Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Thin Solid Films 528 (2013) 93–105

Contents lists available at SciVerse ScienceDirect

Thin Solid Films


journal homepage: www.elsevier.com/locate/tsf

Surface engineering and modification of biomaterials


Paul K. Chu ⁎
Department of Physics & Materials Science, City University of Hong Kong, Tat Chee Avenue, Kowloon, Hong Kong, China

a r t i c l e i n f o a b s t r a c t

Available online 7 November 2012 Surface engineering using plasma, grafting, and related techniques is an important area in biomaterials research
and biomedical engineering. The burgeoning technology enables the modification of selected surface character-
Keywords: istics while the favorable bulk materials properties can be retained. In this invited mini-review, recent work
Surface engineering: Plasma immersion ion related to surface modification of biomaterials by plasma-based and related techniques conducted in the Plasma
implantation and deposition Laboratory at City University of Hong Kong is described. Examples of new applications include enhancement of
Biomaterials
antimicrobial properties and cytocompatibility of plasma and surface-treated and nanostructured biomaterials,
Ti-based alloys
Biomedical polymers
corrosion resistance of plasma-treated biodegradable metals, as well as targeted drug delivery capability and
magnetic properties of surface-modified silica nanospheres and polymeric micelles.
© 2012 Elsevier B.V. All rights reserved.

1. Introduction 2. Antimicrobial properties and cytocompatibility of


surface-treated and nano-structured biomaterials
Many biological reactions occur in a physiological medium
containing various ions and liquids. For surgically implanted biomate- Titanium-based implants such as joint prostheses, fracture fixation
rials, the important reactions occur at the interface between the devices, and dental implants are commonly used in surgeries [17], but
materials and tissues/body fluids, and the surface properties of bioma- ensuing bacterial infection constitutes one of the biggest complications
terials determine their suitability. Properties of relevance include and implant failure. Bacteria can form a biofilm on the implant surface
physiochemical, topographical, and mechanical characteristics, biocom- subsequently leading to bacterial infection and related deleterious
patibility, as well as bio-functionalities. Most bulk materials are not nat- effects [18]. Therefore, it is crucial to prevent initial bacteria adhesion
ural biomaterials. That is to say, they may possess excellent mechanical onto the implant surface by rendering the biomaterials antibacterial.
properties, but they may not interface well with biological tissues and Silver is a biocide that is known to kill a broad spectrum of bacterial
fluids in vivo. Hence, surface modification plays a very important role and fungal species including antibiotic resistant strains [19]. In this
because selected surface properties such as cytocompatibility and anti- respect, silver nanoparticles (Ag NPs) have been shown to be more
microbial characteristics can be enhanced while the relevant bulk mate- reactive than the bulk counterpart because of the larger active surface
rials attributes like mechanical strength, robustness, and inertness can area [20].
be preserved. Ag plasma immersion ion implantation (Ag-PIII) has been employed
Plasma-based techniques such as plasma immersion ion implanta- to enhance the antibacterial properties of polymeric materials [21,22]
tion and deposition (PIII&D) [1,2] are widely used in the surface and recently applied to titanium in our laboratory to investigate the bio-
modification of various types of materials and industrial components, logical actions of Ag nanoparticles (NPs) on two strains of bacteria,
including biomaterials [3–13]. In this mini-review, recent work related namely Staphylococcus aureus and Escherichia coli, as well as mammalian
to biomaterials research in the Plasma Laboratory of City University of cells (osteoblast-like cell line MG63) [23]. During Ag-PIII, Ag NPs are cre-
Hong Kong [14–16] is described. Topics discussed include antimicrobial ated and embedded in titanium. The nanoparticles precipitate on and be-
properties and cytocompatibility of surface and plasma-treated neath the titanium surface via a local nucleation process in the solid
biometals and nano-structured biomaterials, corrosion resistance of solution of α-Ti(Ag). The morphology of S. aureus and E. coli seeded on
plasma-treated biodegradable metals, as well as targeted drug the untreated commercial Ti (Cp Ti) and Ag-PIII surfaces is examined
delivery capability and magnetic properties of surface-treated silica by SEM and the micrographs are depicted in Fig. 1. Bacteria division pre-
nanospheres and polymeric micelles. dominates on the Cp Ti surface but not on the Ag-PIII samples, indicating
that the Ag-PIII surfaces are cytostatic or cytocidal to bacteria. The at-
tached bacteria density measured on Cp Ti, 0.5h-Ag-PIII (Ag-PIII for 0.
5 h), 1.0h-Ag-PIII (Ag-PIII for 1 h), and1.5h-Ag-PIII (Ag-PIII for 1.5 h)
are 3.3×105, 8.1×104, 1.7×105, and 4.6×104 cells/mm2. The evolution
⁎ Tel.: +852 34427724; fax: +852 34420542, +852 34420538. pattern is consistent with that revealed by the zeta potentials at a pH of
E-mail address: paul.chu@cityu.edu.hk. 7.0 [23], indicating that bacteria attachment correlates with the surface

0040-6090/$ – see front matter © 2012 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.tsf.2012.07.144
94 P.K. Chu / Thin Solid Films 528 (2013) 93–105

Fig. 1. SEM morphology of the S. aureus species seeded on the various surface SEI images at (i) low magnification and (i-1) and high magnification together with (i-2) the corre-
sponding BES image (i = a, b, c, and d represent Cp Ti, 0.5h-Ag-PIII, 1.0h-Ag-PIII, and 1.5h-Ag-PIII, respectively and the bacteria concentration is 108 cfu/ml) [23].

zeta potential change. Binary and multiple fissions (black arrow in Fig. 1b-2, c-2, and d-2) is not as sharp as that on Cp Ti (Fig. 1a-2) and
Fig. 1a-2) are observed but not small dead-cell colonies (black arrow in some of them even merge with the background (white arrow in
Fig. 1a). However, more large dead bacteria colonies (black arrow in Fig. 1d-1 and d-2). The data indicate that bacteria death on the Ag-PIII
Fig. 1b, c, and d) and a small amount of fission (black arrow in Fig. 1b-2 samples probably arises from disruption of the bacterial membrane
and c-2) are observed on the Ag-PIII surfaces, implying that S. aureus integrity.
can proliferate well on Cp Ti but cannot survive on the Ag-PIII samples. To further assess the bactericidal effects after Ag-PIII, the bacteria
Furthermore, although some bacteria are found on the Ag-PIII surfaces, are detached from the samples, re-cultivated on agar, and evaluated
the membrane contrast (see the back-scattered electron images in by the bacteria counting method. Fig. 2 shows the bacteria colonies
P.K. Chu / Thin Solid Films 528 (2013) 93–105 95

Fig. 2. Re-cultivated bacterial colonies on agar: S. aureus colonies are detached from (a) Cp Ti, and (b) 0.5h-Ag-PIII at a re-cultivated bacteria concentration of 107 cfu/ml. The E. coli
colonies are dissociated from (c) Cp Ti and (d) 0.5h-Ag-PIII at a re-cultivated bacteria concentration of 106 cfu/ml [23].

on the Cp Ti and 0.5h-Ag-PIII samples. The amounts of S. aureus and proliferation on all the Ag-PIII surfaces is slightly lower than that on
E. coli on 0.5h-Ag-PIII are reduced by approximately 93% and 95% Cp Ti after culturing for 1 day [23]. However, after culturing for 3, 7,
after 24 h. The quantities of silver released from the Ag-PIII samples at and 9 days, the cells on all the Ag-PIII samples exhibit higher prolifer-
37 °C after 60 days are determined by inductively-coupled plasma optical ation rates and viability indicating that Ag-PIII introduces very little
emission spectroscopy (ICP-OES) to be less than a 10 ppb, thereby adverse or even more favorable effects on osteoblast proliferation.
providing evidence that leaching of Ag + from the NPs is minimal. The cell morphology on the Cp Ti and 0.5h-Ag-PIII samples after cul-
The antibacterial action of Ag NPs has been observed to depend on turing for 1 and 3 days is displayed in Fig. 3. Individual and clustered
the availability of silver ions [20]. Moreover, oxidizing species osteoblasts are observed to cover both surfaces and there is also com-
instead of reducing agents have been found to be responsible for plete cell spreading revealing good cytocompatibility after Ag-PIII.
the antibacterial effects [24]. The extremely low silver concentration Besides Al-PIII, titania nanotubes (TiO2-NTs) incorporated with
determined experimentally rules out the inherent toxicity of Ag + to silver (Ag) nanoparticles render good antimicrobial characteristics
bacteria, and it is believed that the antibacterial agent is metallic on Ti biomedical implants [26]. Bone tissues are composed of hierar-
(Ag 0) in lieu of Ag + released from the Ag NPs. The underlying factors chical nanocomposites and so the proper nano-topography can pro-
contributing to the antimicrobial results may be similar to that of the mote osteoconductivity from the biomimetic point of view [27–37].
silver and clay nanohybrids observed by Su et al. [25] that the Hence, ordered titania nanotubes (TiO2-NTs) fabricated on Ti im-
nanohybrids of silver and clay induce bacteria death via physical plants by anodization have attracted much attention [30–37] because
contact. TiO2-NTs have an elastic modulus of approximately 36–43 GPa [36]
Besides the antimicrobial properties, the surface cytocompatibility that is much closer to that of natural bones and are expected to
is another important factor affecting implant suitability and durability have better biomechanical compatibility than many current artificial
in vivo. It is assessed by culturing MG63 cells on the Cp Ti and Ag-PIII biomaterials. Our independent study has revealed that the hierarchi-
samples and determination using the alamarBlue™ assay. Cell cal hybrid TiO2 micropit/nanotube structures induce more balanced
96 P.K. Chu / Thin Solid Films 528 (2013) 93–105

Fig. 3. SEM pictures showing the MG63 cells cultured on Cp Ti after (a) 1 day and (c) 3 days and on 0.5h-Ag-PIII after (b) 1 day and (d) 3 days {cell density in the suspension =
1 × 105 cell/ml} [23].

promotion of multiple cell functions [30]. Therefore, TiO2-NTs are 24 h, there are large amounts of viable bacteria on the flat Ti and
promising bioactive coatings that can foster direct bone–implant smaller amounts on TiO2-NTs. In comparison, the amounts of viable
bonding and enhance the host defense on the implant surface, but bacteria are smaller on the NT-Ag samples and nearly no viable bacte-
again, their antimicrobial properties and cytocompatibility need to ria can be observed from NT-Ag1.5 and NT-Ag2.0.
be assessed and improved. In our experiments, Ag nanoparticles ad- The cytotoxicity results obtained by the LDH activity in the super-
here tightly to the wall of the TiO2-NTs prepared by immersion in a natants after incubation for 1 and 4 days are compared in Fig. 6a.
silver nitrate solution followed by ultraviolet light radiation, and the After the first day, neither TiO2-NTs nor NT-Ag exhibits obvious en-
amount of Ag in the NTs can be varied by changing the AgNO3 con- hancement in the LDH activity. After 4 days, NT-Ag0.5, NT-Ag1.0,
centration and immersion time. According to the time release profile and NT-Ag1.5 show slightly higher LDH activity than TiO2-NTs, but
[26], the amounts of Ag leached into the immersion fluid are quite the difference is statistically insignificant. However, higher LDH activ-
high but diminish gradually with immersion time. After two weeks, ity is observed from NT-Ag2.0. Hence, the NT-Ag samples can lead to
the amount of released Ag is about one-half of that measured in the some cytotoxicity that is related to the amount of incorporated Ag. All
first day. in all, the results demonstrate that Ag nanoparticles can be incorpo-
The antibacterial effectiveness against planktonic bacteria in the rated into TiO2-NTs on Ti implants using a simple procedure involving
medium (Rp) and antibacterial rates for adherent bacteria on the AgNO3 immersion and UV irradiation. The Ag nanoparticles adhere
specimens (Ra) for as long as 30 days is shown in Fig. 4b and c, re- strongly to the inner walls of the nanotubes along the entire length
spectively. The TiO2-NTs have an Rp value of about 30% which is con- and the size and amount of Ag nanoparticles can be regulated by
stant with time. Compared to TiO2-NTs, the NT-Ag samples (number varying the AgNO3 concentration and immersion time. The NT-Ag
following NT-Ag in the sample designation is the AgNO3 concentra- samples effectively kill all the planktonic bacteria in the culture
tion) have higher Rp values of about 100% during the first 4 days. medium during the first several days and the capability to prevent
The Rp values of the NT-Ag samples decrease gradually and that of bacterial adhesion is maintained for 30 days. The only exception is
NT-Ag0.5 diminishes more quickly. After 30 days, the Rp value of when the Ag concentration is very small. Because the adopted
NT-Ag0.5 drops to 30% whereas those of the other NT-Ag samples experimental conditions are much harsher than those encountered
are about 50%. Ag incorporation is effective in preventing bacteria col- normally in vivo, the materials are expected to be effective for a
onization on the specimens throughout the 30 days as indicated by much longer time in actual practice. This long-term effect bodes
Fig. 4c. The samples have Ra values of 100% which do not decrease well for prevention of post-surgery infection. Although the NT-Ag
significantly during this time, with the exception of NT-Ag0.5 that ex- samples show some cytotoxicity, it can be reduced by controlling
hibits a big reduction after 10 days to a value of about 5% after the Ag release rate and the properties can be further tailored to
30 days. The other three NT-Ag samples still have Ra values of about obtain both long-term antibacterial ability and biointegration. The
80% after 30 days. The capability of the NT-Ag samples to prevent materials are thus very attractive to biomedical implants due to
viable bacteria colonization is verified by fluorescence staining as prevention of implant associated infection and promotion of
shown in Fig. 5. After 7 days of repeated bacteria invasion every osseointegration.
P.K. Chu / Thin Solid Films 528 (2013) 93–105 97

surgical implantation, can be too quick and must be improved. Recently,


a new magnesium alloy, Mg–3Nd–0.2Zn–0.4Zr, has been shown to have
excellent mechanical properties and undergo uniform corrosion in the
physiological environment [45]. Although a surface coating can mitigate
corrosion on magnesium alloys [46–49], layer delamination and interfa-
cial mismatch can cause failure and many coatings are also not biode-
gradable. In comparison, ion implantation has many advantages. For
instance, the risk of layer delamination is drastically reduced due to
the graded interface and the surface properties including biocompati-
bility and cytotoxicity can be selectively controlled by implanting a
suitable amount of the proper elements. In particular, oxygen plasma
immersion ion implantation (O-PIII) has been demonstrated to improve
the corrosion resistance of AZ31 magnesium alloy [50]. This plasma
technique has recently been applied in our laboratory to modify the sur-
face of Mg–3Nd–0.2Zn–0.4Zr alloy and the synergistic effects offered by
surface alloying by metal (Al and Cr) ion implantation prior to O-PIII
have also been studied [51]. Atomic force microscopy reveals no appre-
ciable difference in the surface morphology before and after O-PIII. Fig. 7
shows the polarization curves of the samples in simulated body fluids.
The cathodic polarization curve represents cathodic hydrogen evolution
via water reduction, whereas the anodic curve indicates materials dis-
solution. It can be observed that after O-PIII, no significant shift in the
corrosion potential occurs and the corrosion current density decreases
slightly. After Al or Cr surface alloying, the corrosion potential becomes
higher. In particular, after Cr addition, the corrosion current density is
reduced implying that Cr surface alloying enhances the surface corro-
sion resistance of the O-PIII alloy.
Electrochemical impedance spectra (EIS) are further acquired to
determine the electrochemical reaction between the sample and sim-
ulated body fluids [SBF]. Fig. 8 displays the EIS spectra as well as two
equivalent circuit models to fit the data. According to the polarization
and EIS results, the protective effect rendered by the surface layer on
the Mg–Nd–Zn–Zr alloy is not significant although O-PIII produces
thicker surface-modified layers. According to the model used to sim-
ulate the EIS data, transverse propagation instead of longitudinal pen-
etration occurs in the weak regions in the surface layer in the SBF,
implying that the surface layer consisting of mainly magnesium
oxide is not stable. After surface alloying, the electrochemical corro-
sion behavior is altered as indicated by the decreased corrosion cur-
rent density and increased charge transfer. Usually, magnesium
oxide is more easily formed on magnesium alloys than other metal
oxides [52] but here, the surface composition changes due to metal
ion implantation and O-PIII replaces natural oxidation. The energetic
oxygen ions penetrate the surface and react with Mg and Cr (or Al).
On account of the existence of Cr (or Al) in the modified layer, Mg dif-
fusion from the substrate to surface is impeded and metallic Mg in the
top surface is finally depleted by oxidation. Hence, Cr (or Al) coexists
with Mg in the oxidized state in the surface layer. The surface oxide
film is more stable in SBF in comparison with the layers on the
untreated and O-treated sample according to the electrochemical
tests. The retardation effect rendered by Al addition is weaker than
that of Cr. All in all, although O-PIII increases the surface oxide thick-
ness on the Mg alloy, there is no significant improvement in the cor-
rosion resistance. On the other hand, surface alloying with Al or Cr
Fig. 4. (a) Non-cumulative silver release profiles from NT-Ag into PBS, (b) antibacterial
can further improve the performance of the O-PIII samples and the
rates against planktonic bacteria in the medium (Rp), and (C) antibacterial rates
against adherent bacteria on the specimen (Ra). The number after NT-Ag is the results demonstrate that surface alloying in conjunction with O-PIII
AgNO3 solution concentration [26]. is a promising approach to modify the electrochemical corrosion be-
havior of Mg alloys in the physiological environment.
The Mg-Y-Re (WE43) alloy is another promising biodegradable metal
suitable for cardiovascular stents [53–55]. Yttrium (Y) and rare earth
3. Corrosion resistance of plasma-treated biodegradable metals (Re) elements are incorporated into the magnesium alloy to strengthen
the materials and improve the corrosion resistance. In order to improve
Magnesium alloys such as AZ91, WE43 and LAE442 are potential the surface corrosion resistance, an Al2O3 layer is produced on the WE43
biodegradable materials due to their natural biodegradability and magnesium alloy by PIII. To overcome the mismatch in the mechanical
Young's moduli that are similar to those of human bones [38–44]. properties between Al2O3 and Mg as well as the negative electrochemical
However, their corrosion resistance, especially in the early stage after potential of Mg [56,57], an implanted Al layer serves as a transition layer
98 P.K. Chu / Thin Solid Films 528 (2013) 93–105

Ti TiO2-NTs

NT-Ag0.5 NT-Ag1.0

NT-Ag1.5 NT-Ag2.0

Fig. 5. Images revealing the bacterial viability after incubation for 7 days according to acridine (orange) and ethidium bromide fluorescence staining, with the living bacteria are
green while dead ones appear orange [26].

between the Al2O3 layer and magnesium substrate [58]. Fig. 9 shows the 4. Targeted drug delivery capability and magnetic properties of
potentiodynamic polarization curves of the samples in SBF. According to surface-modified silica nanospheres and polymeric micelles
the calculation, the untreated sample has a negative corrosion potential
in SBF of −1.972 V and in comparison with the untreated sample, the Nano-medicine is on the verge of revolutionizing disease diagnosis
corrosion potential of the implanted sample is nobler. The corrosion re- and therapy [60–64]. Among the various nanomaterials, hollow silica
sistance is related to the corrosion current density. The corrosion current nanoparticles with the mesoporous structure are especially attractive
density of the implanted sample is 4.468×10−5 A/cm2 and much lower because of the large surface area, highly accessible pores, bioinertness,
than that of the un-implanted sample (6.025×10−4 A/cm2) suggesting and biocompatibility. Moreover, they can serve as efficient carriers for
reduced magnesium degradation. The high degradation rates on the small drug molecules, genes, and proteins [65–73]. Chitosan, a natural
magnesium alloys in SBF originate primarily from corrosion by chloride cationic polysaccharide composed of β-(1–4)-linked glucosamine
ions. Mg(OH)2 formed by the reaction between surface MgO and water units together with N-acetyl-D-glucosamine units obtained by exhaus-
is slightly soluble and chloride ions can transform it into more soluble tive deacetylation of chitin [74], has favorable biodegradable, nontoxic,
MgCl2 [59]. On the other hand, Al2O3 is relatively stable in the aqueous and antimicrobial properties [75,76]. It is possible to combine the
medium and can effectively resist attack by chloride ions. Therefore, advantages of silica nanoparticles and chitosan. We have recently
the Al2O3 layer introduces a barrier to improve the corrosion resistance developed a one-step method to fabricate mono-dispersed SiO2
of the WE43 alloy in chloride-containing solutions. Although the modi- HNPs (hollow nanospheres) and then decorate them with chitosan
fied layer is not very thick, significant improvement in the corrosion using the surface cross-linking reaction with (3-glycidyloxypropyl)
resistance can be achieved. trimethoxysilane (GTPMS) [77]. The mono-dispersed cationic PS
P.K. Chu / Thin Solid Films 528 (2013) 93–105 99

Fig. 8. Electrochemical impedance spectra (EIS) obtained from the Mg–Nd–Zn–Zr alloy
samples in SBF and equivalent circuit models to fit the corresponding curves [51].

2) to produce the desired nanocarriers for targeted TNF-α (tumor necro-


sis factor α) drug delivery to tumor cells. As shown in Fig. 10, in the acidic
micro-environment inside solid tumors, the chitosan polymer chains
swell in the medium opening the pores of the nanocarriers so that the
loaded drugs can be easily released from the nanocarriers. On the con-
trary, at a higher pH, the chitosan polymer chains are deprotonated and
collapse to form a shield layer on the porous surface on the nanocarriers.
This thus blocks and restricts drugs release from the hollow interior under
normal circumstances.
To demonstrate that the CS-SiO2 HNPs conjugated with antibody
molecules can target MCF-7 breast cancer cells in vitro, SEM is
conducted after the cells are treated with the CS-SiO2 HNPs conjugated
with/without the antibody. As shown in Fig. 11e and f, the CS-SiO2 HNPs
conjugated to the antibody can bind to the cell membrane even after
Fig. 6. (a) LDH activity in the culture media and (b) cell numbers determined by DNA washing (shown by red arrows in Fig. 11f). It may be because the anti-
assay after osteoblast culturing for 1 and 4 days on the specimens [26]. body molecules on the CS-SiO2 HNPs have the ability to bind to the an-
tigen ErbB 2 on the cell membranes of the MCF-7 breast cancer cells.
Therefore, the CS-SiO2 HNPs located at the cell membranes of the
(polystyrene) nanospheres are fabricated by emulsifier-free emulsion targeted cells can keep the local TNF-α concentration higher thereby
polymerization using PVP (polyvinylpyrrolidone) as the stabilizer and increasing the binding efficacy of the TNF-α to the receptors.
AIBA (α,α′-azodiisobutyramidine dihydrochloride) as the cationic initia- In vivo breast cancer therapy experiments are conducted on mice.
tor. The resulting SiO2 hollow nanoparticles (HNPs) with a pH-sensitive The drugs are intraperitoneally administered to the mice once every
polyelectrolyte layer are conjugated to the antibody molecule (to ErbB two days. After two weeks, no mice die and the tumors collected

Fig. 7. Polarization curves acquired from the treated and untreated Mg–Nd–Zn–Zr alloy
samples in SBF [51]. Fig. 9. Polarization curves of WE43 and Al–O implanted WE43 Mg alloy in SBF [58].
100 P.K. Chu / Thin Solid Films 528 (2013) 93–105

Fig. 10. Schematic diagram showing the formation of nanocarriers (CS-SiO2-TNF-α conjugated with antibody) and drug release behavior at different pH values [77].

from three groups are weighed. As shown in Fig. 12, when only the of the nanocarriers composed of CS-SiO2 HNPs, TNF-α, and antibody
TNF-α agent is intraperitoneally administered to the mice, the total has also been investigated by our group in vivo, and no inflammation
weight of the tumors in the TNF group (Fig. 12a) is nearly 90% of and leukocyte infiltration abnormality can be observed from our
that of the control group (Fig. 12c). However, the total weight of pathological study on viscera organs including the liver, spleen, intes-
the tumors from the nano group (Fig. 12b) is about 45% of that of tine, and kidney.
the control group (Fig. 12c) and smaller than that of the TNF group Magnetic nanoparticles (MNPs) play important roles in biotechnol-
(Fig. 12a). The results demonstrate that TNF-α loaded on the ogy and nanomedicine applications such as targeted drug delivery
CS-SiO2 HNPs can effectively suppress tumor growth when compared [79–81], bimolecular labeling and separation [82,83], and magnetic res-
to the situation when only TNF-α is administered. The enhanced onance imaging (MRI) [84–86]. In order to integrate MNPs into biolog-
tumor growth inhibition observed from the nano group is believed ical system, surface coatings are applied to improve the colloidal
to stem from sustained TNF-α release from the CS-SiO2 HNPs in stability, cytocompatibility, and immunogenicity of MNPs. As illustrated
vivo as well as accumulation of the drug nanocarriers inside the in Fig. 13, stable magnetomicelles can be produced by self-assembly of
tumor microenvironment. The nanocarriers can penetrate the micro- fluorine-containing amphiphilic poly(HFMA-g-PEGMA) copolymers
circulation system of the solid tumors into the extracellular matrix [HFMA stands for 2, 2, 3, 4, 4, 4-hexafluorobutyl methacrylate and
(ECM) through the recognition between antibody on the CS-SiO2 PEGMA is methoxy poly (ethylene glycol) monomethacrylate] with
HNPs and antigen-ErbB 2 epitope on the MCF-7 cancer cell mem- oleic acid modified Fe3O4 nanoparticles in an aqueous medium [87].
branes. Hence, the nanocarriers aggregate on the membrane of the On the surface-modified structure, the fluorocarbon segments form a
MCF-7 breast cancer cells. As described in the previous paragraph, stable hydrophobic core to adsorb oil-based nanoparticles, and the
on account of the acidic local environment around the cancerous PEGMA segments enhance the solubility in water thus reducing the
cells, the chitosan chains on the drug nanocarriers swell and the cytotoxicity of the hybrid micelles.
pores of the CS-SiO2 HNPs are opened. Consequently, the loaded To assess the use of magnetomicelles as nanocarriers in drug delivery,
TNF-α is released from the CS-SiO2 HNPs making the TNF-α concen- in vitro release of the hydrophobic drug, 5-FU, commonly administered
tration around the cell membrane higher than that in other areas. in tumor therapy [88,89] is evaluated under physiological conditions.
With the released TNF-α binding to the tumor necrosis factor recep- After 5-FU is incorporated into the magnetomicelles, the drug loading
tor (TNFR), the apoptosis signal transmitting pathway cascade is acti- content is calculated to be 20.94 wt.%. The in vitro drug release behavior
vated further enhancing the tumor suppression effect. However, of 5-FU from the magnetomicelles in PBS at a pH of 7.4 and 37 °C is illus-
when only the TNF-α protein molecules are administered to the trated in Fig. 14 with the release of free 5-FU in PBS under the same con-
TNF group, the killing efficacy diminishes. It may be because the ditions serving as control. The data reveal initial burst release from the
TNF-α protein molecules cannot persistently stay inside the core 5-FU-loaded magnetomicelles (~30 wt.% of the initial loaded amount)
area of the solid tumors due to abnormal vasculature [78]. The safety followed by sustained release (~30–83 wt.% of the initial loaded
P.K. Chu / Thin Solid Films 528 (2013) 93–105 101

Fig. 11. SEM images showing the in vitro targeted MCF-7 breast cancer cells: (a, b) Blank control group without nanospheres; (c, d) control group treated with the CS-SiO2 HNPs
without the antibody; (e, f) experimental group treated by the CS-SiO2 HNPs conjugated with the antibody. The higher magnification SEM images in (b), (d) and (f) show the se-
lected area in (a), (c) and (e), respectively. The CS-SiO2 HNPs conjugated with the antibody is located on the cell membrane (red arrows, f). The same concentration of the blank
CS-SiO2 HNPs, and CS-SiO2 HNPs conjugated with the antibody is used to treat the MCF-7 breast cancer cells in the wells [77].

amount). The initial burst release may be ascribed to drug molecules lo- showing that the magnetomicelles lead to an appreciable negative
cated within the hydrophilic shell or at the interface between the micelle contrast enhancement (dark signal) in MRI. The efficacy of an MRI
core and shell [90]. In contrast to the results obtained from the free re- negative contrast agent is usually evaluated by the transverse
lease of 5-FU, the drug release rate from the magnetomicelles is impeded relaxivity rate, r2 (1 / T2), which presents the efficiency of the magne-
by the interactions between the hydrophobic drug molecules and hydro- tite nanoparticles to shorten the proton relaxation time. The relaxa-
phobic segments of the magnetomicelles. The results demonstrate that tion rates shown in Fig. 15b indicate that they vary linearly with
the magnetomicelles can indeed be used as nanocarriers in drug iron concentrations. In general, particulate MRI contrast agents de-
delivery. rived from magnetic nanoparticles such as Feridex and Resovit accu-
Iron oxide nanoparticles are known to shorten the transverse re- mulate in the liver and spleen thus facilitating the use of MRI to
laxation time of water protons and have been used as a negative con- diagnose liver and spleen diseases. To demonstrate the feasibility of
trast agent in magnetic resonance imaging (MRI) [91–93]. To evaluate these magnetomicelles as in vivo MRI probes, mice models are used.
the ability of these magnetomicelles to enhance MRI, the T2-weighted The in vivo MRI of a mouse liver with magnetomicelles as the contrast
MR images for different iron concentrations in an aqueous medium agent is exhibited in Fig. 16a. The liver is darkened significantly after
are acquired on a clinical 3 T MRI instrument. As shown in Fig. 15a, 10 min of magnetomicelle injection and a high contrast of liver tissue
sensitive and concentration-dependent dark areas can be seen from persists during the 4-h observation, thereby furnishing evidence that
the MR images of the aqueous medium with the magnetomicelles. the magnetomicelles are indeed effective T2 MRI contrast agents in
The decreased intensity is significant at high iron concentrations, MR diagnostic imaging.
102 P.K. Chu / Thin Solid Films 528 (2013) 93–105

Fig. 12. Photographs of tumors collected from mice in the three groups: (a) TNF group: only TNF-α agent administered to the mice; (b) nano group: nanocarriers composed of
CS-SiO2 HNPs and TNF-α and antibody injected into the mice; (c) control group: PBS (pH 7.4) injected into the mice [77].

5. Conclusion and cytocompatibility of plasma and surface-treated and nanostruc-


tured biomaterials, corrosion resistance of plasma-treated biodegrad-
Plasma-based and related surface modification methods are able metals, and targeted drug delivery capability and magnetic
important to biomaterials research and biomedical engineering. properties of surface-modified silica nanospheres and polymeric
New applications include enhancement of antimicrobial properties micelles. Significant applications are constantly being developed and

Fig. 13. Schematic showing the formation of magnetomicelles in assembly of fluorine-containing amphiphilic poly (HFMA-g-PEGMA) copolymers and oleic acid modified Fe3O4
nanoparticles in an aqueous solution [87].
P.K. Chu / Thin Solid Films 528 (2013) 93–105 103

[4] X. Lu, S.S.K. Iyer, C.M. Hu, N.W. Cheung, J. Min, Z.N. Fan, P.K. Chu, Appl. Phys. Lett.
71 (1997) 2767.
[5] P.K. Chu, S.B. Felch, P. Kellerman, F. Sinclair, L.A. Larson, B. Mizuno, Solid State
Technol. 42 (1999) 77 (55 and 42 (1999)).
[6] P.K. Chu, J.Y. Chen, L.P. Wang, N. Huang, Mater. Sci. Eng., R 36 (2002) 143.
[7] D.R. McKenzie, K. Newton-McGee, P. Ruch, M.M. Bilek, B.K. Gan, Surf. Coat.
Technol. 186 (1–2) (2004) 245.
[8] M. Ueda, K.G. Kostov, A.F. Beloto, N.F. Leite, K.G. Grigorov, Surf. Coat. Technol. 186
(1–2) (2004) 295.
[9] P. Chen, S.S. Lau, P.K. Chu, K. Henttinen, T. Suni, I. Suni, N.D. Theodore, T. Alford,
J.W. Mayer, L. Shao, M. Nastasi, Appl. Phys. Lett. 87 (2005) 111910.
[10] X.Y. Liu, P.K. Chu, C.X. Ding, Mater. Sci. Eng., R 47 (2–4) (2004) 49.
[11] Z.F. Di, P.K. Chu, M. Zhu, R.K.Y. Fu, S.H. Luo, L. Shao, M. Nastasi, P. Chen, T.L. Alford,
J.W. Mayer, M. Zhang, W.L. Liu, Z.T. Song, C.L. Lin, Appl. Phys. Lett. 88 (2006) 142108.
[12] X.Y. Liu, P.K. Chu, C.X. Ding, Mater. Sci. Eng., R 70 (3–6) (2010) 275.
[13] P.K. Chu, Surf. Coat. Technol. 204 (18–19) (2010) 2853.
[14] P.K. Chu, B.Y. Tang, Y.C. Cheng, P.K. Ko, Rev. Sci. Instrum. 68 (4) (1997) 1866.
[15] P.K. Chu, S. Qin, C. Chan, N.W. Cheung, P.K. Ko, IEEE Trans. Plasma Sci. 26 (1) (1998) 79.
[16] P.K. Chu, B.Y. Tang, L.P. Wang, X.F. Wang, S.Y. Wang, N. Huang, Rev. Sci. Instrum.
72 (3) (2001) 1660.
[17] M. Geetha, A.K. Singh, R. Asokamani, S.K. Gogia, Prog. Mater. Sci. 54 (2009) 397.
[18] J.W. Costerton, P.S. Stewart, E.P. Greenberg, Science 284 (1999) 1318.
Fig. 14. Release profiles of free 5-FU and 5-FU-loaded magnetomicelles in PBS at a pH [19] A. Agarwal, T.L. Weis, M.J. Schurr, N.G. Faith, C.J. Czuprynski, J.F. McAnulty, C.J.
of 7.4 and 37 °C [87]. Murphy, N.L. Abbott, Biomaterials 21 (2010) 680.
[20] X. Chen, H.J. Schluesener, Toxicol. Lett. 176 (2008) 1.
[21] W. Zhang, Y. Luo, H.Y. Wang, J. Jiang, S.H. Pu, P.K. Chu, Acta Biomater. 4 (2008) 2028.
[22] W. Zhang, Y. Luo, H.Y. Wang, S.H. Pu, P.K. Chu, Surf. Coat. Technol. 203 (2009) 2550.
materials scientists and engineers must continue to cope with the [23] H.L. Cao, X.Y. Liu, F.H. Meng, P.K. Chu, Biomaterials 32 (3) (2011) 693.
ever increasing demands by the biomedical community. [24] C.N. Lok, C.M. Ho, R. Chen, Q.Y. He, W.Y. Yu, H. Sun, P.K.H. Tam, J.F. Chiu, C.M. Che,
J. Biol. Inorg. Chem. 12 (2007) 527.
[25] H.L. Su, C.C. Chou, D.J. Hung, S.H. Lin, I.C. Pao, J.H. Lin, F.L. Huang, R.X. Dong, J.J. Lin,
Acknowledgments Biomaterials 30 (2009) 5979.
[26] L.Z. Zhao, S.L. Mei, W. Wang, P.K. Chu, Y.M. Zhang, Z.F. Wu, J. Biomed. Mater. Res.
Part A 96 (1) (2011) 100.
The work was financially supported by Hong Kong Research
[27] T.J. Webster, J.U. Ejiofor, Biomaterials 25 (2004) 4731.
Grants Council (RGC) General Research Funds (GRF) CityU 112510 [28] C.S. Hajicharalambous, J. Lichter, W.T. Hix, M. Swierczewska, M.F. Rubner, P.
and RGC Special Equipment Grant SEG_CityU05. Rajagopalan, Biomaterials 30 (2009) 4029.
[29] G. Mendonca, D.B. Mendonca, F.J. Aragao, L.F. Cooper, Biomaterials 29 (2008) 3822.
[30] L.Z. Zhao, S.L. Mei, P.K. Chu, Y.M. Zhang, Z. Wu, Biomaterials 31 (2010) 5072.
References [31] L.Z. Zhao, S.L. Mei, W. Wang, P.K. Chu, Z. Wu, Y.M. Zhang, Biomaterials 31 (2010)
2055.
[1] J.R. Conrad, J.L. Radtke, R.A. Dodd, F.J. Worzala, N.C. Tran, J. Appl. Phys. 62 (1987) 4591. [32] S.H. Oh, R.R. Finones, C. Daraio, L.H. Chen, S. Jin, Biomaterials 26 (2005) 4938.
[2] In: A. Anders (Ed.), Handbook of Plasma Immersion Ion Implantation and Depo- [33] S.H. Oh, C. Daraio, L.H. Chen, T.R. Pisanic, R.R. Finones, S. Jin, J. Biomed. Mater. Res.
sition, Wiley, New York, 2000. Part A 78 (2006) 97.
[3] P.K. Chu, S. Qin, C. Chan, N.W. Cheung, L.A. Larson, Mater. Sci. Eng., R R17 (6–7) [34] K.S. Brammer, S.H. Oh, C.J. Cobb, L.M. Bjursten, H. van der Heyde, S. Jin, Acta
(1996) 207. Biomater. 5 (2009) 3215.

Fig. 15. (a) T2-weighted MR images acquired from the aqueous solution containing the magnetomicelles and different concentrations of Fe; (b) T2 relaxation rate (1 / T2) as a func-
tion of Fe concentrations [87].
104 P.K. Chu / Thin Solid Films 528 (2013) 93–105

Fig. 16. (a) T2-weighted images of mice at different time points before and after tail vein administration of magnetomicelles and (b) relative signal enhancement values in the liver,
kidney, spleen, and muscle before and after injection of the magnetomicelles [87].

[35] K. Das, S. Bose, A. Bandyopadhyay, J. Biomed. Mater. Res. Part A 90 (2009) 225. [62] D. Peer, J.M. Karp, S. Hong, O.C. FaroKhzad, R. Margalit, R. Langer, Nat.
[36] G.A. Crawford, N. Chawla, K. Das, S. Bose, A. Bandyopadhyay, Acta Biomater. 3 (2007) Nanotechnol. 2 (2007) 751.
359. [63] J.L. West, N.J. Halas, Annu. Rev. Biomed. Eng. 5 (2003) 285.
[37] K.C. Popat, M. Eltgroth, T.J. Latempa, C.A. Grimes, T.A. Desai, Biomaterials 28 [64] J.H. Lee, Y.M. Huh, Y. Jun, J. Seo, J. Jang, H.T. Song, S.J. Kim, E.J. Cho, H.G. Yoon, J.S.
(2007) 4880. Suh, J.W. Cheon, Nat. Med. 13 (2007) 95.
[38] F. Witte, N. Hort, C. Vogt, S. Cohen, K.U. Kainer, R. Willumeit, F. Feyerabend, Curr. [65] Y.F. Zhu, J.L. Shi, W.H. Shen, X.P. Dong, J.W. Feng, M.L. Ruan, Y.S. Li, Angew. Chem.
Opin. Solid State Mater. Sci. 12 (2008) 63. Int. Ed. 44 (2005) 5083.
[39] B. Zberg, P.J. Uggowitzer, J.F. Löffler, Nat. Mater. 8 (2009) 887. [66] X.M. Jiang, T.L. Ward, Y.S. Cheng, J.W. Liu, C.J. Brinker, Chem. Commun. 46 (2010) 3019.
[40] X. Gu, Y. Zheng, Y. Cheng, S. Zhong, T. Xi, Biomaterials 30 (2009) 484. [67] Y.F. Zhu, J.L. Shi, W.H. Shen, H.R. Chen, X.P. Dong, M.L. Ruan, Nanotechnology 16
[41] X. Gu, Y. Zheng, S. Zhong, T. Xi, J. Wang, W. Wang, Biomaterials 31 (2010) 1093. (2005) 2633.
[42] E. Zhang, L. Xu, K. Yang, Scr. Mater. 53 (2005) 523. [68] N.E. Botterhuis, Q.Y. Sun, P. Magusin, R.A. van Santen, N. Sommerdijk, Chem. Eur.
[43] H.M. Wong, W.K. Yeung, K.O. Lam, V. Tam, P.K. Chu, D.K. Luk, M.C. Cheung, Bioma- J. 12 (2006) 1448.
terials 31 (2010) 2084. [69] J. Yang, J. Lee, J. Kang, K. Lee, J.S. Suh, H.G. Yoon, Y.M. Huh, S. Haam, Langmuir 24
[44] F. Witte, V. Kaese, H. Haferkamp, E. Switzer, A. Meyer-Lindenberg, C.J. Wirth, H. (2008) 3417.
Windhagen, Biomaterials 26 (2005) 3557. [70] E.Y. Yan, Y. Ding, C.J. Chen, R.T. Li, Y. Hu, X.Q. Jiang, Chem. Commun. 19 (2009) 2718.
[45] X. Zhang, G. Yuan, L. Mao, J. Niu, P. Fu, W. Ding, J. Mech. Behav. Biomed. Mater. [71] C.L. Lay, H.Q. Liu, D.C. Wu, Y. Liu, Chem. Eur. J. 16 (2010) 3001.
(2011), http://dx.doi.org/10.1016/j.jmbbm.2011.05.026. [72] J. Kim, H.S. Kim, N. Lee, T. Kim, H. Kim, T. Yu, I.C. Song, W.K. Moon, T. Hyeon,
[46] G. Song, Corros. Sci. 49 (2007) 1696. Angew. Chem. Int. Ed. 47 (2008) 8438.
[47] G. Wu, X. Zeng, G. Li, S. Yao, X. Wang, Mater. Lett. 60 (2006) 674. [73] J. Zhu, J.W. Tang, L.Z. Zhao, X.F. Zhou, Y.H. Wang, C.Z. Yu, Small 6 (2010) 276.
[48] X. Zeng, G. Wu, S. Yao, Mater. Lett. 60 (2006) 2252. [74] N. Bhattarai, J. Gunn, M.Q. Zhang, Adv. Drug Delivery Rev. 62 (2010) 83.
[49] G. Wu, W. Dai, H. Zheng, A. Wang, Surf. Coat. Technol. 205 (2010) 2067. [75] Y.B. Schuetz, R. Gurny, O. Jordan, Eur. J. Pharm. Biopharm. 68 (2008) 19.
[50] C. Wei, C.Z. Gong, X.B. Tian, S.Q. Yang, R.K. Fu, P.K. Chu, Plasma Sci. Technol. 11 [76] J. Wu, M. Sailor, Adv. Funct. Mater. 19 (2009) 733.
(2009) 33. [77] Z.W. Deng, Z.P. Zhen, X.X. Hu, S.L. Wu, Z.S. Xu, P.K. Chu, Biomaterials 32 (2011) 4976.
[51] G.S. Wu, K. Feng, A. Shanaghi, Y. Zhao, R.Z. Xu, G.Y. Yuan, P.K. Chu, Surf. Coat. [78] A.I. Minchinton, I.F. Tannock, Nat. Rev. Cancer 6 (2006) 583.
Technol. (2012), http://dx.doi.org/10.1016/j.surfcoat.2012.01.001. [79] J.D.G. Durán, J.L. Arias, V. Gallardo, A.V. Delgado, J. Pharm. Sci. 97 (2008) 2948.
[52] X.Q. Zeng, Q.D. Wang, Y.Z. Lu, W.J. Ding, Scr. Mater. 43 (2000) 403. [80] S. Purushotham, R.V. Ramanujan, Acta Biomater. 6 (2010) 502.
[53] A. Drynda, N. Deinet, N. Braun, M. Peuster, J. Biomed. Mater. Res. Part A 91A (2009) [81] J. Zhang, R.D.K. Misra, Acta Biomater. 3 (2007) 838.
360. [82] M. Takahashi, Y. Akiyama, J. Ikezumi, T. Nagata, T. Yoshino, A. Iizuka, K.
[54] F. Feyerabend, J. Fischer, J. Holtz, F. Witte, R. Willumeit, H. Drücker, C. Vogt, N. Yamaguchi, T. Matsunaga, Bioconjugate Chem. 20 (2009) 304.
Hort, Acta Biomater. 6 (2010) 1834. [83] Q.Q. Gai, F. Qu, Z.J. Liu, R.J. Dai, Y.K. Zhang, J. Chromatogr. A 1217 (2010) 5035.
[55] C. Mario, H. Griffiths, O. Goktekin, N. Peeters, J. Verbist, M. Bosiers, K. Deloose, B. [84] L. Zhang, H. Xue, C.L. Gao, L. Carr, J.N. Wang, B.C. Chu, S.Y. Jiang, Biomaterials 31
Heublein, R. Rohde, V. Kasese, C. Ilsley, R. Erbel, J. Interv. Cardiovasc. 17 (2004) (2010) 6582.
391. [85] R.Y. Hong, B. Feng, L.L. Chen, G.H. Liu, H.Z. Li, Y. Zheng, D.G. Wei, Biochem. Eng. J.
[56] Y.C. Xin, C.L. Liu, W.J. Zhang, J. Jiang, G.Y. Tang, X.B. Tian, P.K. Chu, J. Electrochem. 42 (2008) 290.
Soc. 155 (2008) C178. [86] K. Kluchova, R. Zboril, J. Tucek, M. Pecova, L. Zajoncova, I. Safarik, M. Mashlan, I.
[57] J. Zhang, W. Zhang, C. Yan, K. Du, F. Wang, Electrochim. Acta 55 (2009) 560. Markova, D. Jancik, M. Sebela, H. Bartonkova, V. Bellesi, P. Novak, D. Petridis, Bioma-
[58] Y. Zhao, G.S. Wu, H.B. Pan, K.W.K. Yeung, P.K. Chu, Mater. Chem. Phys. 132 (2012) terials 30 (2009) 2855.
187. [87] X.L. Li, H. Li, G.Q. Liu, Z.W. Deng, S.L. Wu, P.H. Li, Z.S. Xu, H.B. Xu, P.K. Chu, Bioma-
[59] Y.C. Xin, K.F. Huo, T. Hu, G.Y. Tang, P.K. Chu, Acta Biomater. 4 (2008) 2008. terials 33 (2012) 3013.
[60] M. Ferrari, Nat. Rev. Cancer 5 (2005) 161. [88] M. Prabaharan, J.J. Grailer, S. Pilla, D.A. Steeber, S. Gong, Macromol. Biosci. 9
[61] D. Jones, Nat. Rev. Drug Discovery 6 (2007) 174. (2009) 515.
P.K. Chu / Thin Solid Films 528 (2013) 93–105 105

[89] S. Sivakumar, V. Bansal, C. Cortez, S.F. Chong, A.N. Zelikin, F. Caruso, Adv. Mater. [92] S. Sun, Adv. Mater. 18 (2006) 393.
21 (2009) 1820. [93] J.H. Maeng, D.H. Lee, K.H. Jung, Y.H. Bae, I.S. Park, S. Jeong, Y.S. Jeon, C.K. Shim, W.Y.
[90] C. Allen, D. Maysinger, A. Eisenberg, Colloids Surf., B 16 (1999) 3. Kim, J.A. Kim, J.M. Lee, Y.M. Lee, J.H. Kim, W.H. Kim, S.S. Hong, Biomater 31 (2010)
[91] Y.W. Jun, J.W. Seo, J. Cheon, Acc. Chem. Res. 41 (2008) 179. 586.

You might also like