Functionalized graphene oxide nanosheets with folic acid and silk fibroin as a novel nanobiocomposite for biomedical applications

Blog

HomeHome / Blog / Functionalized graphene oxide nanosheets with folic acid and silk fibroin as a novel nanobiocomposite for biomedical applications

Jun 14, 2023

Functionalized graphene oxide nanosheets with folic acid and silk fibroin as a novel nanobiocomposite for biomedical applications

Scientific Reports volume 12, Article number: 6205 (2022) Cite this article 3126 Accesses 10 Citations 9 Altmetric Metrics details In this paper, a novel graphene oxide-folic acid/silk fibroin

Scientific Reports volume 12, Article number: 6205 (2022) Cite this article

3126 Accesses

10 Citations

9 Altmetric

Metrics details

In this paper, a novel graphene oxide-folic acid/silk fibroin (GO-FA/SF) nanobiocomposite scaffold was designed and fabricated using affordable and non-toxic materials. The GO was synthesized using the hummer method, covalently functionalized with FA, and then easily conjugated with extracted SF via the freeze-drying process. For characterization of the scaffold, several techniques were employed: Fourier-transform infrared (FT-IR), field emission scanning electron microscopy (FE-SEM), energy dispersive X-ray (EDX), and thermogravimetric analysis (TGA). The cell viability method, hemolysis, and anti-biofilm assays were performed, exploring the biological capability of the nanobiocomposite. The cell viability percentages were 96.67, 96.35 and 97.23% for 24, 48, and 72 h, respectively, and its hemolytic effect was less than 10%. In addition, it was shown that this nanobiocomposite prevents the formation of Pseudomonas aeruginosa biofilm and has antibacterial activity.

Graphene, a 2D layered structure of carbon atoms, hexagonally arranged with a large surface area, has been extensively studied over the last decades. It has generated considerable interest due to its unique chemical properties1,2,3,4,5. Graphene is structurally capable of modifying and functionalizing its carbon platform to yield analog products, including graphene nanosheets, graphene oxide (GO), and reduced graphene oxide (rGO)6. According to the Hummers method, graphite reacts with strong oxidizing agents, such as potassium permanganate and concentrated sulphuric acid to produce a yellow colloidal dispersion named GO, so it can be said that GO is the oxidation product of graphite7. There is a direct relationship between graphene's final structure and chemical properties and its synthesis method and degree of oxidation8,9. There are similarities and differences between GO and graphene. Both of them have layered structures with a honeycomb lattice of carbon atoms bonded together with σ bonds in which the π orbitals of carbon atoms provide a delocalized network of electrons through the lattice. Unlike graphene, the presence of oxygen-containing functional groups at the surface of GO such as epoxides, alcohols, ketone carbonyls, and carboxylic groups10,11,12, able GO to be dispersed in aqueous media due to the formation of hydrogen bonds between polar functional groups and water molecules13,14. Besides, unique properties such as aqueous processability, the inhibition effect of bacterial growth, fluorescence quenching ability, easy scalability, inexpensive synthesis, and the ability to prepare the stable colloidal suspension make GO a potential candidate for various applications15,16,17,18,19,20,21,22,23,24. There is a broad spectrum of bioapplications of GO and its derivatives, including pH-dependent drug delivery, biosensing, bioimaging, and bone and skin tissue engineering25,26,27,28,29,30,31,32.

GO can be combined with synthetic or natural materials such as polymers, proteins, and vitamins to improve its imperfections, e.g., tensile strength, elasticity, and conductivity. Folic acid (FA), a water-soluble vitamin, is commonly used in biomaterial with potential bioapplications such as imaging systems33,34, therapeutic agents35, and drug carriers 36 due to its low cost, compatibility in a biological medium, and non-toxicity37. Conjugation of FA to GO platform via amidation led to the preparation of biocompatible materials with a wide range of applications from biosensors to targeted delivery of therapeutic agents for cancer treatment, including breast, ovarian, lung, and colon cancers38,39,40.

Silk fibroin (SF) of silkworms is a well-known natural protein consisting of fibroin and sericin. The first one structurally constructs the silk fibers, and the second acts as a glue and binds the fibroin fibers together41. For centuries SF was used as a suture, and nowadays, it is considered a potential candidate for biomaterials and scaffolds for tissue engineering since it has significant mechanical strength, elasticity, biocompatibility and biodegradability42,43,44. In comparison with other proteins, silk has superiority such as low risk of infection, low-cost processing, easy isolation and purification, large scale availability, excellent mechanical properties, biodegradation and above all the presence of easily accessible chemical groups for functional modifications with preserving its advantageous intrinsic properties45,46,47,48,49.

Nanobiomaterials have been reported to show favorable potential in different bioapplications including nanomedicine50, pharmacology51, and biomedical fields. Despite the prominent advantages of nanobiomaterials, certain challenges exist in their applications. For instance, they may cause notable biological side effects such as an inflammatory reaction and eventual cell death52. To utilize the bioapplications of these nanomaterials, their toxicity and interactions with biological systems must be evaluated and considered. Owing to these issues, in this study GO was synthesized using hummer method, then it was functionalized with FA through covalent bonds and finally to improve its durability and prepare a composite it was blended with SF to construct final nanobiocomposite (GO-FA/SF) with predicted properties such as biocompatibility, and non-toxicity.

Except for silkworm cocoons, the chemical reagents including dimethyl sulfuric acid, powdered graphite, potassium permanganate, anhydrous sodium carbonate, sodium nitrate, lithium bromide, N,N′-dicyclohexylcarbodiimide (DCC), 4-(dimethylamino)pyridine (DMAP), hydrogen peroxide, sodium chloride and the dialysis membrane (molecular weight cut-off = 14 kDa) were purchased from chemical Sigma-Aldrich company. All solvents used in this study and the fuming hydrochloric acid (37%) were also obtained from Merck company. The microorganism strain Pseudomonas aeruginosa (ATCCR 27853TM) was provided from Persian Type Culture Collection (PTCC, Karaj, Iran). Roswell Park Memorial Institute 1640 growth medium (RPMI) was purchased from Gibco BRL Life Technologies.

Fourier-transform infrared (FT-IR) spectra were recorded on a Shimadzu FT-8400 s model, Japan spectrometer in the region 400–4000 cm−1 using KBr pellets. The morphology characteristic, size distribution, and percentage elemental analysis of samples were conducted via field emission-scanning electron microscopy (FE-SEM); (ZEISS-Sigma VP model, Germany) operating at a 15 kV. Before the examination, samples were mounted onto a metal stub using double-sided carbon adhesive tape and covered with a thin layer of gold, with the aid of a direct current sputter technique (Agar Sputter Coater model, Agar Scientific, England). Furthermore, the sample's elemental composition was characterized by an energy-dispersive X-ray (EDX) device (SAMx model, France). Besides, the distribution pattern of structural elements was determined by elemental mapping images. The X-ray diffraction (XRD) pattern was analyzed using the Brucker X-ray diffractometer device (D8 Advanced Model, Germany), which was equipped with Cu-Kα radiation (λ = 0.154 nm, 40 kV, 40 mA). The thermogravimetric analysis (TGA) was conducted using Bahr-STA 504 instrument (Germany) under the argon atmosphere. The adsorption of the samples was measured using a microplate reader (STAT FAX 2100, BioTek, Winooski, USA) in the process of investigating the biological properties53.

According to our previous works, GO was prepared from native graphite flake and used the modified hummer method23,24,32,54. To activate the terminal carboxyl and carbonyl groups of GO, 1.85 g DCC and 0.3 g DMAP were added to 0.5 g of GO suspension in 100 mL DMSO. The mixture was sonicated for 20 min. In the next step, 0.3 g FA was added to the activated GO, and the suspension was sonicated for a further 20 min, and the obtained mixture was stirred overnight. After that, the mixture was centrifuged and washed two times with distilled water and two times with ethanol to remove DMSO and any nonconjugated FA from GO-FA. Finally, the mixture was washed with acetone and dried at room temperature to obtain powder product.

SF has been extracted from cocoons based on the previously reported methods through degumming, rehydration, and dialysis steps55,56. The FA decorated GO (GO-FA) (0.2 g) was sonicated in distilled water for a few minutes, and it was mixed with SF solution (5 mL). Later, the mixture solution was kept under stirring conditions for 4 h. Afterward, the obtained suspension mixture was poured into a petri dish for a further freeze-drying process. For this purpose, the sample was kept in a freezer (−70 °C) for 24 h. Afterward, the petri dish was placed at a freeze dryer device to sublimate the solvent for 24 h, and the dehydration process was performed at −60 °C with a constant pressure (0.1 bar)57.

First, 50 mg of Graphene Oxide-Folic Acid/Silk Fibroin (GO-FA/SF) nanobioocomposite was dispersed in 1 ml RPMI (Roswell Park Memorial Institute) medium using a shaker incubator for 24 h at 37 °C to obtain nanobiocomposite extract. Next, an MTT assay was performed to determine the toxicity of the GO-FA/SF nanobiocomposite. For this purpose, the Hu02 cell line (human skin fibroblast cells) obtained from the Pasteur Institute of Iran cell bank was cultured in 96-well plates (1 × 105 cell/well) at optimal conditions (37 °C, 5% CO2 in humidified incubator). Next, the growth media (10% FBS) was removed, and the cells were washed twice with Phosphate-buffered saline (PBS). New maintenance RPMI medium including GO-FA/SF extract was added and the cells were incubated for 24, 48, and 72 h. Attached RPMI without nanobiocomposite extract and cells in each well were also considered a negative control. A 10 μL solution of freshly prepared 5 mg/mL MTT (3-(4,5-dimethyl-2-thiazolyl)-2,5-diphenyl-2H-tetrazolium bromide) in PBS was added to each well and allowed to incubate at 37 °C for 4 h. After that, the media with MTT solution was removed, and isopropanol was added at 100 µL/well. Next, the plates were shaken gently to facilitate formazan crystal solubilization58. The absorbance was measured at 590 nm using a microplate reader (STAT FAX 2100, BioTek, Winooski, USA). Finally, the percentage of cell toxicity and cell viability was calculated as follows59:

This study was performed following the principles outlined in the declaration of Helsinki. Also, the experimental methods and the procedure for taking informed satisfaction were approved by Semnan University of Medical Sciences, Ethics Research Committee, and Informed consent was obtained from all participants. First, 50 mg of GO-FA/SF nanobiocomposite was dispersed in 1 ml PBS using a shaker incubator at 37 °C with two extraction times (24 and 72 h). Next, the hemolytic assay was performed to measure the hemocompatibility of the GO-FA/SF nanobiocomposite on human red blood cells (RBCs). A fresh blood sample was taken from an O negative blood type volunteer. A subsequent blood sample was diluted in PBS (1:20). Next, 100 μL of it was added to 100 μL of each GO-FA/SF extract (24 and 72 h) in a 96-well plate with three repetitions. In this test, 1% Triton X-100, which lyses almost all RBCs, was used as a positive control. 0.9% NaCl solution was also considered as a negative control. The plate was incubated at 37 °C for 1 h, and then samples were regained and centrifuged at 3000 rpm for 15 min60,61. The absorbance of each sample was determined by photometric analysis of supernatant at 414 nm using a microplate reader (STAT FAX 2100, BioTek, Winooski, USA) to record hemoglobin release, representing RBC damage. Finally, using the formula below, the hemolysis percentage of the samples was calculated62:

Tissue culture plate (TCP) anti-biofilm assay was used to evaluate the antimicrobial properties of the GO-FA/SF. First, pieces of nanobiocomposite and polystyrene (as a positive control) with dimensions of 1 cm2 were sterilized in 70% aqueous ethanol solution and dried in a sterilized incubator at 37 °C. Then, the selected bacteria (P. aeruginosa ATCC 27,853) at a concentration of 107 colony-forming unit (CFU)/mL in nutrient broth (NB) culture medium were prepared into sterilized tubes, and each piece was placed in them. Both tubes were incubated in a shaker incubator at 150 rpm at 37 °C for 24 h. Samples were then removed from the tubes and washed twice by PBS solution. Next, both samples were stained with 0.1% crystal violet solution for 5 min. It was then washed with 33% acetic acid solution to separate the bacteria from their surface. Finally, using a microplate reader (STAT FAX 2100, BioTek, Winooski, USA), the absorbance of the resulting solutions was evaluated at 570 nm63,64.

A t-test accomplished statistical analysis to compare all results by SPSS Statistics 22.0 software (SPSS Inc. Chicago, IL, USA). The values of P \(\ge\) 0.05 (*), P \(\le\) 0.05 (**), and P \(\le\) 0.001 (***) were considered statistically insignificant, significant, and very significant, respectively.

To prepare nano-sized GO-FA/SF biocomposite, the GO nanoflakes, FA vitamin, and fresh SF protein were employed. Four synthetic steps carried out the synthesis procedure. First, GO was synthesized using the modified hummer method. Second, it was functionalized with FA through coupling carboxylic acid of GO with the amine terminal of FA. Third, SF protein was extracted from silkworm cocoons via the degumming method. Finally, the surface-functionalized GO (GO-FA) blended with extracted SF using a freeze-drying process. The whole synthesis procedure is depicted in Fig. 1.

Synthetic process of GO-FA/SF nanobiocomposite.

After the synthesis and preparation of the final nanobiocomposite, its structural characterization was also conducted using various techniques, including FT-IR, FE-SEM, EDX, XRD, and TGA. In the following section, each analysis is discussed as well.

The success of the nanobiocomposite fabrication is confirmed by the analysis of the FT-IR spectra of the starting GO platform, GO-FA conjugate and final GO-FA/SF scaffold. As shown in the spectrum of synthetic GO (Fig. 2A), a broad band at the region of 3400–3350 cm-1 was assigned to the stretching vibration mode of hydroxyl groups (–OH)65. Two distinct absorption bands at 1728 and 1622 cm-1 are attributed to the stretching vibration modes of the carboxyl group (C=O) and unoxidized graphitic domain (C=C), respectively 66. Besides, small absorption bands related to the stretching vibration mode of alcohol groups (C–OH) and the stretching vibration mode of epoxy and alkoxy groups (C-O) were observed around 1222 and 1051 cm-167. Concerning Fig. 2B, amide bond formation during FA conjugation could be characterized by the presence of the stretching vibration of (N–H) at the region of (3470–3320 cm-1), bending vibration of (N–H) at 1570 cm-1 and the stretching vibration of (C-N) of the aliphatic amine group at 1218 cm-168. The characteristic peaks related to GO were observed at 1630, 1238 and 1045 cm-1 attributed to C=C, C–OH and C-OR vibrations respectively. Referring to the literature, it is evident that SF can be characterized by three vibrational bands in three regions including (1660–1625 cm-1), (1540–1520 cm-1), and (1270–1230 cm-1) related to (C=O) stretching vibration mode of amide I, the (N–H) bending vibration mode of amide II and the (C-N) stretching vibration mode of amide III, respectively 54,69. With these in mind, the SF peaks mentioned above appeared at 1660 cm-1 (random coil), 1533 cm-1 (random coil), and 1272 cm-1 (β-sheet) confirmed the presence of SF threads in produced nanobiocomposite structure. As shown in Fig. 2C, assigning absorption bands at 1280 and 1078 cm−1 were attributed to the C–OH and C-OR stretching vibration modes of GO. Also, its better to mention that, the characteristic peaks of GO related to C=O and C=C vibrations were overlapped by sharp peaks attributed to SF random coil vibrations at the region of 1660–1533 cm−1.

FT-IR spectra of GO (A), GO-FA (B) and GO-FA/SF nanobiocomposite scaffold (C).

The structural characterization of synthesized GO-FA and GO-FA/SF composites was carried out using energy dispersive x-ray spectroscopy to obtain the elemental composition. As depicted in Fig. 3A and B, the presence of two strong peaks related to carbon and oxygen elements and observing nitrogen signals confirm the preparation of GO and its functionalization with FA, respectively. Most importantly, the increase of the percentage of nitrogen content from 3.3 to 5.8 wt% (Fig. 3B) revealed that the SF was successfully conjuncted to GO-FA composite in the final step since SF brought extra nitrogen element with itself to final bionanocomposite.

EDX spectrum of GO-FA (A) and GO-FA/SF (B) bionanocomposites.

The morphological characterization of starting GO and final GO-FA/SF nanobiocomposite was performed using field emission scanning electron microscopy FE-SEM. As shown in Fig. 4A, the layered structure of GO flakes with wavy morphology is obvious. Considering Fig. 4B, following the functionalization of GO sheets with FA chains and its modification with SF polymers, it is revealed that the modified GO displayed smooth unilamellar morphology with reduced ruggedness, while it was preserved its characteristic structure observed in the native GO flakes.

Scanning electron microscopy (SEM) images of (A) GO, (B) GO-FA/SF nanobiocomposite at the scale of 200 nm.

X-ray diffraction (XRD) was used to investigate the crystal phase and determine the interlayer spacing for GO and final GO-FA/SF composite. As depicted in Fig. 5A, GO showed a sharp characteristic peak at 2θ = 10.7º (002) with the interlayer d spacing of 0.82 nm corresponding to GO sheets70. Figure 5B shows the semi-crystalline nature of GO-FA/SF composite as GO was functionalized with FA and loaded with SF. As illustrated in Fig. 5B, the well-defined peak of GO was observed at 2θ = 7.8º with increased interlayer d spacing of 11.2 nm, which shifted toward a lower diffraction angle compared to native GO. According to the literature, the increase of the oxygen containing groups in GO platform led to smaller degree of angle71. On the other hand, the entrapment of water molecules between the GO layers can enlarge the d-spacing of GO sheets72. Besides, A few diffractions (2θ = 15.6º, 22.0º, 28.8º) can be observed, indicating the semi-crystalline nature of the FA present in the composite structure73. Also, observing the crystalline peaks at 2θ = 9.3°, 20.1° confirms the presence of SF with its silk II crystalline conformation in prepared bionanocomposite74.

XRD pattern of (A) GO, (B) designed GO-FA/SF nanobiocomposite.

The thermogravimetric analysis was carried out to inspect the thermal stability of the prepared GO-FA/SF nanobiocomposite. Figure 6 shows the mass loss of the sample as a function of temperature, and it is revealed that product composite degrades mainly in four steps. According to the published results, the first significant mass loss at almost 170–260 Cº (≈40%) is mainly due to the decomposition of oxygen-containing functional groups of GO such as epoxide, hydroxyl, and carboxyl groups, yielding CO and CO275,76,77. The second mass reduction at a temperature range of 260–310 Cº (≈14%) corresponds to the decomposition of FA moieties that functionalized the GO platform73. So, we can calculate that the composite contains about 0.28 mg FA. The third degradation range at 310–455 Cº (≈12%) could be explained by cleavage of peptide bonds and breaking down of the side chain group of amino acids related to SF protein in nanobiocomposite structure78,79. Following the temperature increase to 800Cº, the mass loss by ≈34% refers to the decomposition of all of the remaining moieties with almost any residual.

Thermogravimetric curve of GO-FA/SF nanobiocomposite.

As shown in Fig. 7, the viability percentages of Hu02 cells treated with GO-FA/SF nanobiocomposite, after 24 and 48 h of incubation were 96.67 and 96.35%, respectively. This value increased to 97.23% after 72 h of incubation. Also, the effect of GO-FA/SF nanobiocomposite on morphology and shape of Hu02 cells after 72 h of incubation was imaged with a reverse microscope. It’s indicated that Hu02 cells retain their fibroblast shape after treatment with GO-FA/SF. Results are the average of three independent experiments. These results illustrate that GO-FA/SF nanobiocomposite had no toxic effects on Hu02 cells.

Picture of 96-well plate for MTT assay on Hu02 cell line (A) and histogram of the cell viability percentage (B) after different incubation times of nanobiocomposite (* = insignificant, P \(\ge\) 0.05 (. Untreated Hu02 cell line morphology (C) and Hu02 cell line morphology after treatment with GO-FA/SF nanobiocomposite after 72 h incubation (D).

The test results, which averaged the results of the three independent experiments, showed that the hemolytic effects of the nanobiocomposite extract after different extraction times (24 and 72 h) were below 10%, while 1% triton X-100 lysed almost all RBCs. (P \(\le\) 0.001) (Fig. 8). Based on the results, it can be said that GO-FA/SF nanobiocomposite is hemocompatible.

Hemolysis histogram of 1% Triton X-100 (positive control), 0.9% NaCl (negative control) and GO-FA/SF nanobiocomposite after different extraction times (24 h and 72 h) (* = insignificant, P \(\ge\) 0.05, *** = very significant, P \(\le\) 0.001)), comes with 96-well plate image.

As shown in Fig. 9, the adsorption rate of polystyrene (as a positive control) at 570 nm was 0.76, which was reduced to 0.63 for GO-FA/SF nanobiocomposite. So, it showed a small amount of anti-biofilm effects. The reported values are the average of three independent repetitions of the experiment.

Anti-biofilm histogram of polystyrene and GO-FA/SF nanobiocomposite pieces (* = insignificant, P \(\ge\) 0.05) comes with 96-well plate image.

This study designed and fabricated a novel nanobiocomposite using non-toxic and inexpensive starting materials. GO was synthesized through the modified hummer method, functionalized with FA, and then easily conjugated with extracted SF to aim this goal. The series of biological tests revealed remarkable features for obtained GO-FA/SF bionanocomposite. The MTT assay showed that treatment of Hu02 cells with GO-FA/SF nanobiocomposite for 24, 48, and 72 h of incubation time resulted in 96.67%, 96.35%, and 97.23% viability percentages with maintaining their fibroblast shape, which overall confirms the biocompatibility and non-toxicity of GO-FA/SF nanobiocomposite toward Hu02 cells. To explain the hemocompatibility of the produced nanobiocomposite, the hemolysis assay was performed, and the hemolysis percentage of 7.88 and 8.62% for 24 h and 72 h of extraction times were obtained. The results of the anti-biofilm assay showed that the fabricated nanobiocomposite scaffold could well inhibit the P. aeruginosa biofilm formation on its surface. Considering the structural characteristics and the satisfying results of biological assays, this novel nanobiocomposite could be regarded as more in biomedical uses such as wound healing and tissue engineering.

Boehm, H.-P., Clauss, A., Fischer, G. O. & Hofmann, U. Das adsorptionsverhalten sehr dünner kohlenstoff-folien. Z. Anorg. Allg. Chem. 316, 119–127 (1962).

Article CAS Google Scholar

Van Bommel, A. J., Crombeen, J. E. & Van Tooren, A. LEED and Auger electron observations of the SiC (0001) surface. Surf. Sci. 48, 463–472 (1975).

Article ADS Google Scholar

Boehm, H. P., Setton, R. & Stumpp, E. (Pergamon, 1986).

Boehm, H. P., Setton, R. & Stumpp, E. Nomenclature and terminology of graphite intercalation compounds (IUPAC Recommendations 1994). Pure Appl. Chem. 66, 1893–1901 (1994).

Article CAS Google Scholar

Wintterlin, J. & Bocquet, M. L. Graphene on metal surfaces. Surf. Sci. 603, 1841–1852 (2009).

Article ADS CAS Google Scholar

Sanchez, V. C., Jachak, A., Hurt, R. H. & Kane, A. B. Biological interactions of graphene-family nanomaterials: An interdisciplinary review. Chem. Res. Toxicol. 25, 15–34 (2012).

Article CAS PubMed Google Scholar

Hummers, W. S. Jr. & Offeman, R. E. Preparation of graphitic oxide. J. Am. Chem. Soc. 80, 1339–1339 (1958).

Article CAS Google Scholar

Shen, L. et al. Analysis of oxidation degree of graphite oxide and chemical structure of corresponding reduced graphite oxide by selecting different-sized original graphite. RSC Adv. 8, 17209–17217 (2018).

Article ADS CAS Google Scholar

Li, C. et al. Controlled synthesis of graphite oxide: Formation process, oxidation kinetics, and optimized conditions. Chem. Eng. Sci. 176, 319–328 (2018).

Article CAS Google Scholar

He, H., Klinowski, J., Forster, M. & Lerf, A. A new structural model for graphite oxide. Chem. Phys. Lett. 287, 53–56 (1998).

Article ADS CAS Google Scholar

Uhl, F. M. & Wilkie, C. A. Preparation of nanocomposites from styrene and modified graphite oxides. Polym. Degrad. Stab. 84, 215–226 (2004).

Article CAS Google Scholar

Stankovich, S. et al. Stable aqueous dispersions of graphitic nanoplatelets via the reduction of exfoliated graphite oxide in the presence of poly (sodium 4-styrenesulfonate). J. Mater. Chem. 16, 155–158 (2006).

Article CAS Google Scholar

Gao, W., Alemany, L. B., Ci, L. & Ajayan, P. M. New insights into the structure and reduction of graphite oxide. Nat. Chem. 1, 403–408 (2009).

Article CAS PubMed Google Scholar

Li, D., Müller, M. B., Gilje, S., Kaner, R. B. & Wallace, G. G. Processable aqueous dispersions of graphene nanosheets. Nat. Nanotechnol. 3, 101–105 (2008).

Article ADS CAS PubMed Google Scholar

Zare, P., Aleemardani, M., Seifalian, A., Bagher, Z. & Seifalian, A. M. Graphene oxide: Opportunities and challenges in biomedicine. Nanomaterials 11, 1083 (2021).

Article CAS PubMed PubMed Central Google Scholar

Behbudi, G. Mini review of graphene oxide for medical detection and applications. AANBT 1, 63–66 (2020).

CAS Google Scholar

Wu, X., Xing, Y., Zeng, K., Huber, K. & Zhao, J. X. Study of fluorescence quenching ability of graphene oxide with a layer of rigid and tunable silica spacer. Langmuir 34, 603–611 (2018).

Article CAS PubMed Google Scholar

Kumar, P. V. et al. Scalable enhancement of graphene oxide properties by thermally driven phase transformation. Nat. Chem. 6, 151–158 (2014).

Article CAS PubMed Google Scholar

Karthick, R., Brindha, M., Selvaraj, M. & Ramu, S. Stable colloidal dispersion of functionalized reduced graphene oxide in aqueous medium for transparent conductive film. J. Colloid Interface Sci. 406, 69–74 (2013).

Article ADS CAS PubMed Google Scholar

Aunkor, M. T. H. et al. Antibacterial activity of graphene oxide nanosheet against multidrug resistant superbugs isolated from infected patients. R. Soc. Open Sci. 7, 200640 (2020).

Article ADS CAS PubMed PubMed Central Google Scholar

Mohammed, H. et al. Antimicrobial mechanisms and effectiveness of graphene and graphene-functionalized biomaterials. A Scope Review. Front. Bioeng. Biotechnol. 8, 465 (2020).

Article PubMed Google Scholar

Liu, S. et al. Antibacterial activity of graphite, graphite oxide, graphene oxide, and reduced graphene oxide: membrane and oxidative stress. ACS Nano 5, 6971–6980 (2011).

Article CAS PubMed Google Scholar

Eivazzadeh-Keihan, R. et al. Fe3O4/GO@melamine-ZnO nanocomposite: A promising versatile tool for organic catalysis and electrical capacitance. Colloids Surf. A Physicochem. Eng. Asp. 587, 124335 (2020).

Article CAS Google Scholar

Eivazzadeh-Keihan, R. et al. Effective combination of rGO and CuO nanomaterials through Poly(p-phenylenediamine) texture: Utilizing it as an excellent supercapacitor. Energy Fuels 35, 10869–10877 (2021).

Article CAS Google Scholar

Anirudhan, T. S., Sekhar, V. C. & Athira, V. S. Graphene oxide based functionalized chitosan polyelectrolyte nanocomposite for targeted and pH responsive drug delivery. Int. J. Biol. Macromol. 150, 468–479 (2020).

Article CAS PubMed Google Scholar

Kim, J., Park, S.-J. & Min, D.-H. Emerging approaches for graphene oxide biosensor. Anal. Chem. 89, 232–248 (2017).

Article CAS PubMed Google Scholar

Zang, Z. et al. Tunable photoluminescence of water-soluble AgInZnS–graphene oxide (GO) nanocomposites and their application in-vivo bioimaging. Sens. Actuators B Chem. 252, 1179–1186 (2017).

Article CAS Google Scholar

Gv, Y. D., Prabhu, A., Anil, S. & Venkatesan, J. Preparation and characterization of dexamethasone loaded sodium alginate-graphene oxide microspheres for bone tissue engineering. J. Drug Deliv. Sci. Technol., 102624 (2021).

Narayanan, K. B., Park, G. T. & Han, S. S. Electrospun poly (vinyl alcohol)/reduced graphene oxide nanofibrous scaffolds for skin tissue engineering. Colloids Surf. B 191, 110994 (2020).

Article CAS Google Scholar

Mokhtarzadeh, A. et al. Nanomaterial-based biosensors for detection of pathogenic virus. TrAC Trends Analyt. Chem. 97, 445–457 (2017).

Article CAS PubMed Google Scholar

Eivazzadeh-Keihan, R. et al. Carbon based nanomaterials for tissue engineering of bone: Building new bone on small black scaffolds: A review. J. Adv. Res. 18, 185–201 (2019).

Article CAS PubMed PubMed Central Google Scholar

Eyvazzadeh-Keihan, R., Bahrami, N., Taheri-Ledari, R. & Maleki, A. Highly facilitated synthesis of phenyl(tetramethyl)acridinedione pharmaceuticals by a magnetized nanoscale catalytic system, constructed of GO, Fe3O4 and creatine. Diam. Relat. Mater. 102, 107661 (2020).

Article ADS CAS Google Scholar

Batra, B., Narwal, V., Kalra, V., Sharma, M. & Rana, J. S. Folic acid biosensors: A review. Process Biochem. 92, 343–354 (2020).

Article CAS Google Scholar

Müller, C. & Schibli, R. Folic acid conjugates for nuclear imaging of folate receptor–positive cancer. J. Nucl. Med. 52, 1–4 (2011).

Article CAS PubMed Google Scholar

Li, X. et al. Folic acid as a versatile motif to construct molecular hydrogelators through conjugations with hydrophobic therapeutic agents. J. Mater. Chem. 22, 21838–21840 (2012).

Article ADS CAS Google Scholar

Karimi, S. & Namazi, H. A photoluminescent folic acid-derived carbon dot functionalized magnetic dendrimer as a pH-responsive carrier for targeted doxorubicin delivery. New J. Chem. 45, 6397–6405 (2021).

Article CAS Google Scholar

Likis, F. Folic acid. J. Midwifery Womens Health 61, 797–798 (2016).

Article Google Scholar

Xu, J., Zeng, F., Wu, H., Yu, C. & Wu, S. Dual-targeting nanosystem for enhancing photodynamic therapy efficiency. ACS Appl. Mater. Interfaces 7, 9287–9296 (2015).

Article CAS PubMed Google Scholar

Vinothini, K., Rajendran, N. K., Ramu, A., Elumalai, N. & Rajan, M. Folate receptor targeted delivery of paclitaxel to breast cancer cells via folic acid conjugated graphene oxide grafted methyl acrylate nanocarrier. Biomed. Pharmacother 110, 906–917 (2019).

Article CAS PubMed Google Scholar

Park, Y. H., Park, S. Y. & In, I. Direct noncovalent conjugation of folic acid on reduced graphene oxide as anticancer drug carrier. J. Ind. Eng. Chem. 30, 190–196 (2015).

Article CAS Google Scholar

Rockwood, D. N. et al. Materials fabrication from Bombyx mori silk fibroin. Nat. Protoc. 6, 1612–1631 (2011).

Article CAS PubMed Google Scholar

Jiang, C. et al. Mechanical properties of robust ultrathin silk fibroin films. Adv. Func. Mater. 17, 2229–2237 (2007).

Article CAS Google Scholar

Horan, R. L. et al. In vitro degradation of silk fibroin. Biomaterials 26, 3385–3393 (2005).

Article CAS PubMed Google Scholar

Yang, Y. et al. Biocompatibility evaluation of silk fibroin with peripheral nerve tissues and cells in vitro. Biomaterials 28, 1643–1652 (2007).

Article CAS PubMed Google Scholar

Farokhi, M., Mottaghitalab, F., Fatahi, Y., Khademhosseini, A. & Kaplan, D. L. Overview of silk fibroin use in wound dressings. Trends Biotechnol. 36, 907–922 (2018).

Article CAS PubMed Google Scholar

Chen, S., Liu, M., Huang, H., Cheng, L. & Zhao, H.-P. Mechanical properties of Bombyx mori silkworm silk fibre and its corresponding silk fibroin filament: A comparative study. Mater. Des. 181, 108077 (2019).

Article Google Scholar

Umuhoza, D. et al. Strategies for tuning the biodegradation of silk fibroin-based materials for tissue engineering applications. ACS Biomater. Sci. Eng. 6, 1290–1310 (2020).

Article CAS PubMed Google Scholar

Wu, J. et al. Functionalization of silk fibroin electrospun scaffolds via BMSC affinity peptide grafting through oxidative self-polymerization of dopamine for bone regeneration. ACS Appl. Mater. Interfaces 11, 8878–8895 (2019).

Article CAS PubMed Google Scholar

Farokhi, M., Mottaghitalab, F., Reis, R. L., Ramakrishna, S. & Kundu, S. C. Functionalized silk fibroin nanofibers as drug carriers: Advantages and challenges. J. Control. Release 321, 324–347 (2020).

Article CAS PubMed Google Scholar

Su, S. & Kang, P. M. Systemic review of biodegradable nanomaterials in nanomedicine. Nanomaterials 10, 656 (2020).

Article CAS PubMed Central Google Scholar

Cao, Y. Applications of cellulose nanomaterials in pharmaceutical science and pharmacology. Express Polym. Lett. 12 (2018).

Padmanabhan, J. & Kyriakides, T. R. Nanomaterials, inflammation, and tissue engineering. Wiley Interdiscip. Rev.: Nanomed. Nanobiotechnol. 7, 355–370 (2015).

CAS Google Scholar

Eivazzadeh-Keihan, R. et al. Chitosan hydrogel/silk fibroin/Mg(OH)2 nanobiocomposite as a novel scaffold with antimicrobial activity and improved mechanical properties. Sci. Rep. 11, 650 (2021).

Article CAS PubMed PubMed Central Google Scholar

Eivazzadeh-Keihan, R., Radinekiyan, F., Madanchi, H., Aliabadi, H. A. M. & Maleki, A. Graphene oxide/alginate/silk fibroin composite as a novel bionanostructure with improved blood compatibility, less toxicity and enhanced mechanical properties. Carbohydr. Polym. 248, 116802 (2020).

Article CAS PubMed Google Scholar

Ebrahimi, A., Sadrjavadi, K., Hajialyani, M., Shokoohinia, Y. & Fattahi, A. Preparation and characterization of silk fibroin hydrogel as injectable implants for sustained release of Risperidone. Drug Dev. Ind. Pharm. 44, 199–205 (2018).

Article CAS PubMed Google Scholar

Huang, W., Ling, S., Li, C., Omenetto, F. G. & Kaplan, D. L. Silkworm silk-based materials and devices generated using bio-nanotechnology. Chem. Soc. Rev. 47, 6486–6504 (2018).

Article CAS PubMed PubMed Central Google Scholar

Mehrabani, M. G. et al. Chitin/silk fibroin/TiO(2) bio-nanocomposite as a biocompatible wound dressing bandage with strong antimicrobial activity. Int. J. Biol. Macromol. 116, 966–976 (2018).

Article CAS PubMed Google Scholar

Bahojb Noruzi, E. et al. Design of a thiosemicarbazide-functionalized Calix[4]arene ligand and related transition metal complexes: Synthesis, characterization, and biological studies. Front. Chem. 7, (2019).

Eivazzadeh-Keihan, R. et al. Investigation of the biological activity, mechanical properties and wound healing application of a novel scaffold based on lignin–agarose hydrogel and silk fibroin embedded zinc chromite nanoparticles. RSC Adv. 11, 17914–17923 (2021).

Article ADS CAS Google Scholar

Eivazzadeh-Keihan, R. et al. Alginate hydrogel-polyvinyl alcohol/silk fibroin/magnesium hydroxide nanorods: A novel scaffold with biological and antibacterial activity and improved mechanical properties. Int. J. Biol. Macromol. 162, 1959–1971 (2020).

Article CAS PubMed Google Scholar

Bahojb Noruzi, E. et al. Synthesis, crystal structure, and biological activity of a multidentate calix [4] arene ligand doubly functionalized by 2-hydroxybenzeledene-thiosemicarbazone. Molecules 25, 370 (2020).

Article CAS PubMed Central Google Scholar

Jaganathan, S. K., Mani, M. P., Ayyar, M., Krishnasamy, N. P. & Nageswaran, G. Blood compatibility and physicochemical assessment of novel nanocomposite comprising polyurethane and dietary carotino oil for cardiac tissue engineering applications. J. Appl. Polym. Sci. 135, 45691 (2018).

Article CAS Google Scholar

Haney, E. F., Trimble, M. J., Cheng, J. T., Vallé, Q. & Hancock, R. E. W. Critical assessment of methods to quantify biofilm growth and evaluate antibiofilm activity of host defence peptides. Biomolecules 8, 29 (2018).

Article CAS PubMed Central Google Scholar

Eivazzadeh-Keihan, R. et al. Hybrid bionanocomposite containing magnesium hydroxide nanoparticles embedded in a carboxymethyl cellulose hydrogel plus silk fibroin as a scaffold for wound dressing applications. ACS Appl. Mater. Interfaces (2021).

Verma, S. & Dutta, R. K. A facile method of synthesizing ammonia modified graphene oxide for efficient removal of uranyl ions from aqueous medium. RSC Adv. 5, 77192–77203 (2015).

Article ADS CAS Google Scholar

Emiru, T. F. & Ayele, D. W. Controlled synthesis, characterization and reduction of graphene oxide: A convenient method for large scale production. Egypt. J. Basic Appl. Sci. 4, 74–79 (2017).

Google Scholar

Khandelwal, M. & Kumar, A. One-step chemically controlled wet synthesis of graphene nanoribbons from graphene oxide for high performance supercapacitor applications. J. Mater. Chem. A 3, 22975–22988 (2015).

Article CAS Google Scholar

Chauhan, G. et al. “Gold nanoparticles composite-folic acid conjugated graphene oxide nanohybrids” for targeted chemo-thermal cancer ablation: In vitro screening and in vivo studies. Eur. J. Pharm. Sci. 96, 351–361 (2017).

Article CAS PubMed Google Scholar

Ha, S.-W., Tonelli, A. E. & Hudson, S. M. Structural studies of bombyx m ori silk fibroin during regeneration from solutions and wet fiber spinning. Biomacromol 6, 1722–1731 (2005).

Article CAS Google Scholar

Hidayah, N. M. S. et al. 1 edn 150002 (AIP Publishing LLC).

Aziz, M., Halim, F. S. A. & Jaafar, J. Preparation and characterization of graphene membrane electrode assembly. J. Teknol. 69 (2014).

Pham, V. H. et al. Chemical functionalization of graphene sheets by solvothermal reduction of a graphene oxide suspension in N-methyl-2-pyrrolidone. J. Mater. Chem. 21, 3371–3377 (2011).

Article CAS Google Scholar

Ruman, U. et al. Synthesis and characterization of chitosan-based nanodelivery systems to enhance the anticancer effect of sorafenib drug in hepatocellular carcinoma and colorectal adenocarcinoma cells. Nanomaterials 11, 497 (2021).

Article CAS PubMed PubMed Central Google Scholar

Zhang, X. & Pan, Z. Microstructure transitions and dry-wet spinnability of silk fibroin protein from waste silk quilt. Polymers 11, 1622 (2019).

Article MathSciNet CAS PubMed Central Google Scholar

Zhang, Y. et al. Facile synthesis of well-dispersed graphene by γ-ray induced reduction of graphene oxide. J. Mater. Chem. 22, 13064–13069 (2012).

Article CAS Google Scholar

Wang, S., Li, Y., Fan, X., Zhang, F. & Zhang, G. β-Cyclodextrin functionalized graphene oxide: An efficient and recyclable adsorbent for the removal of dye pollutants. Front. Chem. Sci. Eng. 9, 77–83 (2015).

Article ADS CAS Google Scholar

Alosmanov, R. et al. Effect of functional groups on the thermal degradation of phosphorus-and phosphorus/nitrogen-containing functional polymers. J. Therm. Anal. Calorim. 130, 799–812 (2017).

Article CAS Google Scholar

Gupta, P. et al. Mimicking form and function of native small diameter vascular conduits using mulberry and non-mulberry patterned silk films. ACS Appl. Mater. Interfaces 8, 15874–15888 (2016).

Article CAS PubMed Google Scholar

Ho, M.-p., Wang, H. & Lau, K.-t. 1–8 (Composites Australia).

Download references

The authors gratefully acknowledge the partial support from the Research Council of the Iran University of Science and Technology. Also the authors thank of Ethics Research Committee and Biotechnology Research Center from Semnan University of Medical Sciences.

Catalysts and Organic Synthesis Research Laboratory, Department of Chemistry, Iran University of Science and Technology, 16846-13114, Tehran, Iran

Reza Eivazzadeh-Keihan, Farkhondeh Alimirzaloo & Ali Maleki

Protein Chemistry Laboratory, Department of Medical Biotechnology, Biotechnology Research Center, Pasteur Institute of Iran, Tehran, Iran

Hooman Aghamirza Moghim Aliabadi

Advanced Chemical Studies Lab, Department of Chemistry, K. N. Toosi University of Technology, Tehran, Iran

Hooman Aghamirza Moghim Aliabadi

Department of Inorganic Chemistry, Faculty of Chemistry, University of Tabriz, Tabriz, Iran

Ehsan Bahojb Noruzi

Department of Chemistry, Iran University of Science and Technology, 16846-13114, Tehran, Iran

Ali Reza Akbarzadeh

Department of Biotechnology, School of Medicine, Semnan University of Medical Sciences, Semnan, Iran

Hamid Madanchi

Drug Design and Bioinformatics Unit, Department of Medical Biotechnology, Biotechnology Research Center, Pasteur Institute of Iran, Tehran, Iran

Hamid Madanchi

Endocrinology and Metabolism Research Center, Endocrinology and Metabolism Clinical Sciences Institute, Tehran University of Medical Sciences, Tehran, Iran

Mohammad Mahdavi

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

You can also search for this author in PubMed Google Scholar

R.E.-K.: Substantial contributions to the conception, Design of the work, have drafted the work, Writing—Review & Editing, Analysis and interpretation of data and wrote the main manuscript. F.A.: Have drafted the work, Analysis and interpretation of data, substantively revised it. H.A.M.A.: Analysis and interpretation of data, substantively revised it, wrote the main manuscript and prepared figures; E.B.N.: Analysis and interpretation of data, substantively revised it and wrote the main manuscript. A.R.A.: Analysis and interpretation of data, substantively revised it. M.M.: The corresponding author of current study, Analysis and interpretation of data, substantively revised it. A.M.: The main corresponding (submitting) author of current study, Substantial contributions to the conception, Design of the work, have drafted the work, Writing—Review & Editing, substantively revised it. H.M.: The corresponding author of current study, Analysis and interpretation of data, substantively revised it. All authors are aware about ethical approval, guidelines followed in the manuscript and informed consent was obtained from all participants. Figure 1 was drawn by our co-author (Ehsan Bahojb Noruzid), Figs. 7, 8 and 9: 3rd co-author (Hooman Aghamirza Moghim Aliabadi).

Correspondence to Ali Maleki, Hamid Madanchi or Mohammad Mahdavi.

The authors declare no competing interests.

Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Open Access This article is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons licence, and indicate if changes were made. The images or other third party material in this article are included in the article's Creative Commons licence, unless indicated otherwise in a credit line to the material. If material is not included in the article's Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by/4.0/.

Reprints and Permissions

Eivazzadeh-Keihan, R., Alimirzaloo, F., Aghamirza Moghim Aliabadi, H. et al. Functionalized graphene oxide nanosheets with folic acid and silk fibroin as a novel nanobiocomposite for biomedical applications. Sci Rep 12, 6205 (2022). https://doi.org/10.1038/s41598-022-10212-0

Download citation

Received: 07 October 2021

Accepted: 21 March 2022

Published: 13 April 2022

DOI: https://doi.org/10.1038/s41598-022-10212-0

Anyone you share the following link with will be able to read this content:

Sorry, a shareable link is not currently available for this article.

Provided by the Springer Nature SharedIt content-sharing initiative

Cancer Nanotechnology (2023)

BMC Microbiology (2023)

Journal of Inorganic and Organometallic Polymers and Materials (2022)

By submitting a comment you agree to abide by our Terms and Community Guidelines. If you find something abusive or that does not comply with our terms or guidelines please flag it as inappropriate.