

[5] Fisher GB, Ryan, PC. The smectite-to-disordered kaolinite transition in a tropical soil chronosequence, Pacific Coast, Costa Rica. Clays Clay Miner. 2006, 54, 571–586.10.1346/CCMN.2006.0540504Search in Google Scholar

[6] Hillier S, Ryan PC. Identification of halloysite (7Å) by ethylene glycol solvation: the ‘MacEwan effect’. Clay Miner. 2002, 37, 487–496.10.1180/0009855023730047Search in Google Scholar

[7] Guimaraes L, Enyashin A, Seifert G, Duarte H. Structural, electronic, and mechanical properties of single-walled halloysite nanotube models. J. Phys. Chem. C 2010, 114, 11358–11363.10.1021/jp100902eSearch in Google Scholar

[8] Lvov Y, Shchukin D, Mohwald H, Price R. Halloysite clay nanotubes for controlled release of protective agents. ACS Nano 2008, 2, 814–820.10.1021/nn800259qSearch in Google Scholar PubMed

[9] Mitchell MJ, Castellanos CA, King MR. Surfactant functionalization induces robust, differential adhesion of tumor cells and blood cells to charged nanotube-coated biomaterials under flow. Biomaterials 2015, 56, 179–186.10.1016/j.biomaterials.2015.03.045Search in Google Scholar PubMed PubMed Central

[10] Vergaro V, Abdullayev E, Lvov Y, Zeitoun A, Cingolani R, Rinaldi R, Leporatti S. Cytocompatibility and uptake of halloysite clay nanotubes. Biomacromolecules 2010, 11, 820–826.10.1021/bm9014446Search in Google Scholar PubMed

[11] Chao C, Liu J, Wang J, Zhang Y, Zhang B, Zhang Y, Xiang X, Chen R. Surface Modification of halloysite nanotubes with dopamine for enzyme immobilization. ACS Appl. Mater. Interfaces 2013, 5, 10559–10564.10.1021/am4022973Search in Google Scholar PubMed

[12] Sandri G, Aguzzi C, Rossi S, Bonferoni MC, Bruni G, Boselli C, Cornaglia AI, Riva F, ViserasC, Caramella C. Halloysite and chitosan oligosaccharide nanocomposite for wound healing. Acta Biomater. 2017, 57, 216–224.10.1016/j.actbio.2017.05.032Search in Google Scholar PubMed

[13] Naumenko EA, Guryanov ID, Yendluri R, Lvov YM, Fakhrullin RF. Clay nanotube-biopolymer composite scaffolds for tissue engineering. Nanoscale 2016, 8, 7257–7271.10.1039/C6NR00641HSearch in Google Scholar PubMed

[14] Liu M, He R, Yang J, Zhao W, Zhou C. Stripe-like clay nanotubes patterns in glass capillary tubes for capture of tumor cells. ACS Appl. Mater. Interfaces 2016, 8, 7709–7719.10.1021/acsami.6b01342Search in Google Scholar PubMed

[15] Price R, Gaber B, Lvov Y. In-vitro release characteristics of tetracycline HCl, khellin and nicotinamide adenine dineculeotide from halloysite; a cylindrical mineral. J. Microencapsul. 2001, 18, 713–722.10.1080/02652040010019532Search in Google Scholar PubMed

[16] Veerabadran NG, Price RR, Lvov YM. Clay nanotubes for drug encapsulation and sustained release. Nano 2007, 2, 215–222.10.1142/S1793292007000441Search in Google Scholar

[17] Massaro M, Lazzara G, Milioto S, Noto R, Riela S. Covalently modified halloysite clay nanotubes: synthesis, properties, biological and medical applications. J. Mater. Chem. B 2017, 5, 2867–2882.10.1039/C7TB00316ASearch in Google Scholar

[18] Liu M, Chang Y, Yang J, You Y, He R, Chen T, Zhou C. Functionalized halloysite nanotube by chitosan grafting for drug delivery of curcumin to achieve enhanced anticancer efficacy. J. Mater. Chem. B 2016, 4, 2253–2263.10.1039/C5TB02725JSearch in Google Scholar

[19] Kadam AA, Jang J, Lee DS. Supermagnetically tuned halloysite nanotubes functionalized with aminosilane for covalent laccase immobilization. ACS Appl. Mater. Interfaces 2017, 9, 15492–15501.10.1021/acsami.7b02531Search in Google Scholar PubMed

[20] Li H, Zhu X, Zhou H, Zhong S. Functionalization of halloysite nanotubes by enlargement and hydrophobicity for sustained release of analgesic. Colloid. Surface. A 2015, 487, 154–161.10.1016/j.colsurfa.2015.09.062Search in Google Scholar

[21] Riela S, Massaro M, Colletti C, Bommarito A, Giordano C, Milioto S, Noto R, Poma P, Lazzara G. Development and characterization of co-loaded curcumin/triazole-halloysite systems and evaluation of their potential anticancer activity. Int. J. Pharmaceut. 2014, 475, 613–623.10.1016/j.ijpharm.2014.09.019Search in Google Scholar PubMed

[22] Li W, Liu D, Zhang H, Correia A, Mäkilä E, Salonen J, Hirvonen J, Santos HA. Microfluidic assembly of a nano-in-micro dual drug delivery platform composed of halloysite nanotubes and a pH-responsive polymer for colon cancer therapy. Acta Biomater. 2017, 48, 238–246.10.1016/j.actbio.2016.10.042Search in Google Scholar

[23] Lvov Y, Devilliers M, Fakhrullin R. The application of halloysite tubule nanoclay in drug delivery. Expert Opin. Drug Del. 2016, 13, 977–986.10.1517/17425247.2016.1169271Search in Google Scholar

[24] Massaro M, Amorati R, Cavallaro G, Guernelli S, Lazzara G, Milioto S, Noto R, Poma P, Riela S. Direct chemical grafted curcumin on halloysite nanotubes as dual-responsive prodrug for pharmacological applications. Colloids Surf. B Biointerfaces 2016, 140, 505–513.10.1016/j.colsurfb.2016.01.025Search in Google Scholar PubMed

[25] Rong T, Xiao J. The catalytic cracking activity of the kaolin-group minerals. Mater. Lett. 2002, 57, 297–301.10.1016/S0167-577X(02)00781-4Search in Google Scholar

[26] Zatta L, Gardolinski J, Wypych F. Raw halloysite as reusable heterogeneous catalyst for esterification of lauric acid. Appl. Clay Sci. 2011, 51, 165–169.10.1016/j.clay.2010.10.020Search in Google Scholar

[27] Sanchez-Ballester NM, Ramesh GV, Tanabe T, Koudelkova E, Liu J, Shrestha LK, Lvov Y, Hill JP, Ariga K, Abe H. Activated interiors of clay nanotubes for agglomeration-tolerant automotive exhaust remediation. J. Mater. Chem. A 2015, 3, 6614–6619.10.1039/C4TA06966HSearch in Google Scholar

[28] Abdullayev E, Sakakibara K, Okamoto K, Wei W, Ariga K, Lvov Y. Natural tubule clay template synthesis of silver nanorods for antibacterial composite coating. ACS Appl. Mater. Interfaces 2011, 3, 4040–4046.10.1021/am200896dSearch in Google Scholar PubMed

[29] Prieto G, Tueysuez H, Duyckaerts N, Knossalla J, Wang G, Schueth F. Hollow nano- and microstructures as catalysts. Chem. Rev. 2016, 116, 14056–14119.10.1021/acs.chemrev.6b00374Search in Google Scholar PubMed

[30] Zeng X, Wang Q, Wang H, Yang Y Catalytically active silver nanoparticles loaded in the lumen of halloysite nanotubes via electrostatic interactions. J. Mater. Sci. 2017, 52, 8391–8400.10.1007/s10853-017-1073-ySearch in Google Scholar

[31] Kenne DG, Ngnie G, Detellier C. PdNPs decoration of halloysite lumen via selective grafting of ionic liquid onto the aluminol surfaces and catalytic application. ACS Appl. Mater. Interfaces 2016, 8, 4862–4869.10.1021/acsami.5b10407Search in Google Scholar PubMed

[32] Wu Q, Sheng Q, Zheng J. Nonenzymatic sensing of glucose using a glassy carbon electrode modified with halloysite nanotubes heavily loaded with palladium nanoparticles. J. Electroanal. Chem. 2015, 762, 51–58.10.1016/j.jelechem.2015.12.030Search in Google Scholar

[33] Yang Z, Zheng X, Zheng J. Non-enzymatic sensor based on a glassy carbon electrode modified with Ag nanoparticles/polyaniline/halloysite nanotube nanocomposites for hydrogen peroxide sensing. RSC Adv. 2016, 6, 58329–58335.10.1039/C6RA06366GSearch in Google Scholar

[34] Ohashi F, Tomura S, Akaku K, Hayashi S, Wada S. Characterization of synthetic imogolite nanotubes as gas storage. J. Mater. Sci. 2004, 39, 1799–1801.10.1023/B:JMSC.0000016188.04444.36Search in Google Scholar

[35] Cavallaro G, Lazzara G, Milioto S, Parisi F, Sanzillo V. Modified halloysite nanotubes: nanoarchitectures for enhancing the capture of oils from vapor and liquid phases. ACS Appl. Mater. Interfaces 2014, 6, 606–612.10.1021/am404693rSearch in Google Scholar PubMed

[36] Duan J, Liu R, Chen T, Zhang B, Liu J. Halloysite nanotube-Fe3O4 composite for removal of methyl violet from aqueous solutions. Desalination 2012, 293, 46–52.10.1016/j.desal.2012.02.022Search in Google Scholar

[37] Kilislioglu A, Bilgin B. Adsorption of uranium on halloysite. Radiochim. Acta 2011, 90, 155–160.10.1524/ract.2002.90.3_2002.155Search in Google Scholar

[38] Schartel B, Hull T. Development of fire-retarded materials – interpretation of cone calorimeter data. Fire Mater. 2007, 31, 327–354.10.1002/fam.949Search in Google Scholar

[39] Chen H, Wang Y, Schiraldi D. Preparation and flammability of poly(vinyl alcohol) composite aerogels. ACS Appl. Mater. Interfaces 2014, 6, 6790–6796.10.1021/am500583xSearch in Google Scholar PubMed

[40] Yuan P, Southon P, Liu Z, Green M, Hook J, Antill S, Kepert C. Functionalization of halloysite clay nanotubes by grafting with gamma-aminopropyltriethoxysilane. J. Phys. Chem. C 2008, 112, 15742–15751.10.1021/jp805657tSearch in Google Scholar

[41] Li C, Liu J, Qu X, Guo B, Yang Z. Polymer-modified halloysite composite nanotubes. J. Appl. Polym. Sci. 2008, 110, 3638–3646.10.1002/app.28879Search in Google Scholar

[42] Cavallaro G, Lazzara G, Milioto S, Parisi F. Hydrophobically modified halloysite nanotubes as reverse micelles for water-in-oil emulsion. Langmuir 2015, 31, 7472–7478.10.1021/acs.langmuir.5b01181Search in Google Scholar PubMed

[43] Massaro M, Riela S, Guernelli S, Parisi F, Lazzara G, Baschieri A, Valgimigli L, Amorati R. A synergic nanoantioxidant based on covalently modified halloysite–trolox nanotubes with intra-lumen loaded quercetin. J. Mater. Chem. B 2016, 4, 2229–2241.10.1039/C6TB00126BSearch in Google Scholar

[44] Riela S, Massaro M, Colletti CG, Lazzara G, Milioto S, Noto R. Halloysite nanotubes as support for metal-based catalysts. J. Mater. Chem. A 2017, 5, 13276–13293.10.1039/C7TA02996ASearch in Google Scholar

[45] Riela S, Rizzo C, Arrigo R, D‘Anna F, Dintcheva NT, Lazzara G, Parisi F, Massaro M, Blasi FD, Spinelli G. Hybrid supramolecular gels of Fmoc-F/halloysite nanotubes: systems for sustained release of camptothecin. J. Mater. Chem. B 2017, 5, 3217–3229.10.1039/C7TB00297ASearch in Google Scholar

[46] Textor M, Ruiz L, Hofer R, Rossi A, Feldman K, Hähner G, Spencer ND. Structural chemistry of self-assembled monolayers of octadecylphosphoric acid on tantalum oxide surfaces. Langmuir 2000, 16, 3257–3271.10.1021/la990941tSearch in Google Scholar

[47] Michel R, Lussi JW, Csucs G, Reviakine I, Danuser G, Ketterer B, Hubbell JA, Textor M, Spencer ND. Selective molecular assembly patterning: a new approach to micro- and nanochemical patterning of surfaces for biological applications. Langmuir 2002, 18, 3281–3287.10.1021/la011715ySearch in Google Scholar

[48] Yah W, Takahara A, Lvov Y. Selective modification of halloysite lumen with octadecylphosphonic acid: new inorganic tubular micelle. J. Am. Chem. Soc. 2012, 134, 1853–1859.10.1021/ja210258ySearch in Google Scholar PubMed

[49] Sahnoune M, Taguet A, Otazaghine B, Kaci M. Inner surface modification of halloysite nanotubes and its influence on morphology and thermal properties of polystyrene/polyamide-11 blends. Polym. Int. 2017, 66, 300–312.10.1002/pi.5266Search in Google Scholar

[50] Tang Y, Deng S, Ye L, Yang C, Yuan Q, Zhang J, Zhao C. Effects of unfolded and intercalated halloysites on mechanical properties of halloysite – epoxy nanocomposites. Compos. Part A 2011, 42, 345–354.10.1016/j.compositesa.2010.12.003Search in Google Scholar

[51] Dawson TM, Dawson VL. Molecular pathways of neurodegeneration in Parkinson’s disease. Science 2003, 302, 819–822.10.1126/science.1087753Search in Google Scholar PubMed

[52] Nutt DJ, Lingford-Hughes A, Erritzoe D, Stokes PRA. The dopamine theory of addiction: 40 years of highs and lows. Nat. Rev. Neurosci. 2015, 16, 305–312.10.1038/nrn3939Search in Google Scholar PubMed

[53] Volkow ND, Koob GF, McLellan AT. Neurobiologic advances from the brain disease model of addiction. New Engl. J. Med. 2016, 374, 363–371.10.1056/NEJMra1511480Search in Google Scholar PubMed PubMed Central

[54] Kang HJ, Song XS, Wang Z, Zhang W, Zhang SF, Li JZ. High-performance and fully renewable soy protein isolate-based film from microcrystalline cellulose via bio-inspired poly(dopamine) surface modification. ACS Sustain. Chem. Eng. 2016, 4, 4354–4360.10.1021/acssuschemeng.6b00917Search in Google Scholar

[55] Wu F, Li JA, Zhang K, He ZK, Yang P, Zou D, Huang N. Multifunctional coating based on hyaluronic acid and dopamine conjugate for potential application on surface modification of cardiovascular implanted devices. ACS Appl. Mater. Interfaces 2016, 8, 109–121.10.1021/acsami.5b07427Search in Google Scholar PubMed

[56] Vaish A, Vanderah DJ, Richter LJ, Dimitriou M, Steffens KL, Walker ML. Dithiol-based modification of poly(dopamine): enabling protein resistance via short-chain ethylene oxide oligomers. Chem. Commun. 2015, 51, 6591–6594.10.1039/C5CC00299KSearch in Google Scholar

[57] Wang B, Xu C, Xie J, Yang Z, Sun S. pH controlled release of chromone from chromone-Fe3O4 nanoparticles. J. Am. Chem. Soc. 2008, 130, 14436–14437.10.1021/ja806519mSearch in Google Scholar PubMed PubMed Central

[58] Yah W, Xu H, Soejima H, Ma W, Lvov Y, Takahara A. Biomimetic dopamine derivative for selective polymer modification of halloysite nanotube lumen. J. Am. Chem. Soc. 2012, 134, 12134–12137.10.1021/ja303340fSearch in Google Scholar PubMed

[59] Liu X, He S, Song G, Jia H, Shi Z, Liu S, Zhang L, Lin J, Nazarenko S. Proton conductivity improvement of sulfonated poly(ether ether ketone) nanocomposite membranes with sulfonated halloysite nanotubes prepared via dopamine-initiated atom transfer radical polymerization. J. Membrane Sci. 2016, 504, 206–219.10.1016/j.memsci.2016.01.023Search in Google Scholar

[60] Jordan A, Gathergood N. Biodegradation of ionic liquids – a critical review. Chem. Soc. Rev. 2015, 44, 8200–8237.10.1039/C5CS00444FSearch in Google Scholar PubMed

[61] Goossens K, Lava K, Bielawski CW, Binnemans, K. Ionic liquid crystals: versatile materials. Chem. Rev. 2016, 116, 4643–4807.10.1021/cr400334bSearch in Google Scholar PubMed

[62] MacFarlane DR, Forsyth M, Howlett PC, Kar M, Passerini S, Pringle JM, Ohno H, Watanabe M, Yan F, Zheng WJ, Zhang SG, Zhang J. Ionic liquids and their solid-state analogues as materials for energy generation and storage. Nat. Rev. Mater. 2016, 1, 15005.10.1038/natrevmats.2015.5Search in Google Scholar

[63] Dedzo GK, Ngnie G, Detellier C. PdNP decoration of halloysite lumen via selective grafting of ionic liquid onto the aluminol surfaces and catalytic application. ACS Appl. Mater. Interfaces 2016, 8, 4862–4869.10.1021/acsami.5b10407Search in Google Scholar PubMed

[64] Burress J, Gadipelli S, Ford J, Simmons J, Zhou W, Yildirim T. Graphene oxide framework materials: theoretical predictions and experimental results. Angew. Chem. Int. Ed. 2010, 49, 8902–8904.10.1002/anie.201003328Search in Google Scholar PubMed

[65] Zhang H, Ren T, Ji Y, Han L, Wu Y, Song H, Bai L, Ba X. Selective modification of halloysite nanotubes with 1-pyrenylboronic acid: a novel fluorescence probe with highly selective and sensitive response to hyperoxide. ACS Appl. Mater. Interfaces 2015, 7, 23805–23811.10.1021/acsami.5b08600Search in Google Scholar PubMed

[66] Shchukin DG, Sukhorukov GB, Price RR, Lvov YM. Halloysite nanotubes as biomimetic nanoreactors. Small 2005, 1, 510–513.10.1002/smll.200400120Search in Google Scholar PubMed

[67] Tully J, Yendluri R, Lvov Y. Halloysite clay nanotubes for enzyme immobilization. Biomacromolecules 2016, 17, 615–621.10.1021/acs.biomac.5b01542Search in Google Scholar PubMed

[68] Sun J, Yendluri R, Liu K, Guo Y, Lvov Y, Yan X. Enzyme-immobilized clay nanotube-chitosan membranes with sustainable biocatalytic activities. Phys. Chem. Chem. Phys. 2016, 19, 562–567.10.1039/C6CP07450BSearch in Google Scholar PubMed

[69] Cavallaro G, Lazzara G, Milioto S. Exploiting the colloidal stability and solubilization ability of clay nanotubes/ionic surfactant hybrid nanomaterials. J. Phys. Chem. C 2012, 116, 21932–21938.10.1021/jp307961qSearch in Google Scholar

[70] Cavallaro G, Lazzara G, Milioto S, Palmisano G, Parisi, F. Halloysite nanotube with fluorinated lumen: non-foaming nanocontainer for storage and controlled release of oxygen in aqueous media. J. Colloid Interface Sci. 2014, 417, 66–71.10.1016/j.jcis.2013.11.026Search in Google Scholar PubMed

[71] Cavallaro G, Lazzara G, Milioto S, Parisi, F. Halloysite nanotubes with fluorinated cavity: an innovative consolidant for paper treatment. Clay Miner. 2016, 51, 445–455.10.1180/claymin.2016.051.3.01Search in Google Scholar

[72] Hu XJ, Liu YG, Zeng GM, Wang H, Hu X, Chen AW, Wang YQ, Guo YM, Li TT, Zhou L. Effect of aniline on cadmium adsorption by sulfanilic acid-grafted magnetic graphene oxide sheets. J. Colloid Interface Sci. 2014, 426, 213–220.10.1016/j.jcis.2014.04.016Search in Google Scholar PubMed

[73] Abdullayev E, Joshi A, Wei W, Zhao Y, Lvov Y. Enlargement of halloysite clay nanotube lumen by selective etching of aluminum oxide. ACS Nano 2012, 6, 7216–7226.10.1021/nn302328xSearch in Google Scholar PubMed

[74] Kumar A, Han SS. PVA-based hydrogels for tissue engineering: a review. Int. J. Polym. Mater. Po. 2016, 66, 159–182.10.1080/00914037.2016.1190930Search in Google Scholar

[75] Karimi A, Wan MAWD. Materials, preparation, and characterization of PVA/MMT nanocomposite hydrogels: a review. Polym. Composite. 2015, 38, 1086–1102.10.1002/pc.23671Search in Google Scholar

[76] Fan L, Yang H, Jing Y, Min P, Jin H. Preparation and characterization of chitosan/gelatin/PVA hydrogel for wound dressings. Carbohyd. Polym. 2016, 146, 427–434.10.1016/j.carbpol.2016.03.002Search in Google Scholar PubMed

[77] Zhang M, Howe RCT, Woodward RI, Kelleher EJR, Torrisi F, Hu G, Popov SV, Taylor JR, Hasan T. Solution processed MoS2-PVA composite for sub-bandgap mode-locking of a wideband tunable ultrafast Er: fiber laser. Nano Res. 2015, 8, 1522–1534.10.1007/s12274-014-0637-2Search in Google Scholar

[78] Majidnia Z, Idris A. Evaluation of cesium removal from radioactive waste water using maghemite PVA – alginate beads. Chem. Eng. J. 2015, 262, 372–382.10.1016/j.cej.2014.09.118Search in Google Scholar

[79] Song HZ, Zhao NN, Qin WC, Duan B, Ding XY, Wen X, Qiu P, Ba XW. High-performance ionic liquid-based nanocomposite polymer electrolytes with anisotropic ionic conductivity prepared by coupling liquid crystal self-templating with unidirectional freezing. J. Mater. Chem. A 2015, 3, 2128–2134.10.1039/C4TA05720ASearch in Google Scholar

[80] Gaaz TS, Sulong AB, Akhtar MN, Kadhum AA, Mohamad AB, Alamiery AA. Properties and applications of polyvinyl alcohol, halloysite nanotubes and their nanocomposites. Molecules 2015, 20, 22833–22847.10.3390/molecules201219884Search in Google Scholar PubMed PubMed Central

[81] Cheng Z, Liu Y, Liu Z. Novel template preparation of carbon nanotubes with natural HNTs employing selective PVA modification. Surf. Coat. Tech. 2016, 307, 633–638.10.1016/j.surfcoat.2016.09.081Search in Google Scholar

[82] Lowe AB. Thiol-ene ‘click’ reactions and recent applications in polymer and materials synthesis. Polym. Chem. 2010, 1, 17–36.10.1039/B9PY00216BSearch in Google Scholar

[83] Jewett JC, Bertozzi CR. Cu-free click cycloaddition reactions in chemical biology. Chem. Soc. Rev. 2010, 39, 1272–1279.10.1039/b901970gSearch in Google Scholar PubMed PubMed Central

[84] Feng C, Li Y, Yang D, Hu J, Zhang X, Huang X. Well-defined graft copolymers: from controlled synthesis to multipurpose applications. Chem. Soc. Rev. 2011, 42, 1282–1295.10.1039/B921358ASearch in Google Scholar

[85] Ge Z, Wang D, Zhou Y, Liu H, Liu S. Synthesis of organic/inorganic hybrid quatrefoil-shaped star-cyclic polymer containing a polyhedral oligomeric silsesquioxane core. Macromolecules 2009, 42, 2903–2910.10.1021/ma802585kSearch in Google Scholar

[86] Zhang H, Zhu X, Wu Y, Song H, Ba X. High-efficiency grafting of halloysite nanotubes by using p-conjugated polyfluorenes via “click” chemistry. J. Mater. Sci. 2015, 50, 4387–4395.10.1007/s10853-015-8993-1Search in Google Scholar

[87] Arcudi F, Cavallaro G, Lazzara G, Massaro M, Milioto S, Noto R, Riela S. Selective functionalization of halloysite cavity by click reaction: structured filler for enhancing mechanical properties of bionanocomposite films. J. Phys. Chem. C 2014, 118, 15095–15101.10.1021/jp504388eSearch in Google Scholar

[88] Yang S, Liu Z, Jiao Y, Liu Y, Ji C, Zhang Y. New insight into PEO modified inner surface of HNTs and its nano-confinement within nanotube. J. Mater. Sci. 2014, 49, 4270–4278.10.1007/s10853-014-8122-6Search in Google Scholar

Received: 2017-7-17
Accepted: 2017-8-29
Published Online: 2017-9-19
Published in Print: 2017-11-27

©2017 Walter de Gruyter GmbH, Berlin/Boston

This article is distributed under the terms of the Creative Commons Attribution Non-Commercial License, which permits unrestricted non-commercial use, distribution, and reproduction in any medium, provided the original work is properly cited.

Downloaded on 10.5.2024 from https://www.degruyter.com/document/doi/10.1515/ntrev-2017-0163/html
Scroll to top button