ACS Publications. Most Trusted. Most Cited. Most Read
My Activity
CONTENT TYPES

Figure 1Loading Img
RETURN TO ISSUEPREVResearch ArticleNEXT

Biomimetic Carbon Fiber Systems Engineering: A Modular Design Strategy To Generate Biofunctional Composites from Graphene and Carbon Nanofibers

  • Mohammadreza Taale
    Mohammadreza Taale
    Biocompatible Nanomaterials, Institute for Materials Science, Kiel University, Kaiserstraße 2, D-24143 Kiel, Germany
  • Fabian Schütt
    Fabian Schütt
    Functional Nanomaterials, Institute for Materials Science, Kiel University, Kaiserstraße 2, D-24143 Kiel, Germany
    More by Fabian Schütt
  • Tian Carey
    Tian Carey
    Cambridge Graphene Centre, University of Cambridge, 9 JJ Thomson Avenue, Cambridge CB3 0FA, U.K.
    More by Tian Carey
  • Janik Marx
    Janik Marx
    Institute of Polymer and Composites, Hamburg University of Technology, Denickestraße 15, D-21073 Hamburg, Germany
    More by Janik Marx
  • Yogendra Kumar Mishra
    Yogendra Kumar Mishra
    Functional Nanomaterials, Institute for Materials Science, Kiel University, Kaiserstraße 2, D-24143 Kiel, Germany
  • Norbert Stock
    Norbert Stock
    Institute of Inorganic Chemistry, Kiel University, Max-Eyth Straße 2, D-24118 Kiel, Germany
    More by Norbert Stock
  • Bodo Fiedler
    Bodo Fiedler
    Institute of Polymer and Composites, Hamburg University of Technology, Denickestraße 15, D-21073 Hamburg, Germany
    More by Bodo Fiedler
  • Felice Torrisi
    Felice Torrisi
    Cambridge Graphene Centre, University of Cambridge, 9 JJ Thomson Avenue, Cambridge CB3 0FA, U.K.
    More by Felice Torrisi
  • Rainer Adelung
    Rainer Adelung
    Functional Nanomaterials, Institute for Materials Science, Kiel University, Kaiserstraße 2, D-24143 Kiel, Germany
    More by Rainer Adelung
  • , and 
  • Christine Selhuber-Unkel*
    Christine Selhuber-Unkel
    Biocompatible Nanomaterials, Institute for Materials Science, Kiel University, Kaiserstraße 2, D-24143 Kiel, Germany
    *E-mail: [email protected]
Cite this: ACS Appl. Mater. Interfaces 2019, 11, 5, 5325–5335
Publication Date (Web):January 2, 2019
https://doi.org/10.1021/acsami.8b17627

Copyright © 2019 American Chemical Society. This publication is licensed under these Terms of Use.

  • Open Access

Article Views

3022

Altmetric

-

Citations

24
LEARN ABOUT THESE METRICS
PDF (9 MB)
Supporting Info (2)»

Abstract

Carbon-based fibrous scaffolds are highly attractive for all biomaterial applications that require electrical conductivity. It is additionally advantageous if such materials resembled the structural and biochemical features of the natural extracellular environment. Here, we show a novel modular design strategy to engineer biomimetic carbon fiber-based scaffolds. Highly porous ceramic zinc oxide (ZnO) microstructures serve as three-dimensional (3D) sacrificial templates and are infiltrated with carbon nanotubes (CNTs) or graphene dispersions. Once the CNTs and graphene coat the ZnO template, the ZnO is either removed by hydrolysis or converted into carbon by chemical vapor deposition. The resulting 3D carbon scaffolds are both hierarchically ordered and free-standing. The properties of the microfibrous scaffolds were tailored with a high porosity (up to 93%), a high Young’s modulus (ca. 0.027–22 MPa), and an electrical conductivity of ca. 0.1–330 S/m, as well as different surface compositions. Cell viability, fibroblast proliferation rate and protein adsorption rate assays have shown that the generated scaffolds are biocompatible and have a high protein adsorption capacity (up to 77.32 ± 6.95 mg/cm3) so that they are able to resemble the extracellular matrix not only structurally but also biochemically. The scaffolds also allow for the successful growth and adhesion of fibroblast cells, showing that we provide a novel, highly scalable modular design strategy to generate biocompatible carbon fiber systems that mimic the extracellular matrix with the additional feature of conductivity.

1. Introduction

ARTICLE SECTIONS
Jump To

Regenerative medicine aims at developing microenvironments for the regrowth of damaged or dysfunctional tissue and organs. New promising strategies of regenerative medicine make use of biomaterial scaffolds that resemble the chemical composition, (1) the topographical structure, and the three-dimensional (3D) micro- and nanoenvironments of extracellular matrix (ECM). (2) The ECM consists of interwoven protein fibers, such as collagens, in different ranges of diameters varying from a few (<5 nm) up to several hundred nanometers for bundled collagen fibrils. (3) The chemical, structural, and mechanical features of the ECM significantly control cell migration, as well as tissue development and maintenance (4) such that finding novel ways to mimic the ECM is a highly important task in biomaterials science. As the diameter (<5 nm) and length (<500 nm) of unbundled collagen fibrils are in the range of those of carbon nanotubes (CNTs), (3) CNTs present an interesting substitute for collagen fibrils.
A general goal of artificially fabricated biomaterial scaffolds is to promote cells to differentiate and proliferate in three dimensions so that they fulfill their functions in the artificial tissue and integrate well after implantation. (5) Particularly, the microstructure and porosity of a scaffold are the key to achieve spatially organized cell growth, besides the induction of specific biological functions in the regenerated tissue. Tailoring pore size, shape, and interconnectivity of a scaffold ensures that cell migration, as well as oxygen and nutrient contribution are similar to the conditions of natural tissues. (6) Often, a large pore size in the range of a few micrometers supports cell migration and ensures the transport of nutrition and waste products. (7)
In neural implants and heart tissue engineering, the electrical conductivity of a scaffold material is often a further requirement necessary for cellular signaling and function. (8) For example, conductivities of 0.03–0.6 S/m have been reported for cardiac muscles. (9) Carbon-based nanomaterials can in principle fulfill such requirements for 3D assemblies. (10,11) CNTs have a high electrical conductivity (up to 67 000 S/cm) (12) and chemical stability (e.g., against acids); (13) while graphene (G) offers high surface area (2630 m2/g) (14) and high electrical conductivity (107–108 S/m). (15) Both CNTs and graphene have attracted significant attention in biomedical applications, ranging from biosensors (16) and drug/gene delivery (17) to targeted bioimaging. (17) In addition, compared to other carbon-based nanomaterials, the high physical aspect ratio of CNTs (up to 3750 length/diameter) and graphene provides a sufficient surface area for the attachment of adhesion ligands and cells. (14,18) Regarding neural tissue engineering, 3D graphene foams (19) and graphene films (18) have contributed to enhance neural stem cell differentiation toward astrocytes and neurons. (20)
An important requirement for such carbon-containing scaffolds is biocompatibility. The biocompatibility of CNTs depends on the concentration of CNTs, (21) their degree of purification, synthesis method, (22) aspect ratio, diameter and number of CNT walls, (23) and their surface functionalization. (24) Although many studies have proven the feasibility of CNTs as biocompatible material, (25) the cytotoxicity of CNTs is still a concern due to residual metal catalysts, amorphous carbon, and CNT aggregation that can occur within the cell. (26) In contrast, graphene has been reported to be biocompatible and is readily applicable for a variety of biological applications, including the use of neuronal cells, (18) cardiomyocytes, (27) and osteoblast (28) cells. Graphene oxide (GO) is an interesting graphene derivative, as it consists of a single atomic carbon layer decorated with hydrophilic functional groups such as carboxylic acid, hydroxyl, and epoxide. (29) In particular, GO has shown a strong tendency to interact with peptides and proteins via physical or chemical bonds. (30) Therefore, graphene and GO have great potential in biomedical applications as they can easily be converted into biofunctional, peptide- or protein-coated surfaces. Moreover, carbon microtube materials, specifically aerographite (AG), are of further interest as biomaterials due to their electrical conductivity (0.2–0.8 S/m) (31) and highly porous (up to 99.99%) (32) 3D interconnected network. A recent study has demonstrated the feasibility of AG as a suitable 3D matrix for cell migration and proliferation. (33) However, a clear pathway to generate a highly porous biocompatible ECM-mimetic scaffold with tunable porosity, electrical conductivity, and suitable mechanics for biomedical applications has so far been missing.
Here we demonstrate a novel modular design strategy to generate hierarchically structured carbon-based, microfibrous scaffold materials with adjustable electrical and mechanical properties that mimic the structure of the extracellular matrix. The materials investigated here are biocompatible, support cell proliferation and adhesion, and open the gateway to future biomaterial development, where biocompatibility and electrical conductivity are vital for cell proliferation and stimulation.

2. Results and Discussion

ARTICLE SECTIONS
Jump To

2.1. Novel Types of Graphitic Scaffolds by CNT and Graphene Infiltration

Different types of fibrous 3D carbon scaffolds have been prepared based on our modular template-mediated method. Figure 1 shows the modular design strategy of our fabrication method. The fabrication uses presintered highly porous (porosity > 93%) ceramic ZnO templates as a sacrificial material. (32) These ZnO templates themselves consist of interconnected tetrapod-shaped ZnO particles. Representative scanning electron microscopy (SEM) images of the ZnO templates are shown in the Supporting Information Figure S2. The highly porous ZnO templates have an interstitial space of approximately 10–100 μm between ZnO filaments, which in turn have diameters between 0.5 and 5 μm. Therefore, the spatial geometry and organization of the microtube-shaped structures of the ZnO template are comparable to that of the ECM. (34)

Figure 1

Figure 1. Schematic illustration of different 3D carbon tube structures. The highly porous ZnO template can be either infiltrated with a nanoparticle dispersion (e.g., graphene, CNT) leading to a homogeneous coating around the tetrapodal particles or converted to a graphitic structure using a chemical vapor deposition (CVD) process (aerographite). The combination of both processes leads to a modular design strategy, especially in terms of conductivity, mechanical stiffness, and surface topography.

The coating of ZnO templates with carbon nanomaterials (e.g., CNTs, graphene) is performed via a simple process to infiltrate the entire 3D template with a CNT dispersion as described by Schütt et al. (35) In addition, here we demonstrate that the feasibility of this technique can be extended also to graphene dispersions. The infiltration process relies on the superhydrophilicity of the ZnO template, (36) which is a direct result of the combination of the hydrophilic character of the individual tetrapod-shaped ZnO microparticles and the high porosity (>93%) of the template. During water evaporation, the nanomaterials form a widely homogeneous coverage (35) around the ZnO microrods (Supporting Information Figures S3 and S4). The amounts of CNTs and graphene flakes covering the ZnO template can be controlled by cyclically repeating the infiltration process several times (Supporting Information, Figures S3 and S4). SEM images revealed that the infiltrated CNTs form a layer made of self-entangled CNTs around the ZnO network (Supporting Information Figure S4). Similarly to CNTs, a dispersion of graphene flakes also forms a homogeneous layer around the ZnO (Supporting Information Figure S3).
To generate freestanding scaffolds from the ZnO templates after coating them with carbon nanomaterials, the ZnO must be carefully removed. We have used three different processes to remove the sacrificial ZnO network after infiltration (Figure 1): (i) hydrolyzing the sacrificial ZnO template by a HCl solution (35) (Supporting Information Figure S5a), (ii) converting the ZnO template via a CVD process (31) (Supporting Information Figure S5b,c), and (iii) using a carbothermal reduction process (37) in combination with glucose (g) as a carbon source. Each of these processes leads to a specific type of carbon-based scaffold. Therefore, the results of these processes (Supporting Information Figure S5) are discussed in the following sections.

2.2. Freestanding Carbon Nanotube Network (CNTT) Scaffolds from Hydrolysis of the Sacrificial ZnO Template by HCl

When HCl is used to dissolve the ZnO in CNT-coated ZnO templates, hierarchically structured CNTT scaffolds are formed. These structures consist of interconnected hollow tubes, which are composed of self-entangled CNT networks, as reported previously. (35) SEM images of the resulting CNTT scaffolds are presented in Figure 2a–c. The presented structures demonstrate a hierarchical architecture, where the porous scaffold are composed of microtetrapods (Figure 2a,b), which in turn consist of nanoscale CNTs (Figure 2c). The HCl-based ZnO dissolution can only be applied for templates infiltrated with CNTs. Templates infiltrated with graphene only led to collapsing structures (data not shown), presumably as the graphene flakes cannot interweave.

Figure 2

Figure 2. SEM images from low to high magnification of the different 3D carbon structures. (a−c) Carbon nanotube tubes (CNTTs), (d−f) aerographite with incorporated CNTs (AG−CNTT) (the red arrows show the grown CNTs during the CVD process), (g−i) aerographite (AG), (j−l) aerographite with incorporated graphene (AG−G) (the yellow arrows point to nanopores on the surface of aerographite), and (m−o) carbon nanotube tubes incorporated into a thick carbon layer (CNTT−g).

2.3. ZnO Conversion to AG via CVD Forms Composites with Embedded Nanoparticles

We also use a CVD process to remove the ZnO, resulting in a thin (∼15 nm) film of graphite around the entire template similar to the graphitic shells in AG. (31) The CVD process can be applied if the ZnO is coated with either CNTs or graphene (Supporting Information Figure S6). Then, the AG serves as an additional backbone. This is the case in all of our CVD-based scaffolds, i.e., in composites of graphene and aerographite (AG–G) and in composites of multiwalled CNTs and aerographite (AG–CNTT) (Figure 2d–l). The modular design of the fabrication process allows us to change and tailor the surface topography of the hollow graphitic microtubes. Up to 5 μm long CNTs are formed perpendicular to the surface of microtubes on AG–G and AG–CNTT during the CVD process (Figure 2f). Such CNTs are not formed on pure AG (Figure 2i). (31) The growth of new carbon nanotubes on AG–G networks (Figure 2l, red arrows) is most likely attributed to the adsorption of carbon atom clusters on the active sites of the graphene surface during the CVD process. (38)

2.4. Carbothermal Reduction Process Leads to Novel Types of Graphitic Structures with Embedded CNTs

In the carbothermal reduction reaction, (37) glucose acts as the carbon source enabling the reduction of ZnO in a quartz tube furnace at 950 °C under argon atmosphere. To do so, ZnO templates were infiltrated with a mixture of glucose and CNTs. The carbothermal reduction of ZnO to Zn(g) (37) leads to ZnO removal and to the formation of graphitic shells, thus resulting in the so-called CNTT–g structure (Figure 2m–o). We have confirmed the removal of ZnO by Raman spectroscopy (Figure 3, CNTT–g graph) (see next paragraph). Furthermore, this process results in a scaffold with a microstructure that is comparable to that of the AG–CNTT scaffold (Figure 2e,n). However, the graphitic shells of the CNTT–g scaffold appear to be thicker than those of the AG–CNTT scaffold (Figure 2n,o). Additionally, in contrast to the AG–CNTT structures, no additional CNTs are grown (Figure 2n) in the CNTT–g case.

Figure 3

Figure 3. Raman spectroscopy of aerographite, ZnO, CNTT, AG–CNTT, CNTT–g, AG–G, graphene, and CNT inks.

Figure 3 shows Raman spectra (Renishaw 1000 InVia) of all of the carbon-based structures. Raman spectroscopy is used to examine the structural fingerprint of each material at a wavelength of 514.5 nm with an incident power of ∼0.1 mW. The graphene and CNT inks were drop-cast onto Si/SiO2 substrates before measurement. For the graphene ink (red curve), the G peak (∼1586) corresponds to the E2g phonon at the Brillouin zone center, and the D peak located at ∼1350 cm–1 corresponds to the breathing modes of the sp2 carbon atoms and requires a defect for its activation. (39) Our graphene inks are produced by liquid-phase exfoliation; therefore we attribute this peak to edge defects rather than to defects in the basal plane. (40) The two-dimensional (2D) peak located at ∼2700 cm–1 is the D peak overtone and can be fitted by a single Lorentzian, indicating electronically decoupled graphene monolayers. (41)
The aerographite (dark blue curve) shows a G peak position Pos(G) ∼ 1600 cm–1 and the absence of a distinct 2D peak, indicating the more defective nature of this sample and a lack of structural order in the aerographite. (42) The AG–G (yellow curve) structure has a similar spectrum to that of aerographite with D and G peaks located at ∼1350 and ∼1600 cm–1, respectively, indicating that the material is mostly composed of aerographite and graphene, given the selective etching of the ZnO scaffold during the CVD reduction. (31) The spectra of the CNT ink (green curve) and CNT ink with glucose (purple curve) show D, G, and 2D peaks at ∼1350, ∼1580, and ∼2700 cm–1, respectively. The peaks from the glucose residue are too weak to be observed. Moreover, the CNTT spectra (gray curve), AG–CNTT (blue curve), and CNTT–g (pink curve) have D, G, and 2D peaks that demonstrate the presence of the CNTs, respectively. Finally, the ZnO spectra (brown curve) display several peaks below 1200 cm–1, predominately created from the intense peak at 1158 cm–1 attributed to the 2A1(LO) and 2E1(LO) modes at the Brillouin zone center. (43) However, no peaks attributed to ZnO are observed in any of the CNTTs or aerographite scaffolds, proving the complete removal of the ZnO template.

2.5. Scaffold Mechanics can be Tailored Over Several Orders of Magnitudes

Figure 4 shows Young’s moduli and electrical conductivities of CNTT, AG–CNTT, CNTT–g, AG, and AG–G. The scaffold with the lowest stiffness is AG with a Young’s modulus of 16 kPa, which is comparable to previously reported values. (31) AG–G has a Young’s modulus of up to 27 kPa, thus adding graphene into the graphitic shells of AG results in a mechanical reinforcement of the scaffold (∼170%). This reinforcement is in a similar range to that of other graphene-reinforced porous networks, e.g., graphene/chitosan composites (44) (∼200%). In contrast to AG–G, AG–CNTT is much stiffer (Figure 4), with Young’s modulus reaching ∼22 MPa.

Figure 4

Figure 4. Young’s modulus (measured under compression) and electrical conductivity of the 3D CNTT, AG–CNTT, CNTT–g, and AG–G scaffolds. All structures containing CNTs have a higher Young’s modulus and conductivity compared to those without CNTs. The values for pure AG and CNTTs were taken from the corresponding publications, (31,35) whereas the other values were measured using a self-built electromechanical testing setup.

This reinforcement by a factor of about 1000 is presumably due to the reinforcement of CNTs on the nanoscale by self-entanglement. (35) Hence, the mechanical reinforcement of AG by CNTs is higher than in other CNT-reinforced porous biomaterials, including gelatin methacrylate (GelMA)–CNT composites (45) and poly(propylenefumarate)–CNT composites. (46) Young’s modulus of CNTT–g (∼4 MPa) is in between the moduli of AG–G and AG–CNTT. These results confirm that the incorporation of CNTs and graphene into 3D scaffolds compensates for the typically low mechanical Young’s moduli of porous structures (47) and that AG provides a stable backbone for CNTT scaffolds. In addition, the structural integrity of AG–G scaffolds is demonstrated during a long-cycle compression test (Figure S7) (Supporting Information Video S1).

2.6. Tailoring Scaffold Conductivity

We also investigated the electrical properties of our scaffolds (Figure 4, gray squares). AG–G and AG have similar electrical conductivities of around 0.5 and 0.2 S/m, whereas the conductivities of AG–CNTT and CNTT–g are about 120 and 130 S/m, respectively (Figure 4). Hence, AG–CNTT and CNTT–g clearly have higher electrical conductivities than CNT-containing electrospun fibrous composites (3.5 S/m), (48) which are applied in cardiac tissue engineering. The increase in conductivity of the scaffolds can be mainly attributed to the high conductivity of CNTs that are embedded in the graphitic shells of AG. This effect is more pronounced in the case of AG–CNTT and CNTT–g, presumably as a result of the conductive pathways formed by the self-entangled CNT networks. It has already been shown that by adjusting the CNT concentration during the infiltration process, the conductivity of CNTT can be tailored between 10–6 and 130 S/m. (35) Indeed, CNTT–g has the highest conductivity (330 S/m) of our scaffolds.

2.7. Scaffolds Strongly Adsorb Proteins

Albumin is an adhesive protein in plasma and can non-specifically bind to low-dimensional carbon-based materials via electrostatic interactions. (49) Therefore, we checked the albumin adsorption capacity of the scaffolds using the bicinchoninic acid (BCA) assay. As shown in Figure 5, within the first 48 h, albumin adsorption is smaller on AG–G than on AG–CNTT, whereas its adsorption is very similar on all scaffolds later on. The highest absolute protein adsorption mass (30.23–22.6 ± 2.76 mg/cm3) was detected during the first 2 days of incubation, and the adsorption amount reduced to approximately two-thirds (23.69–20.64 ± 2.39 mg/cm3) during the third to fourth day of incubation. Overall, protein adsorption is very similar on all tested scaffolds, although CNTT, CNTT–g, and AG–CNTT scaffolds adsorb slightly higher protein amounts (CNTT–g: 77.32 ± 6.95 mg/cm3; CNTT: 70.77 ± 5.33 mg/cm3; AG–CNTT: 68.08 ± 6.73 mg/cm3) than AG–G (64.92 ± 7.2 mg/cm3) (Figure 5a). This might be attributed to the higher protein adsorption capacity of CNTs than graphene flakes due to van der Waals forces and electrostatic interactions. (30)

Figure 5

Figure 5. Protein adsorption on CNTT, AG–CNTT, CNTT–g, and AG–G during 0–48, 49–96, and 97–144 h of incubation with albumin solution (1 mg/mL). (a) Absolute protein adsorption amount per scaffold volume. (b) Absolute protein adsorption amount per scaffold weight.

To compare our results of protein adsorption with other studies, we needed to relate them to the weight of the scaffolds by taking into account their density (Table 1). Figure 5b shows that protein adsorption per weight of CNTT, CNTT–g, and AG–CNTT scaffolds is different for different scaffold types (CNTT–g: 147.9 ± 13.42 mg/g; CNTT: 128.9 ± 9.69 mg/g; AG–CNTT: 115.14 ± 11.38 mg/g). It is also higher than on single-walled CNTs and graphene (∼100 mg/g) (30) and on nanoporous silica (∼70 mg/g). (50) In addition, due to its low density, AG–G adsorbs even more protein per weight (1512.25 mg/g) after 144 h, e.g., about 10 times more than our other scaffolds. The albumin adsorption on AG–G even after 48 h (527 mg/g) is comparable to that of graphene oxide (∼500 mg/g). (30)
Table 1. Full Names and Abbreviations of Fabricated Materials and Scaffolds in Our Study
full name abbreviation density (g/cm3) porosity (%)
aerographite AG ∼200 × 10–6  (31) up to ∼99.9 (31)
carbon nanotube tube CNTT ∼0.064 ∼94
carbon nanotube tube–glucose CNTT–g ∼0.061 ∼90
aerographite–carbon nanotube tube AG–CNTT ∼0.069 ∼95
aerographite–graphene AG–G ∼0.005 ∼96

2.8. Biocompatibility of the Carbon-Based Scaffolds

Biocompatibility of the scaffolds is investigated by methylthiazolyldiphenyl-tetrazolium bromide (MTT) metabolic activity and WST-1 assays, as well as by proliferation studies. Figure 6 shows the results for CNTT, CNTT–g, AG–CNTT, and AG–G samples using an MTT assay, demonstrating that all scaffolds are biocompatible, hosting a similar number of viable cells as the negative control. As cell adhesion is not possible on pristine AG without functionalization, (33) biocompatibility of AG is not investigated again in this study.

Figure 6

Figure 6. Percentage of viable cells (rat embryonic fibroblasts wild type, REF52wt) relative to the negative control, as determined in an MTT assay (four independent experiments, five technical repeats in each of them). The error bars denote standard deviation.

Figure 7 shows the proliferation rate of rat embryonic fibroblast cells (REF52wt) cultured on CNTT, CNTT–g, AG–CNTT, and AG–G samples relative to cells cultured on a culture dish. CNT-containing scaffolds (CNTT, CNTT–g, and AG–CNTT) lead to higher proliferation rates (ca. 300–320%) than graphene-containing structures (AG–G) (∼240%) after 7 days in culture. As fibroblast proliferation depends on matrix stiffness, (51) an increase in stiffness might also lead to an increase in fibroblast proliferation, (52) which can be attributed to the translation of mechanical cues from the matrix into a biochemical one via mechanosensory receptors such as focal adhesions. (53) Specifically, it has been reported that fibroblasts need substrates with a minimum Young’s modulus of 20–30 kPa to spread and 2 MPa to spread and polarize perfectly. (51) Hence, the huge difference in Young’s moduli between CNT-reinforced scaffolds (between 4 MPa in CNTT–g and 23 MPa in CNTT) and graphene-reinforced scaffolds (27 kPa in AG–G) can explain the higher proliferation rate of fibroblast cells on CNTT, AG–CNTT, and CNTT–g compared to AG–G.

Figure 7

Figure 7. Results of WST-1 metabolic tests of cell proliferation rate (REF52wt). Mean values were determined from four independent experiments, each including five technical repeats. The error bars denote standard deviation, the raw data were normalized to the control, and a correction factor was applied to account for unspecific adsorption (see Materials and Methods).

2.9. Carbon-Based Scaffolds as Porous Structures for Cell Growth

Fibroblast cells (REF52wt) were cultured for 7 days on the scaffolds to investigate cellular growth at the surface and inside. SEM images of critically point-dried cells (Figure 8) revealed that cells (highlighted in green) are attached to the surface of the scaffolds. Fibroblasts do not migrate strongly within a 3D network (54) but proliferate so that we typically observe several cells at one location. Cells are sprawled and elongated between the filaments of scaffolds and have a polygonal shape on all four scaffold types. Close-up images on the adhesion sites reveal tightly anchored membranes of cells to CNTs and graphene on the surface of structures, comparable to fibroblasts on AG functionalized with cyclic arginylglycylaspartic acid (RGD) peptides. (33)

Figure 8

Figure 8. SEM images of REF52wt cells after 7 days of culturing within (a–c) AG–CNTT, (d, e) CNTT–g, (g–i) AG–G, and (j–l) CNTT scaffolds. Left: Medium-sized overview images illustrating the growth of cells between the fibers in different directions and planes; middle: zoomed-in images showing well-stretched cells along the fibers and their elongation; and right: close-up images on the adhesion sites, proving the presence of strong contacts between the materials and the cell membrane (yellow arrows).

To investigate cell adhesion at the molecular level, we studied the presence of paxillin in adhesion structures. Paxillin is a component of focal adhesion clusters; (55) therefore, it can be assumed that more paxillin in contact with our scaffolds is related to stronger cell adhesion. We used cells that were stably transfected with yellow fluorescent protein (YFP)-paxillin and imaged the paxillin in fluorescence microscopy. Imaging fluorescence in 3D matrices compared to 2D is challenging. (56) Due to the low intensity of paxillin in the cells and the high light absorbance of our scaffolds, (31) imaging the paxillin adhesion sites was only possible by using long acquisition times (5 s), which resulted in background signals. Nevertheless, paxillin-containing adhesion sites can be distinguished around the filaments of the scaffolds. Based on the fluorescence images (Figures 9a–d and S8), more adhesion clusters can be detected on CNT-reinforced scaffolds than on graphene-reinforced scaffolds. This could again result from the different mechanical properties of the scaffolds, but it is also in agreement with cell studies on multiwalled CNTs showing that NIH-3T3 fibroblasts form larger adhesion clusters on CNTs than on graphene. (57)

Figure 9

Figure 9. High-magnification fluorescence images of REF52 YFP-paxillin cells on 3D scaffolds: (a) CNTT, (b) AG–CNTT, (c) CNTT–g, and (d) AG–G. YFP-paxillin is mainly distributed in small clusters (YFP; yellow), the nucleus was stained with Hoechst (4′,6-diamidino-2-phenylindole, DAPI; blue), and actin fibers were visualized by phalloidin (red). Fluorescence imaging took place in optical sections approximately between 50 and 100 μm from the surface of the material. Adhesion sites are detectable (a, b) as tiny yellow spots mainly around the tube-shaped filaments of CNTT and AG–CNTT structures. Due to the low intensity and small size of YFP-paxillin (c, d), they are not clearly observed, which is also obscured by red (actin fibers) and blue (nucleus) channels in multichannel images. The green dashed arrows illustrate the protrusion direction of the cells.

A further important contribution to cell adhesion is the cytoskeleton, where networks of actin fibers determine cell shape and movements. (58) To investigate cellular actin networks on our scaffolds, we investigated the fluorescence of phalloidin to detect actin fibers. Again, imaging deeply inside the scaffolds was impaired by the strong light absorbance of CNTs and graphene, restricting it to the first 300 μm from the surface. As shown in Figure 9, well-developed actin fibers (in red) are indeed present within the fibroblasts. Furthermore, the cells are polarized and have oriented actin bundles. Figure 9a–c also shows that actin fibers on CNT-reinforced scaffolds (CNTT, AG–CNTT, and CNTT–g) are mainly oriented in the direction of fibroblast protrusions. Although many actin fibers can be detected in cells grown on AG–G (Figure 9d), they are neither well polarized nor elongated like the actin fibers in the cells on CNT-reinforced scaffolds (CNTT, AG–CNTT, and CNTT–g). These results could again originate from the mechanical properties of our materials, similarly to proliferation rate and paxillin clusters.

2.10. ECM-Mimetic Scaffolds

The open porous structure of our scaffolds with large free volumes (>95%) should be highly beneficial for the growth and migration of cells, as the open pores allow the cells to freely migrate and proliferate within the scaffolds. (2) Moreover, in contrast to other studies on 3D porous structures, in which the alignment of CNTs (24) or graphene sheets (59) confined the accessibility of cavities, our carbon framework structures provide accessible interconnected pores from all sides (Figure 2). In addition, the hierarchical organization of structural elements, specifically self-entangled CNTs in the form of microtubes, is reminiscent of the hierarchical nano- and microstructure of the ECM. It should therefore in principle be possible to employ our modular design strategy to generate different composition-dependent structural and mechanical features similar to collagen (60) in the ECM. As our scaffolds strongly adsorb proteins, it should also be possible to adsorb adhesion ligands typically present in ECM proteins, such as RGD. (61) In this way, the scaffolds can be modified such that they finally mimic the ECM structurally and biochemically, but with the additional feature of conductivity.

3. Conclusions

ARTICLE SECTIONS
Jump To

In summary, we have introduced a novel modular design strategy to produce carbon-based scaffolds that mimic the ECM and allow 3D cellular growth. Biocompatibility studies revealed a high proliferation rate of fibroblasts as well as the ability of fibroblasts to develop paxillin-containing adhesion sites. In addition, the cells sprawl and elongate between single filaments of the scaffolds. Tuning electrical conductivity (ca. 0.1–330 S/m), stiffness (ca. 10–3–0.7 MPa), and protein adsorption and porosity (up to ∼99%) of the scaffold provides great possibilities for culturing cells. Based on the proven protein adsorption capacity of the scaffolds, they are suitable for biofunctionalization and addition of other biochemical cues. This is particularly relevant in tissue engineering of electrically excitable tissue, e.g., heart tissue, as our scaffolds have tunable electrical conductivity. In addition, the fabrication procedure is very simple and can in principle be adopted to develop 3D assemblies from other low-dimensional nanomaterials (e.g., bioactive ceramic nanoparticles, polymeric nanofibers) by only changing the nanoparticle dispersion, as long as the nanoparticles are connected via strong physical contacts such as entanglement, fusion, or physical locks. This makes the scaffolds promising candidates as conductive ECM-mimetic materials in many applications from regenerative medicine to 3D cell culture.

4. Materials and Methods

ARTICLE SECTIONS
Jump To

4.1. Fabrication of 3D Carbon Scaffold Materials

Templates of tetrapod-shaped ZnO were fabricated by a previously developed flame transport synthesis method. (32) The resulting loose powder was pressed into a cylindrical shape (h = 3 mm, d = 6 mm) at a density of 0.3 g/cm3. The pellets were subsequently sintered for 5 h at 1150 °C to obtain an interconnected 3D ZnO network. (32) This structure is the sacrificial template used for fabrication of carbon-based scaffolds, i.e., freestanding CNTT, AG–CNTT, and AG–G.
For the fabrication of the CNTT scaffolds, the porous ZnO templates were infiltrated with an aqueous dispersion of multiwalled CNTs (1 wt %, CARBOBYK-9810, BYK Additives & Instruments) using a self-built computer-controlled syringe. The length of the CNTs can be on the order of a few micrometers, with diameters in the range of 20–60 nm. (34) After infiltration of ∼90 μL of the CNT solution, the samples were dried under ambient conditions for at least 1 h. The process was repeated six times so that CNTs covered the ZnO template. Then, the ZnO backbone was removed by immersing the composite in a 1 M HCl solution overnight. The HCl solution was replaced by washing with pure ethanol (five times). Finally, the etched structures were dried in a critical point dryer (EMS 3000) by using the automatic mode. The purge time was set to 15 min to ensure that all ethanol was washed out, leaving freestanding CNTTs. (3)
To alter the CNTT structures further, we added glucose (1 wt %) to the CNT ink used for infiltration. The infiltration into a ZnO template was then repeated four times. After that, the samples were transferred into a quartz tube furnace and heated to 950 °C under argon atmosphere for 2 h. During this process, the ZnO was removed by carbothermal reduction, (45) leading to CNTT–g, which contains both the remnants of the glucose and embedded CNTs.
For the synthesis of AG–CNTT scaffolds, the CNT-coated ZnO templates were exposed to a CVD process, as reported previously. (31) Briefly, the ZnO template was replicated into aerographite at ∼760 °C under a hydrogen and argon atmosphere in the presence of toluene as the source of carbon. (62) Thereby, the CNTs were embedded into the graphite microtubes of aerographite while the ZnO was simultaneously etched by H2.
A similar process was used to generate composites of graphene and aerographite (AG–G). The graphene ink was made by dispersing graphite flakes (12 mg/mL, Sigma-Aldrich No. 332461) and Triton X-100 stabilization agent (1.7 mg/mL) in deionized water, followed by sonication (Fisherbrand FB15069, Max power 800 W) for 9 h. (63,64) The dispersion was then centrifuged (Sorvall WX100 mounting a TH-641 swinging bucket rotor) at 5k rpm for 1 h to remove the thick flakes. The supernatant (i.e., the top 70%) of the dispersion was then decanted to produce the final graphene ink. In Figure S1, optical absorption spectroscopy (Cary 7000) is used to estimate the flake concentration. (65) The flake concentration in the graphene ink is obtained via the Beer–Lambert law, which links the absorbance A = αcl with the beam path length l (m), the flake concentration c (g/L), and the absorption coefficient α (L/(g m)). The graphene ink was diluted 1:20 with water/Triton X-100. An absorption coefficient of α ∼ 1390 L/(g m) for the graphene ink at 660 nm was utilized, (66) estimating a graphene flake concentration of ∼0.09 mg/mL, consistent with previous reports of graphene-based inks. (40,67) The spectrum of the graphene ink is mostly featureless due to the linear dispersion of the Dirac electrons, while the peak in the UV region is a signature of the van Hove singularity in the graphene density of states. (68) The ZnO scaffold was infiltrated 50 times to achieve coverage of graphene around the ZnO tetrapods, due to the low concentration of the ink (0.09 mg/mL). After infiltration, the CVD process (see process specification above) was used to embed the graphene flakes into aerographite microtubes and to remove the ZnO template.

4.2. Mechanical and Electrical Characterizations of 3D Carbon Scaffold Materials

Mechanical and electrical characterizations were performed using a self-built setup consisting of a Maerzhaeuser Wetzlar HS 6.3 micromanipulator, which is driven by a stepper motor, a Kern PLE 310-3N precision balance, and a Keithley 6400 source meter. A self-written LabView program controls all components. To avoid any vibration damping, the whole setup is located on a very rigid aluminum plate in a box filled with sand, which is mounted on a vibration-isolated table. Stress–strain curves were measured by placing the sample in between the micromanipulator and the precision balance. For compression tests, the micromanipulator deforms the sample by a user-defined step size. After each step, the program measures the force of the balance after a short settling time has been elapsed. These steps are repeated until the maximum deformation defined by the user is reached. Afterwards, the direction of the deformation is inverted until the micromanipulator comes back to its original position. Finally, the stress–strain curves are evaluated and Young’s modulus is determined. With respect to the cyclic compression test, the same procedure was repeated several times. The wait time between each compression step was set to 0 s, and the deformation speed was set to 40 μm/s. To demonstrate the structural integrity, a video was recorded during the long-cycle compression test with a USB camera. Furthermore, the same setup also allows to record the current–voltage characteristics using a Keithley 6400 source meter in the four-wire sense mode. For electrical measurements, the carbon structures were connected to thin copper plates on both sides by using conductive silver paste. It is noteworthy that only a thin layer of paste was needed to ensure good electrical connection of the porous material to the measurement setup. Current–voltage curves were measured in a voltage range of up to 5 V. Finally, the resistance was calculated from the obtained data and converted to conductivity to give meaningful values for each structure.

4.3. Cell Culture and Cell Seeding on Scaffolds

Rat embryonic fibroblasts (REF52), both as wild type and stably transfected with YFP-paxillin (REF YFP-Pax), (69) were cultured in Dulbecco’s modification of Eagle medium (Biochrom, Germany) at 37 °C, 5% CO2 at ∼90% humidity. The medium was supplemented with 10% fetal bovine serum (Biochrom, Germany) and 1% penicillin/streptomycin (Sigma-Aldrich, Germany). To expel any traces of remaining zinc from the scaffold fabrication process prior to cell experiments, all samples were immersed in culture medium for 14 days after autoclaving at 121 °C. Shortly before the cell experiments, the cells were immersed in a fresh culture medium, counted with a cell counter (Scepter, Merck Millipore, Germany), and ∼20 000 cells were seeded on each scaffold in 24-well plates. The cells were incubated on the scaffolds for 1, 3, or 7 days.

4.4. Cell Staining

To investigate the cell morphology, proliferation, and adhesion on the scaffolds, cell nuclei were stained with DAPI (Thermo Fisher, Germany), and actin stress fibers were stained with phalloidin (Alexa Fluor 647 phalloidin, Thermo Fisher, Germany). Imaging was carried out using fluorescence microscopy (IX81, Olympus, Germany), and images were processed with cellSens Dimension (Olympus, Germany). For electron microscopy investigations, the cells were fixed by paraformaldehyde (Thermo Fisher, Germany) and dried using critical point drying (EMS 3000). A thin layer of gold was sputtered (Bal-Tec SCD 050, 30 mA, 30 s) onto the sample prior to scanning electron microscopy (SEM) (Ultra Plus Zeiss SEM, 5 kV). The cells in SEM images were highlighted in green (via Adobe Photoshop CC 2017) to distinguish them more easily.

4.5. Viability and Proliferation Assay

The number of living cells on the scaffolds was quantified by a WST-1 proliferation assay after 1, 3, and 7 days of incubation. In this assay, the number of living cells can be acquired from the amount of dye produced via bioreduction of stable tetrazolium salt WST-1 (Sigma-Aldrich, Germany). The amount of formazan is proportional to the number of cells (Cell Proliferation Reagent WST-1 Protocol, Sigma-Aldrich, Germany). The experiments were carried out as follows: After seeding the cells onto each sample (see specification above), the scaffolds were first washed with phosphate buffered saline and then incubated with a WST-1-containing medium for 4 h. The concentration of the formazan dye was quantified by a multiwell spectrophotometer (Bio-Tek μQuant) after removal of the samples from the wells. CNTs tend to react with tetrazolium salts; (70) thus, the proliferation rates were normalized to the absorption of the detected tetrazolium on control samples without cells. The amount of tetrazolium reacting with the scaffolds was determined for each sample as explained in the following: 10 000 cells were cultured for 24 h in 96-well plates prior to adding scaffolds to half of the wells, with the other half serving as scaffold-free control. Then, all of the wells were incubated for an additional time of 4 h with the WST-1 reagent before quantification of the formazan (i.e., product of tetrazolium cell reaction) amount in each well by a multiwell spectrophotometer. The amount of tetrazolium that reacted with the scaffold was calculated for each specimen by subtracting the amount of formazan of scaffold-containing wells from scaffold-free control wells. The difference in absorbance between the scaffolds and controls was used as a correction factor for the data generated with the WST-1 assay.
In addition, the viability of cells was tested according to the ISO 10993 norm. Briefly, 10 000 cells/100 μL REF52 cells were cultured in a 96-well plate for 24 h. For medium extraction, the scaffolds were incubated in a culture medium at 37 °C for 72 h. The cultured cells were incubated with either untreated medium or extracted medium for a further 24 h. To determine cell viability, the colorimetric methylthiazolyldiphenyl-tetrazolium bromide metabolic activity assay (MTT; Sigma-Aldrich, Germany) was used. The cells in untreated medium served as the negative control, and the cells in 20% dimethyl sulfoxide were the positive control. The absorbance was measured at 490 nm (absorption wavelength of formazan) and 600 nm as a reference. The results from the cultured cells with extracted medium were normalized to the values measured for the negative control.

4.6. Protein Adsorption Rate

The protein adsorption rate on the scaffolds was measured by using bovine serum albumin (Pierce; Thermo Fisher, Germany) as a model protein. Protein solution (1 mL, 1 mg/mL) was added per scaffold and incubated at 37 °C in a humidified incubator (CO2 5%, humidity 95%). After 48 h, nonadhered proteins were carefully removed by a pipette and saved for recording. The scaffolds were washed with saline and incubated for a further 48 h after addition of 1 mL of protein solution. The same procedure was repeated after 48 h. The concentration of protein in the supernatant was measured using a Micro BCA protein assay (Pierce; Thermo Fisher, Germany). To do so, 10 μL of supernatant was mixed with 200 μL of working reagent and incubated for 30 min. The absorbance was measured using a microplate reader (Bio-Tek μQuant) at 570 nm. After calibrating the results with a standard curve provided by the BCA protein assay kit (Pierce; Thermo Fisher, Germany), the protein adsorption rate was calculated by subtracting the residual protein concentration from the initial protein concentration.

Supporting Information

ARTICLE SECTIONS
Jump To

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsami.8b17627.

  • Optical absorption spectrum of graphene ink (Figure S1); SEM images of ZnO templates (Figures S2–S4); different 3D carbon tube structures (Figure S5); SEM images of AG–CNTT structures (Figure S6); long-cycle compression test graph of AG–G scaffold (Figure S7); high-magnification fluorescence image of REF52 YFP-paxillin cells (Figure S8); supporting discussion (PDF)

  • Structural integrity of AG–G scaffold (AVI)

Terms & Conditions

Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Author Information

ARTICLE SECTIONS
Jump To

  • Corresponding Author
  • Authors
    • Mohammadreza Taale - Biocompatible Nanomaterials, Institute for Materials Science and , Kiel University, Kaiserstraße 2, D-24143 Kiel, Germany
    • Fabian Schütt - Functional Nanomaterials, Institute for Materials Science and , Kiel University, Kaiserstraße 2, D-24143 Kiel, GermanyOrcidhttp://orcid.org/0000-0003-2942-503X
    • Tian Carey - Cambridge Graphene Centre, University of Cambridge, 9 JJ Thomson Avenue, Cambridge CB3 0FA, U.K.
    • Janik Marx - Institute of Polymer and Composites, Hamburg University of Technology, Denickestraße 15, D-21073 Hamburg, Germany
    • Yogendra Kumar Mishra - Functional Nanomaterials, Institute for Materials Science and , Kiel University, Kaiserstraße 2, D-24143 Kiel, GermanyOrcidhttp://orcid.org/0000-0002-8786-9379
    • Norbert Stock - Institute of Inorganic Chemistry, Kiel University, Max-Eyth Straße 2, D-24118 Kiel, GermanyOrcidhttp://orcid.org/0000-0002-0339-7352
    • Bodo Fiedler - Institute of Polymer and Composites, Hamburg University of Technology, Denickestraße 15, D-21073 Hamburg, Germany
    • Felice Torrisi - Cambridge Graphene Centre, University of Cambridge, 9 JJ Thomson Avenue, Cambridge CB3 0FA, U.K.Orcidhttp://orcid.org/0000-0002-6144-2916
    • Rainer Adelung - Functional Nanomaterials, Institute for Materials Science and , Kiel University, Kaiserstraße 2, D-24143 Kiel, Germany
  • Notes
    The authors declare no competing financial interest.

Acknowledgments

ARTICLE SECTIONS
Jump To

R.A. gratefully acknowledges partial project funding by the Deutsche Forschungsgemeinschaft under contract FOR2093. This project has received funding from the European Union’s Horizon 2020 research and innovation programme under grant agreement No. GrapheneCore2 785219. C.S.-U. was supported by the European Research Council (ERC StG 336104 CELLINSPIRED, ERC PoC 768740 CHANNELMAT) and the German Research Foundation (RTG 2154, SFB 1261 project B7). M.T. acknowledges support from the German Academic Exchange Service (DAAD) through a research grant for doctoral candidates (91526555-57048249). The authors acknowledge funding from EPSRC grants EP/P02534X/1, ERC grant 319277 (Hetero2D), the Trinity College, Cambridge, and the Isaac Newton Trust. We also acknowledge Galen Ream for critical proofreading.

References

ARTICLE SECTIONS
Jump To

This article references 70 other publications.

  1. 1
    O’Brien, F. J. Biomaterials & Scaffolds for Tissue Engineering. Mater. Today 2011, 14, 8895,  DOI: 10.1016/S1369-7021(11)70058-X
  2. 2
    Yang, S.; Leong, K.-F.; Du, Z.; Chua, C.-K. The Design of Scaffolds for Use in Tissue Engineering. Part I. Traditional Factors. Tissue Eng. 2001, 7, 679689,  DOI: 10.1089/107632701753337645
  3. 3
    Shoulders, M. D.; Raines, R. T. Collagen Structure and Stability. Annu. Rev. Biochem. 2009, 78, 929958,  DOI: 10.1146/annurev.biochem.77.032207.120833
  4. 4
    Friedl, P.; Wolf, K. Plasticity of Cell Migration: A Multiscale Tuning Model. J. Cell Biol. 2010, 188, 1119,  DOI: 10.1083/jcb.200909003
  5. 5
    Langer, R.; Vacanti, J. P. Tissue Engineering. Science 1993, 260, 920926,  DOI: 10.1126/science.8493529
  6. 6
    Place, E. S.; Evans, N. D.; Stevens, M. M. Complexity in Biomaterials for Tissue Engineering. Nat. Mater. 2009, 8, 457470,  DOI: 10.1038/nmat2441
  7. 7
    Loh, Q. L.; Choong, C. Three-Dimensional Scaffolds for Tissue Engineering Applications: Role of Porosity and Pore Size. Tissue Eng., Part B 2013, 19, 485502,  DOI: 10.1089/ten.teb.2012.0437
  8. 8
    Ganji, Y.; Li, Q.; Quabius, E. S.; Böttner, M.; Selhuber-Unkel, C.; Kasra, M. Cardiomyocyte Behavior on Biodegradable Polyurethane/Gold Nanocomposite Scaffolds under Electrical Stimulation. Mater. Sci. Eng., C 2016, 59, 1018,  DOI: 10.1016/j.msec.2015.09.074
  9. 9
    Potse, M.; Dubé, B.; Vinet, A. Cardiac Anisotropy in Boundary-Element Models for the Electrocardiogram. Med. Biol. Eng. Comput. 2009, 47, 719729,  DOI: 10.1007/s11517-009-0472-x
  10. 10
    Fang, B.; Yang, J.; Chen, C.; Zhang, C.; Chang, D.; Xu, H.; Gao, C. Carbon Nanotubes Loaded on Graphene Microfolds as Efficient Bifunctional Electrocatalysts for the Oxygen Reduction and Oxygen Evolution Reactions. ChemCatChem 2017, 9, 45204528,  DOI: 10.1002/cctc.201700985
  11. 11
    Li, Z.; Liu, Z.; Sun, H.; Gao, C. Superstructured Assembly of Nanocarbons: Fullerenes, Nanotubes, and Graphene. Chem. Rev. 2015, 115, 70467117,  DOI: 10.1021/acs.chemrev.5b00102
  12. 12
    Lekawa-Raus, A.; Patmore, J.; Kurzepa, L.; Bulmer, J.; Koziol, K. Electrical Properties of Carbon Nanotube Based Fibers and Their Future Use in Electrical Wiring. Adv. Funct. Mater. 2014, 24, 36613682,  DOI: 10.1002/adfm.201303716
  13. 13
    Sahebian, S.; Zebarjad, S. M.; vahdati Khaki, J.; Lazzeri, A. A Study on the Dependence of Structure of Multi-Walled Carbon Nanotubes on Acid Treatment. J. Nanostruct. Chem. 2015, 5, 287293,  DOI: 10.1007/s40097-015-0160-3
  14. 14
    Aliofkhazraei, M.; Ali, N.; Milne, W. I.; Ozkan, C. S.; Mitura, S.; Gervasoni, J. L. Graphene Science Handbook: Electrical and Optical Properties; CRC Press, 2016; p 715.
  15. 15
    Chen, J. H.; Jang, C.; Xiao, S.; Ishigami, M.; Fuhrer, M. S. Intrinsic and Extrinsic Performance Limits of Graphene Devices on SiO2. Nat. Nanotechnol. 2008, 3, 206209,  DOI: 10.1038/nnano.2008.58
  16. 16
    Gorityala, B. K.; Ma, J.; Wang, X.; Chen, P.; Liu, X. W. Carbohydrate Functionalized Carbon Nanotubes and Their Applications. Chem. Soc. Rev. 2010, 39, 29252934,  DOI: 10.1039/b919525b
  17. 17
    Bhunia, S. K.; Saha, A.; Maity, A. R.; Ray, S. C.; Jana, N. R. Carbon Nanoparticle-Based Fluorescent Bioimaging Probes. Sci. Rep. 2013, 3, 1473,  DOI: 10.1038/srep01473
  18. 18
    Fabbro, A.; Scaini, D.; León, V.; Vázquez, E.; Cellot, G.; Privitera, G.; Lombardi, L.; Torrisi, F.; Tomarchio, F.; Bonaccorso, F.; Bosi, S.; Ferrari, A. C.; Ballerini, L.; Prato, M. Graphene-Based Interfaces Do Not Alter Target Nerve Cells. ACS Nano 2016, 10, 615623,  DOI: 10.1021/acsnano.5b05647
  19. 19
    Call, T. P.; Carey, T.; Bombelli, P.; Lea-Smith, D. J.; Hooper, P.; Howe, C. J.; Torrisi, F. Platinum-Free, Graphene Based Anodes and Air Cathodes for Single Chamber Microbial Fuel Cells. J. Mater. Chem. A 2017, 5, 2387223886,  DOI: 10.1039/C7TA06895F
  20. 20
    Li, N.; Zhang, Q.; Gao, S.; Song, Q.; Huang, R.; Wang, L.; Liu, L.; Dai, J.; Tang, M.; Cheng, G. Three-Dimensional Graphene Foam as a Biocompatible and Conductive Scaffold for Neural Stem Cells. Sci. Rep. 2013, 3, 1604,  DOI: 10.1038/srep01604
  21. 21
    Mooney, E.; Dockery, P.; Greiser, U.; Murphy, M.; Barron, V. Carbon Nanotubes and Mesenchymal Stem Cells: Biocompatibility, Proliferation and Differentiation. Nano Lett. 2008, 8, 21372143,  DOI: 10.1021/nl073300o
  22. 22
    Yang, S. T.; Luo, J.; Zhou, Q.; Wang, H. Pharmacokinetics, Metabolism and Toxicity of Carbon Nanotubes for Bio-Medical Purposes. Theranostics 2012, 2, 271282,  DOI: 10.7150/thno.3618
  23. 23
    Jia, G.; Wang, H.; Yan, L.; Wang, X.; Pei, R.; Yan, T.; Zhao, Y.; Guo, X. Cytotoxicity of Carbon Nanomaterials: Single-Wall Nanotube, Multi-Wall Nanotube, and Fullerene. Environ. Sci. Technol. 2005, 39, 13781383,  DOI: 10.1021/es048729l
  24. 24
    Correa-Duarte, M. A.; Wagner, N.; Rojas-Chapana, J.; Morsczeck, C.; Thie, M.; Giersig, M. Fabrication and Biocompatibility of Carbon Nanotube-Based 3D Networks as Scaffolds for Cell Seeding and Growth. Nano Lett. 2004, 4, 22332236,  DOI: 10.1021/nl048574f
  25. 25
    Edwards, S. L.; Church, J. S.; Werkmeister, J. A.; Ramshaw, J. A. M. Tubular Micro-Scale Multiwalled Carbon Nanotube-Based Scaffolds for Tissue Engineering. Biomaterials 2009, 30, 17251731,  DOI: 10.1016/j.biomaterials.2008.12.031
  26. 26
    Liu, Y.; Zhao, Y.; Sun, B.; Chen, C. Understanding the Toxicity of Carbon Nanotubes. Acc. Chem. Res. 2013, 46, 702713,  DOI: 10.1021/ar300028m
  27. 27
    Cohen-Karni, T.; Qing, Q.; Li, Q.; Fang, Y.; Lieber, C. M. Graphene and Nanowire Transistors for Cellular Interfaces and Electrical Recording. Nano Lett. 2010, 10, 10981102,  DOI: 10.1021/nl1002608
  28. 28
    Akhavan, O.; Ghaderi, E.; Shahsavar, M. Graphene Nanogrids for Selective and Fast Osteogenic Differentiation of Human Mesenchymal Stem Cells. Carbon 2013, 59, 200211,  DOI: 10.1016/j.carbon.2013.03.010
  29. 29
    Shin, S. R.; Li, Y. C.; Jang, H. L.; Khoshakhlagh, P.; Akbari, M.; Nasajpour, A.; Zhang, Y. S.; Tamayol, A.; Khademhosseini, A. Graphene-Based Materials for Tissue Engineering. Adv. Drug Delivery Rev. 2016, 105, 255274,  DOI: 10.1016/j.addr.2016.03.007
  30. 30
    Smith, S. C.; Ahmed, F.; Gutierrez, K. M.; Frigi Rodrigues, D. A Comparative Study of Lysozyme Adsorption with Graphene, Graphene Oxide, and Single-Walled Carbon Nanotubes: Potential Environmental Applications. Chem. Eng. J. 2014, 240, 147154,  DOI: 10.1016/j.cej.2013.11.030
  31. 31
    Mecklenburg, M.; Schuchardt, A.; Mishra, Y. K.; Kaps, S.; Adelung, R.; Lotnyk, A.; Kienle, L.; Schulte, K. Aerographite: Ultra Lightweight, Flexible Nanowall, Carbon Microtube Material with Outstanding Mechanical Performance. Adv. Mater. 2012, 24, 34863490,  DOI: 10.1002/adma.201200491
  32. 32
    Mishra, Y. K.; Kaps, S.; Schuchardt, A.; Paulowicz, I.; Jin, X.; Gedamu, D.; Freitag, S.; Claus, M.; Wille, S.; Kovalev, A.; Gorb, S. N.; Adelung, R. Fabrication of Macroscopically Flexible and Highly Porous 3D Semiconductor Networks from Interpenetrating Nanostructures by a Simple Flame Transport Approach. Part. Part. Syst. Charact. 2013, 30, 775783,  DOI: 10.1002/ppsc.201300197
  33. 33
    Lamprecht, C.; Taale, M.; Paulowicz, I.; Westerhaus, H.; Grabosch, C.; Schuchardt, A.; Mecklenburg, M.; Böttner, M.; Lucius, R.; Schulte, K.; Adelung, R.; Selhuber-Unkel, C. A Tunable Scaffold of Microtubular Graphite for 3D Cell Growth. ACS Appl. Mater. Interfaces 2016, 8, 1498014985,  DOI: 10.1021/acsami.6b00778
  34. 34
    Mouw, J. K.; Ou, G.; Weaver, V. M. Extracellular Matrix Assembly: A Multiscale Deconstruction. Nat. Rev. Mol. Cell Biol. 2014, 15, 771785,  DOI: 10.1038/nrm3902
  35. 35
    Schütt, F.; Signetti, S.; Krüger, H.; Röder, S.; Smazna, D.; Kaps, S.; Gorb, S. N.; Mishra, Y. K.; Pugno, N. M.; Adelung, R. Hierarchical Self-Entangled Carbon Nanotube Tube Networks. Nat. Commun. 2017, 8, 1215  DOI: 10.1038/s41467-017-01324-7
  36. 36
    Feng, X.; Feng, L.; Jin, M.; Zhai, J.; Jiang, L.; Zhu, D. Reversible Super-Hydrophobicity to Super-Hydrophilicity Transition of Aligned ZnO Nanorod Films. J. Am. Chem. Soc. 2004, 126, 6263,  DOI: 10.1021/ja038636o
  37. 37
    Chen, H.-K. Kinetic Study on the Carbothermic Reduction of Zinc Oxide. Scand. J. Metall. 2001, 30, 292296,  DOI: 10.1034/j.1600-0692.2001.300503.x
  38. 38
    Zhao, J.; Xing, B.; Yang, H.; Pan, Q.; Li, Z.; Liu, Z. Growth of Carbon Nanotubes on Graphene by Chemical Vapor Deposition. New Carbon Mater. 2016, 31, 3136,  DOI: 10.1016/S1872-5805(16)60002-1
  39. 39
    Ferrari, A. C.; Robertson, J. Resonant Raman Spectroscopy of Disordered, Amorphous, and Diamondlike Carbon. Phys. Rev. B: Condens. Matter Mater. Phys. 2001, 64, 075414,  DOI: 10.1103/PhysRevB.64.075414
  40. 40
    Carey, T.; Cacovich, S.; Divitini, G.; Ren, J.; Mansouri, A.; Kim, J. M.; Wang, C.; Ducati, C.; Sordan, R.; Torrisi, F. Fully Inkjet-Printed Two-Dimensional Material Field-Effect Heterojunctions for Wearable and Textile Electronics. Nat. Commun. 2017, 8, 1202,  DOI: 10.1038/s41467-017-01210-2
  41. 41
    Ferrari, A. C.; Meyer, J. C.; Scardaci, V.; Casiraghi, C.; Lazzeri, M.; Mauri, F.; Piscanec, S.; Jiang, D.; Novoselov, K. S.; Roth, S. Raman Spectrum of Graphene and Graphene Layers. Phys. Rev. Lett. 2006, 97, 187401,  DOI: 10.1103/PhysRevLett.97.187401
  42. 42
    Ferrari, A.; Robertson, J. Interpretation of Raman Spectra of Disordered and Amorphous Carbon. Phys. Rev. B: Condens. Matter Mater. Phys. 2000, 61, 1409514107,  DOI: 10.1103/PhysRevB.61.14095
  43. 43
    Cuscó, R.; Alarcón-Lladó, E.; Ibáñez, J.; Artús, L.; Jiménez, J.; Wang, B.; Callahan, M. Temperature Dependence of Raman Scattering in ZnO. Phys. Rev. B: Condens. Matter Mater. Phys. 2007, 75, 165202,  DOI: 10.1103/PhysRevB.75.165202
  44. 44
    Fan, H.; Wang, L.; Zhao, K.; Li, N.; Shi, Z.; Ge, Z.; Jin, Z. Fabrication, Mechanical Properties, and Biocompatibility of Graphene-Reinforced Chitosan Composites. Biomacromolecules 2010, 11, 23452351,  DOI: 10.1021/bm100470q
  45. 45
    Ahadian, S.; Ramón-Azcón, J.; Estili, M.; Liang, X.; Ostrovidov, S.; Shiku, H.; Ramalingam, M.; Nakajima, K.; Sakka, Y.; Bae, H. H.; Matsue, T.; Khademhosseini, A. Hybrid Hydrogels Containing Vertically Aligned Carbon Nanotubes with Anisotropic Electrical Conductivity for Muscle Myofiber Fabrication. Sci. Rep. 2015, 4, 4271,  DOI: 10.1038/srep04271
  46. 46
    Shi, X.; Sitharaman, B.; Pham, Q. P.; Liang, F.; Wu, K.; Edward Billups, W.; Wilson, L. J.; Mikos, A. G. Fabrication of Porous Ultra-Short Single-Walled Carbon Nanotube Nanocomposite Scaffolds for Bone Tissue Engineering. Biomaterials 2007, 28, 40784090,  DOI: 10.1016/j.biomaterials.2007.05.033
  47. 47
    Chen, Z.; Ren, W.; Gao, L.; Liu, B.; Pei, S.; Cheng, H. M. Three-Dimensional Flexible and Conductive Interconnected Graphene Networks Grown by Chemical Vapour Deposition. Nat. Mater. 2011, 10, 424428,  DOI: 10.1038/nmat3001
  48. 48
    Zhao, G.; Zhang, X.; Lu, T. J.; Xu, F. Recent Advances in Electrospun Nanofibrous Scaffolds for Cardiac Tissue Engineering. Adv. Funct. Mater. 2015, 25, 57265738,  DOI: 10.1002/adfm.201502142
  49. 49
    Zuo, G.; Kang, S. G.; Xiu, P.; Zhao, Y.; Zhou, R. Interactions between Proteins and Carbon-Based Nanoparticles: Exploring the Origin of Nanotoxicity at the Molecular Level. Small 2013, 9, 15461556,  DOI: 10.1002/smll.201201381
  50. 50
    Suh, C. W.; Kim, M. Y.; Choo, J. B.; Kim, J. K.; Kim, H. K.; Lee, E. K. Analysis of Protein Adsorption Characteristics to Nano-Pore Silica Particles by Using Confocal Laser Scanning Microscopy. J. Biotechnol. 2004, 112, 267277,  DOI: 10.1016/j.jbiotec.2004.05.005
  51. 51
    Prager-Khoutorsky, M.; Lichtenstein, A.; Krishnan, R.; Rajendran, K.; Mayo, A.; Kam, Z.; Geiger, B.; Bershadsky, A. D. Fibroblast Polarization Is a Matrix-Rigidity-Dependent Process Controlled by Focal Adhesion Mechanosensing. Nat. Cell Biol. 2011, 13, 14571465,  DOI: 10.1038/ncb2370
  52. 52
    El-Mohri, H.; Wu, Y.; Mohanty, S.; Ghosh, G. Impact of Matrix Stiffness on Fibroblast Function. Mater. Sci. Eng., C 2017, 74, 146151,  DOI: 10.1016/j.msec.2017.02.001
  53. 53
    Sun, Y.; Chen, C. S.; Fu, J. Forcing Stem Cells to Behave: A Biophysical Perspective of the Cellular Microenvironment. Annu. Rev. Biophys. 2012, 41, 519542,  DOI: 10.1146/annurev-biophys-042910-155306
  54. 54
    Moreno-Arotzena, O.; Borau, C.; Movilla, N.; Vicente-Manzanares, M.; García-Aznar, J. M. Fibroblast Migration in 3D Is Controlled by Haptotaxis in a Non-Muscle Myosin II-Dependent Manner. Ann. Biomed. Eng. 2015, 43, 30253039,  DOI: 10.1007/s10439-015-1343-2
  55. 55
    Turner, C. E. Paxillin and Focal Adhesion Signalling. Nat. Cell Biol. 2000, 2, E231E236,  DOI: 10.1038/35046659
  56. 56
    Chiu, C.-L.; Aguilar, J. S.; Tsai, C. Y.; Wu, G.; Gratton, E.; Digman, M. M. A.; Kuo, J.; Han, X.; Hsiao, C.; Iii, J. Y. Nanoimaging of Focal Adhesion Dynamics in 3D. PLoS One 2014, 9, e99896,  DOI: 10.1371/journal.pone.0099896
  57. 57
    Ryoo, S. R.; Kim, Y. K.; Kim, M. H.; Min, D. H. Behaviors of NIH-3T3 Fibroblasts on Graphene/Carbon Nanotubes: Proliferation, Focal Adhesion, and Gene Transfection Studies. ACS Nano 2010, 4, 65876598,  DOI: 10.1021/nn1018279
  58. 58
    Cooper, G. M. The Cell: A Molecular Approach; ASM Press, 2000; Vol. 10.
  59. 59
    Serrano, M. C.; Patiño, J.; García-Rama, C.; Ferrer, M. L.; Fierro, J. L. G.; Tamayo, A.; Collazos-Castro, J. E.; Del Monte, F.; Gutiérrez, M. C. 3D Free-Standing Porous Scaffolds Made of Graphene Oxide as Substrates for Neural Cell Growth. J. Mater. Chem. B 2014, 2, 56985706,  DOI: 10.1039/C4TB00652F
  60. 60
    Avery, N. C.; Bailey, A. J.; Fratzl, P. Collagen: Structure and Mechanics; Springer, 2008.
  61. 61
    LeBaron, R. G.; Athanasiou, K. A. Extracellular Matrix Cell Adhesion Peptides: Functional Applications in Orthopedic Materials. Tissue Eng. 2000, 6, 85103,  DOI: 10.1089/107632700320720
  62. 62
    Marx, J.; Lewke, M. R. D.; Smazna, D.; Mishra, Y. K.; Adelung, R.; Schulte, K.; Fiedler, B. Processing, Growth Mechanism and Thermodynamic Calculations of Carbon Foam with a Hollow Tetrapodal Morphology – Aerographite. Appl. Surf. Sci. 2019, 470, 535542,  DOI: 10.1016/j.apsusc.2018.11.016
  63. 63
    Wang, F.; Torrisi, F.; Jiang, Z.; Popa, D.; Hasan, T.; Sun, Z.; Cho, W.; Ferrari, A. C. Graphene Passively Q-Switched Two-Micron Fiber Lasers. In Conference on Lasers and Electro-Optics (CLEO); OSA Technical Digest (Optical Society of America): San Jose, CA, 2012; p JW2A 72.
  64. 64
    Purdie, D. G.; Popa, D.; Wittwer, V. J.; Jiang, Z.; Bonacchini, G.; Torrisi, F.; Milana, S.; Lidorikis, E.; Ferrari, A. C. Few-Cycle Pulses from a Graphene Mode-Locked All-Fiber Laser. Appl. Phys. Lett. 2015, 106, 253101,  DOI: 10.1063/1.4922397
  65. 65
    Bianchi, V.; Carey, T.; Viti, L.; Li, L.; Linfield, E. H.; Davies, A. G.; Tredicucci, A.; Yoon, D.; Karagiannidis, P. G.; Lombardi, L.; Tomarchio, F.; Ferrari, A. C.; Torrisi, F.; Vitiello, M. S. Terahertz Saturable Absorbers from Liquid Phase Exfoliation of Graphite. Nat. Commun. 2017, 8, 15763,  DOI: 10.1038/ncomms15763
  66. 66
    Lotya, M.; Hernandez, Y.; King, P. J.; Smith, R. J.; Nicolosi, V.; Karlsson, L. S.; Blighe, F. M.; De, S.; Wang, Z.; McGovern, I. T.; Duesberg, G. S.; Coleman, J. N. Liquid Phase Production of Graphene by Exfoliation of Graphite in Surfactant/Water Solutions. J. Am. Chem. Soc. 2009, 131, 36113620,  DOI: 10.1021/ja807449u
  67. 67
    Carey, T.; Jones, C.; Le Moal, F.; Deganello, D.; Torrisi, F. Spray Coating Thin Films on Three-Dimensional Surfaces for a Semi-Transparent Capacitive Touch Device. ACS Appl. Mater. Interfaces 2018, 10, 1994819956,  DOI: 10.1021/acsami.8b02784
  68. 68
    Kravets, V. G.; Grigorenko, A. N.; Nair, R. R.; Blake, P.; Anissimova, S.; Novoselov, K. S.; Geim, A. K. Spectroscopic Ellipsometry of Graphene and an Exciton-Shifted van Hove Peak in Absorption. Phys. Rev. B: Condens. Matter Mater. Phys. 2010, 81, 155413,  DOI: 10.1103/PhysRevB.81.155413
  69. 69
    Selhuber-Unkel, C.; Erdmann, T.; López-García, M.; Kessler, H.; Schwarz, U. S.; Spatz, J. P. Cell Adhesion Strength Is Controlled by Intermolecular Spacing of Adhesion Receptors. Biophys. J. 2010, 98, 543551,  DOI: 10.1016/j.bpj.2009.11.001
  70. 70
    Wörle-Knirsch, J. M.; Pulskamp, K.; Krug, H. F. Oops They Did It Again! Carbon Nanotubes Hoax Scientists in Viability Assays. Nano Lett. 2006, 6, 12611268,  DOI: 10.1021/nl060177c

Cited By

ARTICLE SECTIONS
Jump To

This article is cited by 24 publications.

  1. Anna Gapeeva, Haoyi Qiu, Ala Cojocaru, Christine Arndt, Tehseen Riaz, Fabian Schütt, Christine Selhuber-Unkel, Yogendra Kumar Mishra, Aysegül Tura, Svenja Sonntag, Stefanie Gniesmer, Salvatore Grisanti, Sören Kaps, Rainer Adelung. Tetrapodal ZnO-Based Composite Stents for Minimally Invasive Glaucoma Surgery. ACS Biomaterials Science & Engineering 2023, 9 (3) , 1352-1361. https://doi.org/10.1021/acsbiomaterials.2c01203
  2. Christine Arndt, Margarethe Hauck, Irene Wacker, Berit Zeller-Plumhoff, Florian Rasch, Mohammadreza Taale, Ali Shaygan Nia, Xinliang Feng, Rainer Adelung, Rasmus R. Schröder, Fabian Schütt, Christine Selhuber-Unkel. Microengineered Hollow Graphene Tube Systems Generate Conductive Hydrogels with Extremely Low Filler Concentration. Nano Letters 2021, 21 (8) , 3690-3697. https://doi.org/10.1021/acs.nanolett.0c04375
  3. Kunal Mondal, Tanmoy Maitra, Alok Kumar Srivastava, Gorakh Pawar, Michael D. McMurtrey, Ashutosh Sharma. 110th Anniversary: Particle Size Effect on Enhanced Graphitization and Electrical Conductivity of Suspended Gold/Carbon Composite Nanofibers. Industrial & Engineering Chemistry Research 2020, 59 (5) , 1944-1952. https://doi.org/10.1021/acs.iecr.9b06592
  4. Huaibin Qing, Yuan Ji, Wenfang Li, Guoxu Zhao, Qingzhen Yang, Xiaohui Zhang, Zhengtang Luo, Tian Jian Lu, Guorui Jin, Feng Xu. Microfluidic Printing of Three-Dimensional Graphene Electroactive Microfibrous Scaffolds. ACS Applied Materials & Interfaces 2020, 12 (2) , 2049-2058. https://doi.org/10.1021/acsami.9b17948
  5. Florian Rasch, Fabian Schütt, Lena M. Saure, Sören Kaps, Julian Strobel, Oleksandr Polonskyi, Ali Shaygan Nia, Martin R. Lohe, Yogendra K. Mishra, Franz Faupel, Lorenz Kienle, Xinliang Feng, Rainer Adelung. Wet-Chemical Assembly of 2D Nanomaterials into Lightweight, Microtube-Shaped, and Macroscopic 3D Networks. ACS Applied Materials & Interfaces 2019, 11 (47) , 44652-44663. https://doi.org/10.1021/acsami.9b16565
  6. Mohammadreza Taale, Diana Krüger, Emmanuel Ossei-Wusu, Fabian Schütt, Muhammad Atiq Ur Rehman, Yogendra Kumar Mishra, Janik Marx, Norbert Stock, Bodo Fiedler, Aldo R. Boccaccini, Regine Willumeit-Römer, Rainer Adelung, Christine Selhuber-Unkel. Systematically Designed Periodic Electrophoretic Deposition for Decorating 3D Carbon-Based Scaffolds with Bioactive Nanoparticles. ACS Biomaterials Science & Engineering 2019, 5 (9) , 4393-4404. https://doi.org/10.1021/acsbiomaterials.9b00102
  7. Mohammadreza Taale, Barbara Schamberger, Miguel A. Monclus, Christian Dolle, Fereydoon Taheri, Dario Mager, Yolita M. Eggeler, Jan G. Korvink, Jon M. Molina‐Aldareguia, Christine Selhuber‐Unkel, Andrés Díaz Lantada, Monsur Islam. Microarchitected Compliant Scaffolds of Pyrolytic Carbon for 3D Muscle Cell Growth. Advanced Healthcare Materials 2024, 13 (9) https://doi.org/10.1002/adhm.202303485
  8. Monsur Islam, Christine Selhuber-Unkel, Jan G. Korvink, Andrés Díaz Lantada. Engineered living carbon materials. Matter 2023, 6 (5) , 1382-1403. https://doi.org/10.1016/j.matt.2023.03.018
  9. Yuexuan Li, Hiromu Hamasaki, Kaori Hirahara. Effect of annealing treatment on the mechanical properties of spiked-shell aerographite particles. Carbon 2023, 203 , 523-533. https://doi.org/10.1016/j.carbon.2022.11.079
  10. Tian Carey, Abdelnour Alhourani, Ruiyuan Tian, Shayan Seyedin, Adrees Arbab, Jack Maughan, Lidija Šiller, Dominik Horvath, Adam Kelly, Harneet Kaur, Eoin Caffrey, Jong M. Kim, Hanne R. Hagland, Jonathan N. Coleman. Cyclic production of biocompatible few-layer graphene ink with in-line shear-mixing for inkjet-printed electrodes and Li-ion energy storage. npj 2D Materials and Applications 2022, 6 (1) https://doi.org/10.1038/s41699-021-00279-0
  11. Kamil G. Gareev, Denis S. Grouzdev, Veronika V. Koziaeva, Nikita O. Sitkov, Huile Gao, Tatiana M. Zimina, Maxim Shevtsov. Biomimetic Nanomaterials: Diversity, Technology, and Biomedical Applications. Nanomaterials 2022, 12 (14) , 2485. https://doi.org/10.3390/nano12142485
  12. Monsur Islam, Andrés Díaz Lantada, Dario Mager, Jan G. Korvink. Carbon‐Based Materials for Articular Tissue Engineering: From Innovative Scaffolding Materials toward Engineered Living Carbon. Advanced Healthcare Materials 2022, 11 (1) https://doi.org/10.1002/adhm.202101834
  13. Jiaqi Wang, Shu Zhang, Han Cao, Junzhou Ma, Lintianyang Huang, Shujun Yu, Xiaoying Ma, Gang Song, Muqing Qiu, Xiangxue Wang. Water purification and environmental remediation applications of carbonaceous nanofiber-based materials. Journal of Cleaner Production 2022, 331 , 130023. https://doi.org/10.1016/j.jclepro.2021.130023
  14. Yuexuan Li, Hiromu Hamasaki, Kaori Hirahara. Effect of Annealing Treatment on the Mechanical Properties of Spiked-Shell Aerographite Particles. SSRN Electronic Journal 2022, 24 https://doi.org/10.2139/ssrn.4199853
  15. Birte Hindenlang, Anna Gapeeva, Martina J. Baum, Sören Kaps, Lena M. Saure, Florian Rasch, Jörg Hammel, Julian Moosmann, Malte Storm, Rainer Adelung, Fabian Schütt, Berit Zeller-Plumhoff. Evaporation kinetics in highly porous tetrapodal zinc oxide networks studied using in situ SRµCT. Scientific Reports 2021, 11 (1) https://doi.org/10.1038/s41598-021-99624-y
  16. D. Sharma, P. Shandilya, N.K. Saini, P. Singh, V.K. Thakur, R.V. Saini, D. Mittal, G. Chandan, V. Saini, A.K. Saini. Insights into the synthesis and mechanism of green synthesized antimicrobial nanoparticles, answer to the multidrug resistance. Materials Today Chemistry 2021, 19 , 100391. https://doi.org/10.1016/j.mtchem.2020.100391
  17. Lakshmi Padmavathi Pydipati, Bharathi Depuru, Suvarna Latha Anchapakula, Hemavathi Buddareddy. Effectiveness of Plant-Based Biomaterials in the Field of Biomedical Engineering. 2021, 735-743. https://doi.org/10.1007/978-981-16-1941-0_75
  18. Christina Schmitt, Florian Rasch, François Cossais, Janka Held-Feindt, Ralph Lucius, Adrian Romani Vázquez, Ali Shaygan Nia, Martin R Lohe, Xinliang Feng, Yogendra K Mishra, Rainer Adelung, Fabian Schütt, Kirsten Hattermann. Glial cell responses on tetrapod-shaped graphene oxide and reduced graphene oxide 3D scaffolds in brain in vitro and ex vivo models of indirect contact. Biomedical Materials 2021, 16 (1) , 015008. https://doi.org/10.1088/1748-605X/aba796
  19. O. Platnieks, S. Gaidukovs, N. Neibolts, A. Barkane, G. Gaidukova, V.K. Thakur. Poly(butylene succinate) and graphene nanoplatelet–based sustainable functional nanocomposite materials: structure-properties relationship. Materials Today Chemistry 2020, 18 , 100351. https://doi.org/10.1016/j.mtchem.2020.100351
  20. Francesco De Nicola, Ilenia Viola, Lorenzo Donato Tenuzzo, Florian Rasch, Martin R. Lohe, Ali Shaygan Nia, Fabian Schütt, Xinliang Feng, Rainer Adelung, Stefano Lupi. Wetting Properties of Graphene Aerogels. Scientific Reports 2020, 10 (1) https://doi.org/10.1038/s41598-020-58860-4
  21. Raj Kumar, Kunal Mondal, Pritam Kumar Panda, Ajeet Kaushik, Reza Abolhassani, Rajeev Ahuja, Horst-Günter Rubahn, Yogendra Kumar Mishra. Core–shell nanostructures: perspectives towards drug delivery applications. Journal of Materials Chemistry B 2020, 8 (39) , 8992-9027. https://doi.org/10.1039/D0TB01559H
  22. Ying Luo. Three-dimensional scaffolds. 2020, 343-360. https://doi.org/10.1016/B978-0-12-818422-6.00020-4
  23. Lang Ma, Mi Zhou, Chao He, Shuang Li, Xin Fan, Chuanxiong Nie, Hongrong Luo, Li Qiu, Chong Cheng. Graphene-based advanced nanoplatforms and biocomposites from environmentally friendly and biomimetic approaches. Green Chemistry 2019, 21 (18) , 4887-4918. https://doi.org/10.1039/C9GC02266J
  24. Reza Mohammadinejad, Hajar Maleki, Eneko Larrañeta, André R. Fajardo, Amirala Bakhshian Nik, Amin Shavandi, Amir Sheikhi, Mansour Ghorbanpour, Mehdi Farokhi, Praveen Govindh, Etienne Cabane, Susan Azizi, Amir Reza Aref, Masoud Mozafari, Mehdi Mehrali, Sabu Thomas, João F. Mano, Yogendra Kumar Mishra, Vijay Kumar Thakur. Status and future scope of plant-based green hydrogels in biomedical engineering. Applied Materials Today 2019, 16 , 213-246. https://doi.org/10.1016/j.apmt.2019.04.010
  • Abstract

    Figure 1

    Figure 1. Schematic illustration of different 3D carbon tube structures. The highly porous ZnO template can be either infiltrated with a nanoparticle dispersion (e.g., graphene, CNT) leading to a homogeneous coating around the tetrapodal particles or converted to a graphitic structure using a chemical vapor deposition (CVD) process (aerographite). The combination of both processes leads to a modular design strategy, especially in terms of conductivity, mechanical stiffness, and surface topography.

    Figure 2

    Figure 2. SEM images from low to high magnification of the different 3D carbon structures. (a−c) Carbon nanotube tubes (CNTTs), (d−f) aerographite with incorporated CNTs (AG−CNTT) (the red arrows show the grown CNTs during the CVD process), (g−i) aerographite (AG), (j−l) aerographite with incorporated graphene (AG−G) (the yellow arrows point to nanopores on the surface of aerographite), and (m−o) carbon nanotube tubes incorporated into a thick carbon layer (CNTT−g).

    Figure 3

    Figure 3. Raman spectroscopy of aerographite, ZnO, CNTT, AG–CNTT, CNTT–g, AG–G, graphene, and CNT inks.

    Figure 4

    Figure 4. Young’s modulus (measured under compression) and electrical conductivity of the 3D CNTT, AG–CNTT, CNTT–g, and AG–G scaffolds. All structures containing CNTs have a higher Young’s modulus and conductivity compared to those without CNTs. The values for pure AG and CNTTs were taken from the corresponding publications, (31,35) whereas the other values were measured using a self-built electromechanical testing setup.

    Figure 5

    Figure 5. Protein adsorption on CNTT, AG–CNTT, CNTT–g, and AG–G during 0–48, 49–96, and 97–144 h of incubation with albumin solution (1 mg/mL). (a) Absolute protein adsorption amount per scaffold volume. (b) Absolute protein adsorption amount per scaffold weight.

    Figure 6

    Figure 6. Percentage of viable cells (rat embryonic fibroblasts wild type, REF52wt) relative to the negative control, as determined in an MTT assay (four independent experiments, five technical repeats in each of them). The error bars denote standard deviation.

    Figure 7

    Figure 7. Results of WST-1 metabolic tests of cell proliferation rate (REF52wt). Mean values were determined from four independent experiments, each including five technical repeats. The error bars denote standard deviation, the raw data were normalized to the control, and a correction factor was applied to account for unspecific adsorption (see Materials and Methods).

    Figure 8

    Figure 8. SEM images of REF52wt cells after 7 days of culturing within (a–c) AG–CNTT, (d, e) CNTT–g, (g–i) AG–G, and (j–l) CNTT scaffolds. Left: Medium-sized overview images illustrating the growth of cells between the fibers in different directions and planes; middle: zoomed-in images showing well-stretched cells along the fibers and their elongation; and right: close-up images on the adhesion sites, proving the presence of strong contacts between the materials and the cell membrane (yellow arrows).

    Figure 9

    Figure 9. High-magnification fluorescence images of REF52 YFP-paxillin cells on 3D scaffolds: (a) CNTT, (b) AG–CNTT, (c) CNTT–g, and (d) AG–G. YFP-paxillin is mainly distributed in small clusters (YFP; yellow), the nucleus was stained with Hoechst (4′,6-diamidino-2-phenylindole, DAPI; blue), and actin fibers were visualized by phalloidin (red). Fluorescence imaging took place in optical sections approximately between 50 and 100 μm from the surface of the material. Adhesion sites are detectable (a, b) as tiny yellow spots mainly around the tube-shaped filaments of CNTT and AG–CNTT structures. Due to the low intensity and small size of YFP-paxillin (c, d), they are not clearly observed, which is also obscured by red (actin fibers) and blue (nucleus) channels in multichannel images. The green dashed arrows illustrate the protrusion direction of the cells.

  • References

    ARTICLE SECTIONS
    Jump To

    This article references 70 other publications.

    1. 1
      O’Brien, F. J. Biomaterials & Scaffolds for Tissue Engineering. Mater. Today 2011, 14, 8895,  DOI: 10.1016/S1369-7021(11)70058-X
    2. 2
      Yang, S.; Leong, K.-F.; Du, Z.; Chua, C.-K. The Design of Scaffolds for Use in Tissue Engineering. Part I. Traditional Factors. Tissue Eng. 2001, 7, 679689,  DOI: 10.1089/107632701753337645
    3. 3
      Shoulders, M. D.; Raines, R. T. Collagen Structure and Stability. Annu. Rev. Biochem. 2009, 78, 929958,  DOI: 10.1146/annurev.biochem.77.032207.120833
    4. 4
      Friedl, P.; Wolf, K. Plasticity of Cell Migration: A Multiscale Tuning Model. J. Cell Biol. 2010, 188, 1119,  DOI: 10.1083/jcb.200909003
    5. 5
      Langer, R.; Vacanti, J. P. Tissue Engineering. Science 1993, 260, 920926,  DOI: 10.1126/science.8493529
    6. 6
      Place, E. S.; Evans, N. D.; Stevens, M. M. Complexity in Biomaterials for Tissue Engineering. Nat. Mater. 2009, 8, 457470,  DOI: 10.1038/nmat2441
    7. 7
      Loh, Q. L.; Choong, C. Three-Dimensional Scaffolds for Tissue Engineering Applications: Role of Porosity and Pore Size. Tissue Eng., Part B 2013, 19, 485502,  DOI: 10.1089/ten.teb.2012.0437
    8. 8
      Ganji, Y.; Li, Q.; Quabius, E. S.; Böttner, M.; Selhuber-Unkel, C.; Kasra, M. Cardiomyocyte Behavior on Biodegradable Polyurethane/Gold Nanocomposite Scaffolds under Electrical Stimulation. Mater. Sci. Eng., C 2016, 59, 1018,  DOI: 10.1016/j.msec.2015.09.074
    9. 9
      Potse, M.; Dubé, B.; Vinet, A. Cardiac Anisotropy in Boundary-Element Models for the Electrocardiogram. Med. Biol. Eng. Comput. 2009, 47, 719729,  DOI: 10.1007/s11517-009-0472-x
    10. 10
      Fang, B.; Yang, J.; Chen, C.; Zhang, C.; Chang, D.; Xu, H.; Gao, C. Carbon Nanotubes Loaded on Graphene Microfolds as Efficient Bifunctional Electrocatalysts for the Oxygen Reduction and Oxygen Evolution Reactions. ChemCatChem 2017, 9, 45204528,  DOI: 10.1002/cctc.201700985
    11. 11
      Li, Z.; Liu, Z.; Sun, H.; Gao, C. Superstructured Assembly of Nanocarbons: Fullerenes, Nanotubes, and Graphene. Chem. Rev. 2015, 115, 70467117,  DOI: 10.1021/acs.chemrev.5b00102
    12. 12
      Lekawa-Raus, A.; Patmore, J.; Kurzepa, L.; Bulmer, J.; Koziol, K. Electrical Properties of Carbon Nanotube Based Fibers and Their Future Use in Electrical Wiring. Adv. Funct. Mater. 2014, 24, 36613682,  DOI: 10.1002/adfm.201303716
    13. 13
      Sahebian, S.; Zebarjad, S. M.; vahdati Khaki, J.; Lazzeri, A. A Study on the Dependence of Structure of Multi-Walled Carbon Nanotubes on Acid Treatment. J. Nanostruct. Chem. 2015, 5, 287293,  DOI: 10.1007/s40097-015-0160-3
    14. 14
      Aliofkhazraei, M.; Ali, N.; Milne, W. I.; Ozkan, C. S.; Mitura, S.; Gervasoni, J. L. Graphene Science Handbook: Electrical and Optical Properties; CRC Press, 2016; p 715.
    15. 15
      Chen, J. H.; Jang, C.; Xiao, S.; Ishigami, M.; Fuhrer, M. S. Intrinsic and Extrinsic Performance Limits of Graphene Devices on SiO2. Nat. Nanotechnol. 2008, 3, 206209,  DOI: 10.1038/nnano.2008.58
    16. 16
      Gorityala, B. K.; Ma, J.; Wang, X.; Chen, P.; Liu, X. W. Carbohydrate Functionalized Carbon Nanotubes and Their Applications. Chem. Soc. Rev. 2010, 39, 29252934,  DOI: 10.1039/b919525b
    17. 17
      Bhunia, S. K.; Saha, A.; Maity, A. R.; Ray, S. C.; Jana, N. R. Carbon Nanoparticle-Based Fluorescent Bioimaging Probes. Sci. Rep. 2013, 3, 1473,  DOI: 10.1038/srep01473
    18. 18
      Fabbro, A.; Scaini, D.; León, V.; Vázquez, E.; Cellot, G.; Privitera, G.; Lombardi, L.; Torrisi, F.; Tomarchio, F.; Bonaccorso, F.; Bosi, S.; Ferrari, A. C.; Ballerini, L.; Prato, M. Graphene-Based Interfaces Do Not Alter Target Nerve Cells. ACS Nano 2016, 10, 615623,  DOI: 10.1021/acsnano.5b05647
    19. 19
      Call, T. P.; Carey, T.; Bombelli, P.; Lea-Smith, D. J.; Hooper, P.; Howe, C. J.; Torrisi, F. Platinum-Free, Graphene Based Anodes and Air Cathodes for Single Chamber Microbial Fuel Cells. J. Mater. Chem. A 2017, 5, 2387223886,  DOI: 10.1039/C7TA06895F
    20. 20
      Li, N.; Zhang, Q.; Gao, S.; Song, Q.; Huang, R.; Wang, L.; Liu, L.; Dai, J.; Tang, M.; Cheng, G. Three-Dimensional Graphene Foam as a Biocompatible and Conductive Scaffold for Neural Stem Cells. Sci. Rep. 2013, 3, 1604,  DOI: 10.1038/srep01604
    21. 21
      Mooney, E.; Dockery, P.; Greiser, U.; Murphy, M.; Barron, V. Carbon Nanotubes and Mesenchymal Stem Cells: Biocompatibility, Proliferation and Differentiation. Nano Lett. 2008, 8, 21372143,  DOI: 10.1021/nl073300o
    22. 22
      Yang, S. T.; Luo, J.; Zhou, Q.; Wang, H. Pharmacokinetics, Metabolism and Toxicity of Carbon Nanotubes for Bio-Medical Purposes. Theranostics 2012, 2, 271282,  DOI: 10.7150/thno.3618
    23. 23
      Jia, G.; Wang, H.; Yan, L.; Wang, X.; Pei, R.; Yan, T.; Zhao, Y.; Guo, X. Cytotoxicity of Carbon Nanomaterials: Single-Wall Nanotube, Multi-Wall Nanotube, and Fullerene. Environ. Sci. Technol. 2005, 39, 13781383,  DOI: 10.1021/es048729l
    24. 24
      Correa-Duarte, M. A.; Wagner, N.; Rojas-Chapana, J.; Morsczeck, C.; Thie, M.; Giersig, M. Fabrication and Biocompatibility of Carbon Nanotube-Based 3D Networks as Scaffolds for Cell Seeding and Growth. Nano Lett. 2004, 4, 22332236,  DOI: 10.1021/nl048574f
    25. 25
      Edwards, S. L.; Church, J. S.; Werkmeister, J. A.; Ramshaw, J. A. M. Tubular Micro-Scale Multiwalled Carbon Nanotube-Based Scaffolds for Tissue Engineering. Biomaterials 2009, 30, 17251731,  DOI: 10.1016/j.biomaterials.2008.12.031
    26. 26
      Liu, Y.; Zhao, Y.; Sun, B.; Chen, C. Understanding the Toxicity of Carbon Nanotubes. Acc. Chem. Res. 2013, 46, 702713,  DOI: 10.1021/ar300028m
    27. 27
      Cohen-Karni, T.; Qing, Q.; Li, Q.; Fang, Y.; Lieber, C. M. Graphene and Nanowire Transistors for Cellular Interfaces and Electrical Recording. Nano Lett. 2010, 10, 10981102,  DOI: 10.1021/nl1002608
    28. 28
      Akhavan, O.; Ghaderi, E.; Shahsavar, M. Graphene Nanogrids for Selective and Fast Osteogenic Differentiation of Human Mesenchymal Stem Cells. Carbon 2013, 59, 200211,  DOI: 10.1016/j.carbon.2013.03.010
    29. 29
      Shin, S. R.; Li, Y. C.; Jang, H. L.; Khoshakhlagh, P.; Akbari, M.; Nasajpour, A.; Zhang, Y. S.; Tamayol, A.; Khademhosseini, A. Graphene-Based Materials for Tissue Engineering. Adv. Drug Delivery Rev. 2016, 105, 255274,  DOI: 10.1016/j.addr.2016.03.007
    30. 30
      Smith, S. C.; Ahmed, F.; Gutierrez, K. M.; Frigi Rodrigues, D. A Comparative Study of Lysozyme Adsorption with Graphene, Graphene Oxide, and Single-Walled Carbon Nanotubes: Potential Environmental Applications. Chem. Eng. J. 2014, 240, 147154,  DOI: 10.1016/j.cej.2013.11.030
    31. 31
      Mecklenburg, M.; Schuchardt, A.; Mishra, Y. K.; Kaps, S.; Adelung, R.; Lotnyk, A.; Kienle, L.; Schulte, K. Aerographite: Ultra Lightweight, Flexible Nanowall, Carbon Microtube Material with Outstanding Mechanical Performance. Adv. Mater. 2012, 24, 34863490,  DOI: 10.1002/adma.201200491
    32. 32
      Mishra, Y. K.; Kaps, S.; Schuchardt, A.; Paulowicz, I.; Jin, X.; Gedamu, D.; Freitag, S.; Claus, M.; Wille, S.; Kovalev, A.; Gorb, S. N.; Adelung, R. Fabrication of Macroscopically Flexible and Highly Porous 3D Semiconductor Networks from Interpenetrating Nanostructures by a Simple Flame Transport Approach. Part. Part. Syst. Charact. 2013, 30, 775783,  DOI: 10.1002/ppsc.201300197
    33. 33
      Lamprecht, C.; Taale, M.; Paulowicz, I.; Westerhaus, H.; Grabosch, C.; Schuchardt, A.; Mecklenburg, M.; Böttner, M.; Lucius, R.; Schulte, K.; Adelung, R.; Selhuber-Unkel, C. A Tunable Scaffold of Microtubular Graphite for 3D Cell Growth. ACS Appl. Mater. Interfaces 2016, 8, 1498014985,  DOI: 10.1021/acsami.6b00778
    34. 34
      Mouw, J. K.; Ou, G.; Weaver, V. M. Extracellular Matrix Assembly: A Multiscale Deconstruction. Nat. Rev. Mol. Cell Biol. 2014, 15, 771785,  DOI: 10.1038/nrm3902
    35. 35
      Schütt, F.; Signetti, S.; Krüger, H.; Röder, S.; Smazna, D.; Kaps, S.; Gorb, S. N.; Mishra, Y. K.; Pugno, N. M.; Adelung, R. Hierarchical Self-Entangled Carbon Nanotube Tube Networks. Nat. Commun. 2017, 8, 1215  DOI: 10.1038/s41467-017-01324-7
    36. 36
      Feng, X.; Feng, L.; Jin, M.; Zhai, J.; Jiang, L.; Zhu, D. Reversible Super-Hydrophobicity to Super-Hydrophilicity Transition of Aligned ZnO Nanorod Films. J. Am. Chem. Soc. 2004, 126, 6263,  DOI: 10.1021/ja038636o
    37. 37
      Chen, H.-K. Kinetic Study on the Carbothermic Reduction of Zinc Oxide. Scand. J. Metall. 2001, 30, 292296,  DOI: 10.1034/j.1600-0692.2001.300503.x
    38. 38
      Zhao, J.; Xing, B.; Yang, H.; Pan, Q.; Li, Z.; Liu, Z. Growth of Carbon Nanotubes on Graphene by Chemical Vapor Deposition. New Carbon Mater. 2016, 31, 3136,  DOI: 10.1016/S1872-5805(16)60002-1
    39. 39
      Ferrari, A. C.; Robertson, J. Resonant Raman Spectroscopy of Disordered, Amorphous, and Diamondlike Carbon. Phys. Rev. B: Condens. Matter Mater. Phys. 2001, 64, 075414,  DOI: 10.1103/PhysRevB.64.075414
    40. 40
      Carey, T.; Cacovich, S.; Divitini, G.; Ren, J.; Mansouri, A.; Kim, J. M.; Wang, C.; Ducati, C.; Sordan, R.; Torrisi, F. Fully Inkjet-Printed Two-Dimensional Material Field-Effect Heterojunctions for Wearable and Textile Electronics. Nat. Commun. 2017, 8, 1202,  DOI: 10.1038/s41467-017-01210-2
    41. 41
      Ferrari, A. C.; Meyer, J. C.; Scardaci, V.; Casiraghi, C.; Lazzeri, M.; Mauri, F.; Piscanec, S.; Jiang, D.; Novoselov, K. S.; Roth, S. Raman Spectrum of Graphene and Graphene Layers. Phys. Rev. Lett. 2006, 97, 187401,  DOI: 10.1103/PhysRevLett.97.187401
    42. 42
      Ferrari, A.; Robertson, J. Interpretation of Raman Spectra of Disordered and Amorphous Carbon. Phys. Rev. B: Condens. Matter Mater. Phys. 2000, 61, 1409514107,  DOI: 10.1103/PhysRevB.61.14095
    43. 43
      Cuscó, R.; Alarcón-Lladó, E.; Ibáñez, J.; Artús, L.; Jiménez, J.; Wang, B.; Callahan, M. Temperature Dependence of Raman Scattering in ZnO. Phys. Rev. B: Condens. Matter Mater. Phys. 2007, 75, 165202,  DOI: 10.1103/PhysRevB.75.165202
    44. 44
      Fan, H.; Wang, L.; Zhao, K.; Li, N.; Shi, Z.; Ge, Z.; Jin, Z. Fabrication, Mechanical Properties, and Biocompatibility of Graphene-Reinforced Chitosan Composites. Biomacromolecules 2010, 11, 23452351,  DOI: 10.1021/bm100470q
    45. 45
      Ahadian, S.; Ramón-Azcón, J.; Estili, M.; Liang, X.; Ostrovidov, S.; Shiku, H.; Ramalingam, M.; Nakajima, K.; Sakka, Y.; Bae, H. H.; Matsue, T.; Khademhosseini, A. Hybrid Hydrogels Containing Vertically Aligned Carbon Nanotubes with Anisotropic Electrical Conductivity for Muscle Myofiber Fabrication. Sci. Rep. 2015, 4, 4271,  DOI: 10.1038/srep04271
    46. 46
      Shi, X.; Sitharaman, B.; Pham, Q. P.; Liang, F.; Wu, K.; Edward Billups, W.; Wilson, L. J.; Mikos, A. G. Fabrication of Porous Ultra-Short Single-Walled Carbon Nanotube Nanocomposite Scaffolds for Bone Tissue Engineering. Biomaterials 2007, 28, 40784090,  DOI: 10.1016/j.biomaterials.2007.05.033
    47. 47
      Chen, Z.; Ren, W.; Gao, L.; Liu, B.; Pei, S.; Cheng, H. M. Three-Dimensional Flexible and Conductive Interconnected Graphene Networks Grown by Chemical Vapour Deposition. Nat. Mater. 2011, 10, 424428,  DOI: 10.1038/nmat3001
    48. 48
      Zhao, G.; Zhang, X.; Lu, T. J.; Xu, F. Recent Advances in Electrospun Nanofibrous Scaffolds for Cardiac Tissue Engineering. Adv. Funct. Mater. 2015, 25, 57265738,  DOI: 10.1002/adfm.201502142
    49. 49
      Zuo, G.; Kang, S. G.; Xiu, P.; Zhao, Y.; Zhou, R. Interactions between Proteins and Carbon-Based Nanoparticles: Exploring the Origin of Nanotoxicity at the Molecular Level. Small 2013, 9, 15461556,  DOI: 10.1002/smll.201201381
    50. 50
      Suh, C. W.; Kim, M. Y.; Choo, J. B.; Kim, J. K.; Kim, H. K.; Lee, E. K. Analysis of Protein Adsorption Characteristics to Nano-Pore Silica Particles by Using Confocal Laser Scanning Microscopy. J. Biotechnol. 2004, 112, 267277,  DOI: 10.1016/j.jbiotec.2004.05.005
    51. 51
      Prager-Khoutorsky, M.; Lichtenstein, A.; Krishnan, R.; Rajendran, K.; Mayo, A.; Kam, Z.; Geiger, B.; Bershadsky, A. D. Fibroblast Polarization Is a Matrix-Rigidity-Dependent Process Controlled by Focal Adhesion Mechanosensing. Nat. Cell Biol. 2011, 13, 14571465,  DOI: 10.1038/ncb2370
    52. 52
      El-Mohri, H.; Wu, Y.; Mohanty, S.; Ghosh, G. Impact of Matrix Stiffness on Fibroblast Function. Mater. Sci. Eng., C 2017, 74, 146151,  DOI: 10.1016/j.msec.2017.02.001
    53. 53
      Sun, Y.; Chen, C. S.; Fu, J. Forcing Stem Cells to Behave: A Biophysical Perspective of the Cellular Microenvironment. Annu. Rev. Biophys. 2012, 41, 519542,  DOI: 10.1146/annurev-biophys-042910-155306
    54. 54
      Moreno-Arotzena, O.; Borau, C.; Movilla, N.; Vicente-Manzanares, M.; García-Aznar, J. M. Fibroblast Migration in 3D Is Controlled by Haptotaxis in a Non-Muscle Myosin II-Dependent Manner. Ann. Biomed. Eng. 2015, 43, 30253039,  DOI: 10.1007/s10439-015-1343-2
    55. 55
      Turner, C. E. Paxillin and Focal Adhesion Signalling. Nat. Cell Biol. 2000, 2, E231E236,  DOI: 10.1038/35046659
    56. 56
      Chiu, C.-L.; Aguilar, J. S.; Tsai, C. Y.; Wu, G.; Gratton, E.; Digman, M. M. A.; Kuo, J.; Han, X.; Hsiao, C.; Iii, J. Y. Nanoimaging of Focal Adhesion Dynamics in 3D. PLoS One 2014, 9, e99896,  DOI: 10.1371/journal.pone.0099896
    57. 57
      Ryoo, S. R.; Kim, Y. K.; Kim, M. H.; Min, D. H. Behaviors of NIH-3T3 Fibroblasts on Graphene/Carbon Nanotubes: Proliferation, Focal Adhesion, and Gene Transfection Studies. ACS Nano 2010, 4, 65876598,  DOI: 10.1021/nn1018279
    58. 58
      Cooper, G. M. The Cell: A Molecular Approach; ASM Press, 2000; Vol. 10.
    59. 59
      Serrano, M. C.; Patiño, J.; García-Rama, C.; Ferrer, M. L.; Fierro, J. L. G.; Tamayo, A.; Collazos-Castro, J. E.; Del Monte, F.; Gutiérrez, M. C. 3D Free-Standing Porous Scaffolds Made of Graphene Oxide as Substrates for Neural Cell Growth. J. Mater. Chem. B 2014, 2, 56985706,  DOI: 10.1039/C4TB00652F
    60. 60
      Avery, N. C.; Bailey, A. J.; Fratzl, P. Collagen: Structure and Mechanics; Springer, 2008.
    61. 61
      LeBaron, R. G.; Athanasiou, K. A. Extracellular Matrix Cell Adhesion Peptides: Functional Applications in Orthopedic Materials. Tissue Eng. 2000, 6, 85103,  DOI: 10.1089/107632700320720
    62. 62
      Marx, J.; Lewke, M. R. D.; Smazna, D.; Mishra, Y. K.; Adelung, R.; Schulte, K.; Fiedler, B. Processing, Growth Mechanism and Thermodynamic Calculations of Carbon Foam with a Hollow Tetrapodal Morphology – Aerographite. Appl. Surf. Sci. 2019, 470, 535542,  DOI: 10.1016/j.apsusc.2018.11.016
    63. 63
      Wang, F.; Torrisi, F.; Jiang, Z.; Popa, D.; Hasan, T.; Sun, Z.; Cho, W.; Ferrari, A. C. Graphene Passively Q-Switched Two-Micron Fiber Lasers. In Conference on Lasers and Electro-Optics (CLEO); OSA Technical Digest (Optical Society of America): San Jose, CA, 2012; p JW2A 72.
    64. 64
      Purdie, D. G.; Popa, D.; Wittwer, V. J.; Jiang, Z.; Bonacchini, G.; Torrisi, F.; Milana, S.; Lidorikis, E.; Ferrari, A. C. Few-Cycle Pulses from a Graphene Mode-Locked All-Fiber Laser. Appl. Phys. Lett. 2015, 106, 253101,  DOI: 10.1063/1.4922397
    65. 65
      Bianchi, V.; Carey, T.; Viti, L.; Li, L.; Linfield, E. H.; Davies, A. G.; Tredicucci, A.; Yoon, D.; Karagiannidis, P. G.; Lombardi, L.; Tomarchio, F.; Ferrari, A. C.; Torrisi, F.; Vitiello, M. S. Terahertz Saturable Absorbers from Liquid Phase Exfoliation of Graphite. Nat. Commun. 2017, 8, 15763,  DOI: 10.1038/ncomms15763
    66. 66
      Lotya, M.; Hernandez, Y.; King, P. J.; Smith, R. J.; Nicolosi, V.; Karlsson, L. S.; Blighe, F. M.; De, S.; Wang, Z.; McGovern, I. T.; Duesberg, G. S.; Coleman, J. N. Liquid Phase Production of Graphene by Exfoliation of Graphite in Surfactant/Water Solutions. J. Am. Chem. Soc. 2009, 131, 36113620,  DOI: 10.1021/ja807449u
    67. 67
      Carey, T.; Jones, C.; Le Moal, F.; Deganello, D.; Torrisi, F. Spray Coating Thin Films on Three-Dimensional Surfaces for a Semi-Transparent Capacitive Touch Device. ACS Appl. Mater. Interfaces 2018, 10, 1994819956,  DOI: 10.1021/acsami.8b02784
    68. 68
      Kravets, V. G.; Grigorenko, A. N.; Nair, R. R.; Blake, P.; Anissimova, S.; Novoselov, K. S.; Geim, A. K. Spectroscopic Ellipsometry of Graphene and an Exciton-Shifted van Hove Peak in Absorption. Phys. Rev. B: Condens. Matter Mater. Phys. 2010, 81, 155413,  DOI: 10.1103/PhysRevB.81.155413
    69. 69
      Selhuber-Unkel, C.; Erdmann, T.; López-García, M.; Kessler, H.; Schwarz, U. S.; Spatz, J. P. Cell Adhesion Strength Is Controlled by Intermolecular Spacing of Adhesion Receptors. Biophys. J. 2010, 98, 543551,  DOI: 10.1016/j.bpj.2009.11.001
    70. 70
      Wörle-Knirsch, J. M.; Pulskamp, K.; Krug, H. F. Oops They Did It Again! Carbon Nanotubes Hoax Scientists in Viability Assays. Nano Lett. 2006, 6, 12611268,  DOI: 10.1021/nl060177c
  • Supporting Information

    Supporting Information

    ARTICLE SECTIONS
    Jump To

    The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsami.8b17627.

    • Optical absorption spectrum of graphene ink (Figure S1); SEM images of ZnO templates (Figures S2–S4); different 3D carbon tube structures (Figure S5); SEM images of AG–CNTT structures (Figure S6); long-cycle compression test graph of AG–G scaffold (Figure S7); high-magnification fluorescence image of REF52 YFP-paxillin cells (Figure S8); supporting discussion (PDF)

    • Structural integrity of AG–G scaffold (AVI)


    Terms & Conditions

    Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

Pair your accounts.

Export articles to Mendeley

Get article recommendations from ACS based on references in your Mendeley library.

You’ve supercharged your research process with ACS and Mendeley!

STEP 1:
Click to create an ACS ID

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

Please note: If you switch to a different device, you may be asked to login again with only your ACS ID.

MENDELEY PAIRING EXPIRED
Your Mendeley pairing has expired. Please reconnect