• Alzheimer's disease & dementia
  • Arthritis & Rheumatism
  • Attention deficit disorders
  • Autism spectrum disorders
  • Biomedical technology
  • Diseases, Conditions, Syndromes
  • Endocrinology & Metabolism
  • Gastroenterology
  • Gerontology & Geriatrics
  • Health informatics
  • Inflammatory disorders
  • Medical economics
  • Medical research
  • Medications
  • Neuroscience
  • Obstetrics & gynaecology
  • Oncology & Cancer
  • Ophthalmology
  • Overweight & Obesity
  • Parkinson's & Movement disorders
  • Psychology & Psychiatry
  • Radiology & Imaging
  • Sleep disorders
  • Sports medicine & Kinesiology
  • Vaccination
  • Breast cancer
  • Cardiovascular disease
  • Chronic obstructive pulmonary disease
  • Colon cancer
  • Coronary artery disease
  • Heart attack
  • Heart disease
  • High blood pressure
  • Kidney disease
  • Lung cancer
  • Multiple sclerosis
  • Myocardial infarction
  • Ovarian cancer
  • Post traumatic stress disorder
  • Rheumatoid arthritis
  • Schizophrenia
  • Skin cancer
  • Type 2 diabetes
  • Full List »

Journal of Biomedical Materials Research

The Journal of Biomedical Materials Research is an international, interdisciplinary, English-language publication of original contributions concerning studies of the preparation, performance, and evaluation of biomaterials; the chemical, physical, toxicological, and mechanical behavior of materials in physiological environments; and the response of blood and tissues to biomaterials. The Journal publishes peer-reviewed articles on all relevant biomaterial topics including the science and technology of alloys, polymers, ceramics, and reprocessed animal and human tissues in surgery, dentistry, artificial organs, and other medical devices. The Journal also publishes articles in interdisciplinary areas such as tissue engineering and controlled release technology where biomaterials play a significant role in the performance of the medical device.

Some content from Wikipedia , licensed under CC BY-SA

  • Date 6 hours 12 hours 1 day 3 days all
  • Rank Last day 1 week 1 month all
  • LiveRank Last day 1 week 1 month all
  • Popular Last day 1 week 1 month all

Biomaterial proves capable of accelerating bone regeneration

Researchers at São Paulo State University's Botucatu Institute of Biosciences (IBB-UNESP) in Brazil have developed a novel biomaterial that speeds up osteoblast (bone cell) differentiation. The invention has the potential ...

Nov 21, 2023

journal of biomedical materials research

Researchers develop new model for quickly evaluating potential cervical cancer drugs

Researchers at Oregon State University have created a means of speeding up and improving the evaluation process for drugs used to combat cervical cancer.

Mar 31, 2023

journal of biomedical materials research

E-mail newsletter

Journal of Biomedical Materials Research

journal of biomedical materials research

Subject Area and Category

  • Biomedical Engineering
  • Biomaterials

HeteroCorporation

Publication type

journal of biomedical materials research

The set of journals have been ranked according to their SJR and divided into four equal groups, four quartiles. Q1 (green) comprises the quarter of the journals with the highest values, Q2 (yellow) the second highest values, Q3 (orange) the third highest values and Q4 (red) the lowest values.

CategoryYearQuartile
Biomaterials1999Q1
Biomaterials2000Q1
Biomaterials2001Q1
Biomaterials2002Q1
Biomaterials2003Q1
Biomaterials2004Q1
Biomaterials2005Q1
Biomaterials2006Q3
Biomaterials2018Q4
Biomaterials2019Q4
Biomedical Engineering1999Q1
Biomedical Engineering2000Q1
Biomedical Engineering2001Q1
Biomedical Engineering2002Q1
Biomedical Engineering2003Q1
Biomedical Engineering2004Q1
Biomedical Engineering2005Q1
Biomedical Engineering2006Q2
Biomedical Engineering2018Q4
Biomedical Engineering2019Q4

The SJR is a size-independent prestige indicator that ranks journals by their 'average prestige per article'. It is based on the idea that 'all citations are not created equal'. SJR is a measure of scientific influence of journals that accounts for both the number of citations received by a journal and the importance or prestige of the journals where such citations come from It measures the scientific influence of the average article in a journal, it expresses how central to the global scientific discussion an average article of the journal is.

YearSJR
19991.384
20001.171
20011.460
20021.599
20031.587
20041.074
20051.467
20060.623
20180.124
20190.125

Evolution of the number of published documents. All types of documents are considered, including citable and non citable documents.

YearDocuments
1999355
2000433
2001395
2002428
20030
20040
20050
20060
20180
20190

This indicator counts the number of citations received by documents from a journal and divides them by the total number of documents published in that journal. The chart shows the evolution of the average number of times documents published in a journal in the past two, three and four years have been cited in the current year. The two years line is equivalent to journal impact factor ™ (Thomson Reuters) metric.

Cites per documentYearValue
Cites / Doc. (4 years)19992.190
Cites / Doc. (4 years)20002.434
Cites / Doc. (4 years)20012.689
Cites / Doc. (4 years)20023.150
Cites / Doc. (4 years)20033.277
Cites / Doc. (4 years)20044.070
Cites / Doc. (4 years)20054.513
Cites / Doc. (4 years)20064.465
Cites / Doc. (4 years)20180.000
Cites / Doc. (4 years)20190.000
Cites / Doc. (3 years)19992.190
Cites / Doc. (3 years)20002.350
Cites / Doc. (3 years)20012.608
Cites / Doc. (3 years)20022.948
Cites / Doc. (3 years)20033.124
Cites / Doc. (3 years)20043.800
Cites / Doc. (3 years)20054.355
Cites / Doc. (3 years)20060.000
Cites / Doc. (3 years)20180.000
Cites / Doc. (3 years)20190.000
Cites / Doc. (2 years)19992.006
Cites / Doc. (2 years)20002.183
Cites / Doc. (2 years)20012.288
Cites / Doc. (2 years)20022.679
Cites / Doc. (2 years)20032.796
Cites / Doc. (2 years)20043.666
Cites / Doc. (2 years)20050.000
Cites / Doc. (2 years)20060.000
Cites / Doc. (2 years)20180.000
Cites / Doc. (2 years)20190.000

Evolution of the total number of citations and journal's self-citations received by a journal's published documents during the three previous years. Journal Self-citation is defined as the number of citation from a journal citing article to articles published by the same journal.

CitesYearValue
Self Cites1999441
Self Cites2000542
Self Cites2001619
Self Cites2002516
Self Cites20030
Self Cites20040
Self Cites20050
Self Cites20060
Self Cites20180
Self Cites20190
Total Cites19992061
Total Cites20002446
Total Cites20013044
Total Cites20023487
Total Cites20033924
Total Cites20043127
Total Cites20051864
Total Cites20060
Total Cites20180
Total Cites20190

Evolution of the number of total citation per document and external citation per document (i.e. journal self-citations removed) received by a journal's published documents during the three previous years. External citations are calculated by subtracting the number of self-citations from the total number of citations received by the journal’s documents.

CitesYearValue
External Cites per document19991.722
External Cites per document20001.829
External Cites per document20012.078
External Cites per document20022.511
External Cites per document20033.124
External Cites per document20043.800
External Cites per document20054.355
External Cites per document20060
External Cites per document20180
External Cites per document20190
Cites per document19992.190
Cites per document20002.350
Cites per document20012.608
Cites per document20022.948
Cites per document20033.124
Cites per document20043.800
Cites per document20054.355
Cites per document20060.000
Cites per document20180.000
Cites per document20190.000

International Collaboration accounts for the articles that have been produced by researchers from several countries. The chart shows the ratio of a journal's documents signed by researchers from more than one country; that is including more than one country address.

YearInternational Collaboration
199914.37
200015.94
200118.48
200219.16
20030
20040
20050
20060
20180
20190

Not every article in a journal is considered primary research and therefore "citable", this chart shows the ratio of a journal's articles including substantial research (research articles, conference papers and reviews) in three year windows vs. those documents other than research articles, reviews and conference papers.

DocumentsYearValue
Non-citable documents19997
Non-citable documents20009
Non-citable documents200110
Non-citable documents20028
Non-citable documents20035
Non-citable documents20042
Non-citable documents20052
Non-citable documents20060
Non-citable documents20180
Non-citable documents20190
Citable documents1999934
Citable documents20001032
Citable documents20011157
Citable documents20021175
Citable documents20031251
Citable documents2004821
Citable documents2005426
Citable documents20060
Citable documents20180
Citable documents20190

Ratio of a journal's items, grouped in three years windows, that have been cited at least once vs. those not cited during the following year.

DocumentsYearValue
Uncited documents1999259
Uncited documents2000261
Uncited documents2001284
Uncited documents2002261
Uncited documents2003254
Uncited documents2004122
Uncited documents200555
Uncited documents20060
Uncited documents20180
Uncited documents20190
Cited documents1999682
Cited documents2000780
Cited documents2001883
Cited documents2002922
Cited documents20031002
Cited documents2004701
Cited documents2005373
Cited documents20060
Cited documents20180
Cited documents20190

Evolution of the percentage of female authors.

YearFemale Percent
199926.74
200024.56
200124.53
200225.36
20030.00
20040.00
20050.00
20060.00
20180.00
20190.00

Evolution of the number of documents cited by public policy documents according to Overton database.

DocumentsYearValue
Overton19991
Overton20000
Overton20011
Overton200210
Overton20030
Overton20040
Overton20050
Overton20060
Overton20180
Overton20190

Evoution of the number of documents related to Sustainable Development Goals defined by United Nations. Available from 2018 onwards.

DocumentsYearValue
SDG20180
SDG20190

Scimago Journal & Country Rank

Leave a comment

Name * Required

Email (will not be published) * Required

* Required Cancel

The users of Scimago Journal & Country Rank have the possibility to dialogue through comments linked to a specific journal. The purpose is to have a forum in which general doubts about the processes of publication in the journal, experiences and other issues derived from the publication of papers are resolved. For topics on particular articles, maintain the dialogue through the usual channels with your editor.

Scimago Lab

Follow us on @ScimagoJR Scimago Lab , Copyright 2007-2024. Data Source: Scopus®

journal of biomedical materials research

Cookie settings

Cookie Policy

Legal Notice

Privacy Policy

Journal of Biomedical Materials Research Part A

Journal Abbreviation: J BIOMED MATER RES A Journal ISSN: 1549-3296

About Journal of Biomedical Materials Research Part A

Year Impact Factor (IF) Total Articles Total Cites
2023 (2024 update) 3.9 - -
2022 4.9 - 19115
2021 4.854 - 20842
2020 4.396 323 20820
2019 3.525 249 17615
2018 3.221 304 17201
2017 3.231 327 17464
2016 3.076 292 15670
2015 3.263 393 15442
2014 3.369 464 15555
2013 2.841 402 13513
2012 2.834 384 12128
2011 2.625 270 11292
2010 3.044 612 11035

You may also be interested in the following journals

  • ► Journal of Biomedical Materials Research Part B-Applied Biomaterials
  • ► Faseb Journal
  • ► Acta Biomaterialia
  • ► Journal of Biological Chemistry
  • ► Bone
  • ► Journal of Bone and Mineral Research
  • ► Ieee Transactions on Nanobioscience
  • ► Bmj Open
  • ► Pharmaceutical Research
  • ► Journal of Alloys and Compounds

Top Journals in engineering

  • Nature Nanotechnology
  • Nature Biotechnology
  • Progress in Energy and Combustion Science
  • Materials Today
  • Materials Science & Engineering R-Reports
  • Progress in Materials Science
  • Journal of Statistical Software
  • Advanced Materials
  • IEEE Communications Surveys and Tutorials
  • Annual Review of Materials Research

Journal Impact

Home

  • Biomaterials Forum
  • JBMR-A&B
  • Mentor Program

journal of biomedical materials research

Clinicians, researchers, and students from medical device manufacturers, academic research labs, and operating rooms across the country and around the world belong to SFB. This interdisciplinary gathering is a critical pathway to identifying issues with current technologies, and for translating academic research to clinical practice.

journal of biomedical materials research

2024 Regional Symposia

Join our 2024 Regional Symposia on September 19th-20th at one of these six distinguished universities:  

Northeast: Northeastern University Midwest: Case Western Reserve University Southeast: Georgia Institute of Technology Southwest: University of Texas at Austin Western: University of Colorado, Denver | Anschutz Medical Campus Northwest: University of Washington

REGISTER TODAY!

Q4 2023 Forum Cover

  • Publications

2023 Q4 Biomaterials Forum is Available Now!

The  Biomaterials Forum  Fourth Quarter 2023 issue is now available online. The feature of this issue is a remembrance of Dr. Allan S. Hoffman. It also includes a letter from our president, William R. Wagner, on the role of biomaterials scientists in the evaluation of risk associated with PFAS. Read up on this and more!

Journal of Biomedical Materials Research

  • Fabrication of anodic and atomic layer deposition‐alumina coated titanium implants for effective osteointegration applications Submitted on Fri, 09/06/2024 Journal of Biomedical Materials Research Part A, EarlyView.
  • Collagen binding and mimetic peptide‐functionalized self‐assembled peptide hydrogel enhance chondrogenic differentiation of human mesenchymal stem cells Submitted on Fri, 09/06/2024 Journal of Biomedical Materials Research Part A, EarlyView.

Applied Biomaterials

  • The Effect of Cigarettes Smoke on the Color and Properties of a Silicone for Maxillofacial Prostheses Submitted on Wed, 09/04/2024 Journal of Biomedical Materials Research Part B: Applied Biomaterials, Volume 112, Issue 9, September 2024.
  • Composite suture material based on polylactide accelerates the healing of surgical wounds in in vivo experiment Submitted on Tue, 09/03/2024 Journal of Biomedical Materials Research Part B: Applied Biomaterials, Volume 112, Issue 9, September 2024.

The Materials Science & Technology (MS&T) technical meeting and exhibition is the long-standing, recognized forum for fostering technical innovation at the intersection of materials science, engineering, and application. Each year, MS&T brings together scientists, engineers, students, suppliers, and business leaders to discuss current research and technical applications and to shape the future of materials science and technology. The event's unmatched technical program addresses structure, properties, processing, and performance across the materials community.

MS&T24

Committees & SIGs

Biomaterials researchers and engineers just like yourself envision the future of biomaterials together as volunteer leaders of the Society For Biomaterials. This is YOUR society. Get involved today to share in networking opportunities, build bridges to other scientific groups, and explore the diverse biomaterials community.

Membership Benefits

SFB offers these exciting membership Benefits!

  • Special Interest Groups
  • Professional Development
  • Annual Meetings
  • Student Opportunities
  • Volunteer Opportunities

View All News

SFB Jobs

Employment Opportunities

Members of the Society For Biomaterials may post current job openings. Please email a single page PDF of Word DOC to SFB headquarters . All job postings will be listed on this page for one month, free of charge.

SFB Jobs

Advertisement

Back to Top

W

  • General & Introductory Life Sciences
  • Neuroscience

journal of biomedical materials research

Journal of Biomedical Materials Research Part B: Applied Biomaterials

Print ISSN: 1552-4973

Online ISSN: 1552-4981

Impact Factor: 3.2

Society for Biomaterials

journal of biomedical materials research

Edited By:Jeremy Gilbert

U.S. flag

An official website of the United States government

The .gov means it’s official. Federal government websites often end in .gov or .mil. Before sharing sensitive information, make sure you’re on a federal government site.

The site is secure. The https:// ensures that you are connecting to the official website and that any information you provide is encrypted and transmitted securely.

  • Publications
  • Account settings
  • My Bibliography
  • Collections
  • Citation manager

Save citation to file

Email citation, add to collections.

  • Create a new collection
  • Add to an existing collection

Add to My Bibliography

Your saved search, create a file for external citation management software, your rss feed.

  • Search in PubMed
  • Search in NLM Catalog
  • Add to Search

Human Amniotic Membrane: A review on tissue engineering, application, and storage

Affiliations.

  • 1 Institute for Multiphase Processes, Leibniz University Hannover, Garbsen, Germany.
  • 2 Department of Ophthalmology, University Eye Hospital, Hannover Medical School, Hannover, Germany.
  • 3 German Society for Tissue Transplantation (DGFG), Hannover, Germany.
  • 4 Institute of Transfusion Medicine and Transplant Engineering, Hannover Medical School, Hannover, Germany.
  • PMID: 33319484
  • DOI: 10.1002/jbm.b.34782

Human amniotic membrane (hAM) has been employed as scaffolding material in a wide range of tissue engineering applications, especially as a skin dressing and as a graft for corneal treatment, due to the structure of the extracellular matrix and excellent biological properties that enhance both wound healing and tissue regeneration. This review highlights recent work and current knowledge on the application of native hAM, and/or production of hAM-based tissue-engineered products to create scaffolds mimicking the structure of the native membrane to enhance the hAM performance. Moreover, an overview is presented on the available (cryo) preservation techniques for storage of native hAM and tissue-engineered products that are necessary to maintain biological functions such as angiogenesis, anti-inflammation, antifibrotic and antibacterial activity.

Keywords: cryopreservation; cryoprotective agent (CPA); extracellular matrix (ECM); human amniotic membrane (hAM); scaffold; tissue engineering (TE).

© 2020 The Authors. Journal of Biomedical Materials Research Part B: Applied Biomaterials published by Wiley Periodicals LLC.

PubMed Disclaimer

Similar articles

  • Amniotic miracle: Investigating the unique development and applications of amniotic membrane in wound healing. Guo X, Zhang W, Lu J, Zhu Y, Sun H, Xu D, Xian S, Yao Y, Qian W, Lu B, Shi J, Ding X, Li Y, Tong X, Xiao S, Huang R, Ji S. Guo X, et al. Skin Res Technol. 2024 Jul;30(7):e13860. doi: 10.1111/srt.13860. Skin Res Technol. 2024. PMID: 39073182 Free PMC article. Review.
  • Human Amniotic Membrane with Aligned Electrospun Fiber as Scaffold for Aligned Tissue Regeneration. Hasmad H, Yusof MR, Mohd Razi ZR, Hj Idrus RB, Chowdhury SR. Hasmad H, et al. Tissue Eng Part C Methods. 2018 Jun;24(6):368-378. doi: 10.1089/ten.TEC.2017.0447. Epub 2018 May 24. Tissue Eng Part C Methods. 2018. PMID: 29690856
  • Decellularized human amniotic membrane: From animal models to clinical trials. Milan PB, Amini N, Joghataei MT, Ebrahimi L, Amoupour M, Sarveazad A, Kargozar S, Mozafari M. Milan PB, et al. Methods. 2020 Jan 15;171:11-19. doi: 10.1016/j.ymeth.2019.07.018. Epub 2019 Jul 19. Methods. 2020. PMID: 31326597
  • Protocols for decellularization of human amniotic membrane. Khosravimelal S, Momeni M, Gholipur M, Kundu SC, Gholipourmalekabadi M. Khosravimelal S, et al. Methods Cell Biol. 2020;157:37-47. doi: 10.1016/bs.mcb.2019.11.004. Epub 2019 Dec 18. Methods Cell Biol. 2020. PMID: 32334719
  • Application of conducting polymers to wound care and skin tissue engineering: A review. Talikowska M, Fu X, Lisak G. Talikowska M, et al. Biosens Bioelectron. 2019 Jun 15;135:50-63. doi: 10.1016/j.bios.2019.04.001. Epub 2019 Apr 6. Biosens Bioelectron. 2019. PMID: 30999241 Review.
  • Focal Adhesion Maturation Responsible for Behavioral Changes in Human Corneal Stromal Fibroblasts on Fibrillar Substrates. Zhurenkov KE, Lobov AA, Bildyug NB, Alexander-Sinclair EI, Darvish DM, Lomert EV, Kriger DV, Zainullina BR, Chabina AS, Khorolskaya JI, Perepletchikova DA, Blinova MI, Mikhailova NA. Zhurenkov KE, et al. Int J Mol Sci. 2024 Aug 7;25(16):8601. doi: 10.3390/ijms25168601. Int J Mol Sci. 2024. PMID: 39201288 Free PMC article.
  • Pathological and pharmacovigilance monitoring as toxicological imputations of azithromycin and its residues in broilers. Fotouh A, Abdel-Maguid DS, Abdelhaseib M, Zaki RS, Darweish M. Fotouh A, et al. Vet World. 2024 Jun;17(6):1271-1280. doi: 10.14202/vetworld.2024.1271-1280. Epub 2024 Jun 14. Vet World. 2024. PMID: 39077436 Free PMC article.
  • Human Amniotic Membrane for Dural Repair and Duraplasty: A Systematic Review of Safety and Efficacy. Abbas A, Hamad AA, Ballut OO, El-Gayar RM, Negida A, Raslan AM. Abbas A, et al. Cureus. 2023 Dec 26;15(12):e51117. doi: 10.7759/cureus.51117. eCollection 2023 Dec. Cureus. 2023. PMID: 38274915 Free PMC article. Review.
  • Modification of the Human Amniotic Membrane Using Different Cross-Linking Agents as a Promising Tool for Regenerative Medicine. Skopinska-Wisniewska J, Michalak M, Tworkiewicz J, Tyloch D, Tuszynska M, Bajek A. Skopinska-Wisniewska J, et al. Materials (Basel). 2023 Oct 17;16(20):6726. doi: 10.3390/ma16206726. Materials (Basel). 2023. PMID: 37895710 Free PMC article.

REFERENCESDoh

  • Grey C. Tissue engineering Scaffold Fabrication and Processing Techniques to Improve Cellular Infiltration. Virginia, USA: Virginia Commonwealth University; 2014.
  • Rana D, Arulkumar S, Vishwakarma A, Ramalingam M. Considerations on designing scaffold for tissue engineering. Stem Cell Biology and Tissue Engineering in Dental Sciences. USA: Academic Press; 2015;133-148. https://doi.org/10.1016/B978-0-12-397157-9.00012-6 .
  • Ngadiman NHA, Noordin MY, Idris A, Kurniawan D. A review of evolution of electrospun tissue engineering scaffold: from two dimensions to three dimensions. Proc Inst Mech Eng Part H J Eng Med. 2017;231(7):597-616. https://doi.org/10.1177/0954411917699021 .
  • 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(6):679-689. https://doi.org/10.1089/107632701753337645 .
  • Lim R. Concise review: fetal membranes in regenerative medicine: new tricks from an old dog? Stem Cells Transl Med. 2017;6(9):1767-1776. https://doi.org/10.1002/sctm.16-0447 .

Publication types

  • Search in MeSH

LinkOut - more resources

Full text sources, other literature sources.

  • The Lens - Patent Citations

full text provider logo

  • Citation Manager

NCBI Literature Resources

MeSH PMC Bookshelf Disclaimer

The PubMed wordmark and PubMed logo are registered trademarks of the U.S. Department of Health and Human Services (HHS). Unauthorized use of these marks is strictly prohibited.

Accessibility Links

  • Skip to content
  • Skip to search IOPscience
  • Skip to Journals list
  • Accessibility help

Biomedical Materials

Purpose-led Publishing is a coalition of three not-for-profit publishers in the field of physical sciences: AIP Publishing, the American Physical Society and IOP Publishing.

Together, as publishers that will always put purpose above profit, we have defined a set of industry standards that underpin high-quality, ethical scholarly communications.

We are proudly declaring that science is our only shareholder.

journal of biomedical materials research

Biomedical Materials publishes original research findings and critical reviews that contribute to our knowledge about the composition, properties, and performance of materials for all applications relevant to human healthcare.

Open all abstracts , in this tab

Tiancheng Li et al 2023 Biomed. Mater. 18 052004

Alveolar bone loss is widespread in all age groups and remains a severe hazard to periodontal health. Horizontal alveolar bone loss is the pattern of bone loss more commonly seen in periodontitis. Until now, limited regenerative procedures have been applied to treating horizontal alveolar bone loss in periodontal clinics, making it the least predictable periodontal defect type. This article reviews the literature on recent advances in horizontal alveolar bone regeneration. The biomaterials and clinical and preclinical approaches tested for the regeneration of the horizontal type of alveolar bone are first discussed. Furthermore, current obstacles for horizontal alveolar bone regeneration and future directions in regenerative therapy are presented to provide new ideas for developing an effective multidisciplinary strategy to address the challenge of horizontal alveolar bone loss.

Farah N S Raja et al 2023 Biomed. Mater. 18 045003

With the advent of nanotechnology, there has been an extensive interest in the antimicrobial potential of metals. The rapid and widespread development of antimicrobial-resistant and multidrug-resistant bacteria has prompted recent research into developing novel or alternative antimicrobial agents. In this study, the antimicrobial efficacy of metallic copper, cobalt, silver and zinc nanoparticles was assessed against Escherichia coli (NCTC 10538), S. aureus (ATCC 6538) along with three clinical isolates of Staphylococcus epidermidis (A37, A57 and A91) and three clinical isolates of E. coli (Strains 1, 2 and 3) recovered from bone marrow transplant patients and patients with cystitis respectively. Antimicrobial sensitivity assays, including agar diffusion and broth macro-dilution to determine minimum inhibitory and bactericidal concentrations (MIC/MBC) and time-kill/synergy assays, were used to assess the antimicrobial efficacy of the agents. The panel of test microorganisms, including antibiotic-resistant strains, demonstrated a broad range of sensitivity to the metals investigated. MICs of the type culture strains were in the range of 0.625–5.0 mg ml −1 . While copper and cobalt exhibited no difference in sensitivity between Gram-positive and Gram-negative microorganisms, silver and zinc showed strain specificity. A significant decrease ( p < 0.001) in the bacterial density of E. coli and S. aureus was demonstrated by silver, copper and zinc in as little as two hours. Furthermore, combining metal nanoparticles reduced the time required to achieve a complete kill.

Hanne Meryem Kayadurmus et al 2024 Biomed. Mater. 19 045045

Chronic skin wounds pose a global clinical challenge, necessitating effective treatment strategies. This study explores the potential of 3D printed Poly Lactic Acid (PLA) scaffolds, enhanced with Whey Protein Concentrate (WPC) at varying concentrations (25, 35, and 50% wt), for wound healing applications. PLA's biocompatibility, biodegradability, and thermal stability make it an ideal material for medical applications. The addition of WPC aims to mimic the skin's extracellular matrix and enhance the bioactivity of the PLA scaffolds. Fourier Transform Infrared Spectroscopy results confirmed the successful loading of WPC into the 3D printed PLA-based scaffolds. Scanning Electron Microscopy (SEM) images revealed no significant differences in pore size between PLA/WPC scaffolds and pure PLA scaffolds. Mechanical strength tests showed similar tensile strength between pure PLA and PLA with 50% WPC scaffolds. However, scaffolds with lower WPC concentrations displayed reduced tensile strength. Notably, all PLA/WPC scaffolds exhibited increased strain at break compared to pure PLA. Swelling capacity was highest in PLA with 25% WPC, approximately 130% higher than pure PLA. Scaffolds with higher WPC concentrations also showed increased swelling and degradation rates. Drug release was found to be prolonged with increasing WPC concentration. After seven days of incubation, cell viability significantly increased in PLA with 50% WPC scaffolds compared to pure PLA scaffolds. This innovative approach could pave the way for personalized wound care strategies, offering tailored treatments and targeted drug delivery. However, further studies are needed to optimize the properties of these scaffolds and validate their effectiveness in clinical settings.

Maryke de Villiers et al 2024 Biomed. Mater. 19 055034

The high incidence of malignant melanoma highlights the need for in vitro models that accurately represent the tumour microenvironment, enabling developments in melanoma therapy and drug screening. Despite several advancements in 3D cell culture models, appropriate melanoma models for evaluating drug efficacy are still in high demand. The 3D pneumatic extrusion-based bioprinting technology offers numerous benefits, including the ability to achieve high-throughput capabilities. However, there is a lack of research that combines pneumatic extrusion-based bioprinting with analytical assays to enable efficient drug screening in 3D melanoma models. To address this gap, this study developed a simple and highly reproducible approach to fabricate a 3D A375 melanoma cell culture model using the pneumatic extrusion-based bioprinting technology. To optimise this method, the bioprinting parameters for producing 3D cell cultures in a 96-well plate were adjusted to improve reproducibility while maintaining the desired droplet size and a cell viability of 92.13 ± 6.02%. The cross-linking method was optimised by evaluating cell viability and proliferation of the 3D bioprinted cells in three different concentrations of calcium chloride. The lower concentration of 50 mM resulted in higher cell viability and increased cell proliferation after 9 d of incubation. The A375 cells exhibited a steadier proliferation rate in the 3D bioprinted cell cultures, and tended to aggregate into spheroids, whereas the 2D cell cultures generally formed monolayered cell sheets. In addition, we evaluated the drug responses of four different anti-cancer drugs on the A375 cells in both the 2D and 3D cell cultures. The 3D cell cultures exhibited higher levels of drug resistance in all four tested anti-cancer drugs. This method presents a simple and cost-effective method of producing and analysing 3D cell culture models that do not add additional complexity to current assays and shows considerable potential for advancing 3D cell culture models' drug efficacy evaluations.

Bashiru K Sodipo and Zainab Kasim Mohammed 2024 Biomed. Mater. 19 042010

To improve the translational and clinical applications of gold nanoparticles (GNPs) in medicine there is a need for better understanding of physicochemical properties of the nanoparticles in relation to the systemic parameters and in-vivo performance. This review presents the influence of physicochemical properties (surface charges and size) and route of administration on the biodistribution of GNPs. The role of protein corona (PC) (a unique biological identifier) as a barrier to biodistribution of GNPs, and the advances in engineered GNPs towards improving biodistribution are presented. Proteins can easily adsorb on charged (anionic and cationic) functionalized GNPs in circulation and shape the dynamics of their biodistribution. Non-ionic coatings such as PEG experience accelerated blood clearance (ABC) due to immunogenic response. While zwitterionic coatings provide stealth effects to formation of PC on the GNPs. GNPs with sizes less than 50 nm were found to circulate to several organs while the route of administration of the GNPs determines the serum protein that adsorbs on the nanoparticles.

Hilal Yilmaz et al 2024 Biomed. Mater. 19 045029

Although different fabrication methods and biomaterials are used in scaffold development, hydrogels and electrospun materials that provide the closest environment to the extracellular matrix have recently attracted considerable interest in tissue engineering applications. However, some of the limitations encountered in the application of these methods alone in scaffold fabrication have increased the tendency to use these methods together. In this study, a bilayer scaffold was developed using 3D-printed gelatin methacryloyl (GelMA) hydrogel containing ciprofloxacin (CIP) and electrospun polycaprolactone (PCL)-collagen (COL) patches. The bilayer scaffolds were characterized in terms of chemical, morphological, mechanical, swelling, and degradation properties; drug release, antibacterial properties, and cytocompatibility of the scaffolds were also studied. In conclusion, bilayer GelMA-CIP/PCL-COL scaffolds, which exhibit sufficient porosity, mechanical strength, and antibacterial properties and also support cell growth, are promising potential substitutes in tissue engineering applications.

Chenlu Li et al 2024 Biomed. Mater. 19 045039

Recently, cytokine-induced killer (CIK) cells have a broad application prospect in the comprehensive diagnosis and treatment of tumors owing to their unique characteristics of killing and targeting malignant tumors. Herein, we report a facile strategy for synthesis of monodisperse gold nanostars (GNSs) based on PEGylation and co-loaded with the photosensitizer chlorin e6 (Ce6) to form GNSs-PEG@Ce6 NPs. Then employing CIK cells loading the as-prepared GNSs-PEG@Ce6 NPs to fabricate a CIK cells-based drug delivery system (GNSs-PEG@Ce6-CIK) for lung cancer. Among them, GNSs was functioned as transport media, Ce6 acted as the near-infrared (NIR) fluorescence imaging agent and photodynamic therapy (PDT), and CIK cells served as targeting vectors for immunotherapy, which can increase the efficiency of tumor enrichment and treatment effect. The results of cellular experiments demonstrated that GNSs-PEG@Ce6 NPs had good dispersibility, water solubility and low toxicity under physiological conditions, and the cultured CIK cells had strong anti-tumor properties. Subsequently, GNSs-PEG@Ce6-CIK could effectively inhibit the growth of A549 cells under the exposure of 633 nm laser, which showed stronger killing effect than that of GNSs-PEG@Ce6 NPs or CIK cells. In addition, they showed good tumor targeting and tumor synergistic killing activity in vivo . Therefore, GNSs-PEG@Ce6-CIK was constructed for targeted NIR fluorescence imaging, enhanced PDT and immunotherapy of lung cancer.

Rosemond A Mensah et al 2023 Biomed. Mater. 18 042001

Naturally derived materials are often preferred over synthetic materials for biomedical applications due to their innate biological characteristics, relative availability, sustainability, and agreement with conscientious end-users. The chicken eggshell membrane (ESM) is an abundant resource with a defined structural profile, chemical composition, and validated morphological and mechanical characteristics. These unique properties have not only allowed the ESM to be exploited within the food industry but has also led to it be considered for other novel translational applications such as tissue regeneration and replacement, wound healing and drug delivery. However, challenges still exist in order to enhance the native ESM (nESM): the need to improve its mechanical properties, the ability to combine/join fragments of ESM together, and the addition or incorporation of drugs/growth factors to advance its therapeutic capacity. This review article provides a succinct background to the nESM, its extraction, isolation, and consequent physical, mechanical and biological characterisation including possible approaches to enhancement. Moreover, it also highlights current applications of the ESM in regenerative medicine and hints at future novel applications in which this novel biomaterial could be exploited to beneficial use.

Katherine Pitrolino et al 2024 Biomed. Mater. 19 055025

Bioresorbable chitosan scaffolds have shown potential for osteochondral repair applications. The in vivo degradation of chitosan, mediated by lysozyme and releasing glucosamine, enables progressive replacement by ingrowing tissue. Here the degradation process of a chitosan-nHA based bioresorbable scaffold was investigated for mass loss, mechanical properties and degradation products released from the scaffold when subjected to clinically relevant enzyme concentrations. The scaffold showed accelerated mass loss during the early stages of degradation but without substantial reduction in mechanical strength or structure deterioration. Although not cytotoxic, the medium in which the scaffold was degraded for over 2 weeks showed a transient decrease in mesenchymal stem cell viability, and the main degradation product (glucosamine) demonstrated a possible adverse effect on viability when added at its peak concentration. This study has implications for the design and biomedical application of chitosan scaffolds, underlining the importance of modelling degradation products to determine suitability for clinical translation.

Hideo Saotome et al 2024 Biomed. Mater. 19 045031

The alignment of each cell in human myocardium is considered critical for the efficient movement of cardiac tissue. We investigated 96-well microstripe-patterned plates to align human induced pluripotent stem cell (hiPSC)-derived cardiomyocytes (CMs), which resemble fetal myocardium. The aligned CMs (ACMs) cultured on the microstripe-patterned plates exhibited pathology, motor function, gene expression, and drug response that more closely resembled those of adult cells than did unaligned CMs cultured on a flat plate (FCMs). We used these ACMs to evaluate drug side effects and efficacy, and to determine whether these were similar to adult-like responses. When CMs from patients with hypertrophic cardiomyopathy (HCMs) were seeded and cultured on the microstripe-patterned plates or layered on top of the ACMs, both sets of HCMs showed increased heart rate and synchronized contractions, indicating improved cardiac function. It is suggested that the ACMs could be used for drug screening as cells representative of adult-like CMs and be transplanted in the form of a cell sheet for regenerative treatment of heart failure.

Latest articles

Qi-yan Li et al 2024 Biomed. Mater. 19 065002

The invasion and metastasis of tumors pose significant challenges in the treatment of ovarian cancer (OC), making it difficult to cure. One potential treatment approach that has gained attention is the use of matrix metalloproteinase reactive controlled release micelle preparations. In this study, we developed a novel PEG 5000 -PVGLIG-hyaluronic acid docetaxel/bakuchiol (PP-HA-DTX/BAK) micelles formulation with desirable characteristics such as particle size, narrow polydispersity index, and a ZETA potential of approximately −5 mV. The surface modification with HA facilitates tumor penetration into the tumor interior, while the incorporation of DSPE-PEG 2000 -PVGLIG-PEG 5000 helps conceal DSPE-PEG 2000 -HA, reducing off-target effects and prolonging drug circulation time in vivo . Both in vitro and in vivo experiments demonstrated that these micelles effectively inhibit proliferation, invasion, and metastasis of OC cells while promoting apoptosis. Therefore, our findings suggest that PP-HA-DTX/BAK micelles represent a safe and effective therapeutic strategy for treating OC.

Irem Unalan et al 2024 Biomed. Mater. 19 065001

Electrospinning is a versatile and straightforward technique to produce nanofibrous mats with different morphologies. In addition, by optimizing the solution, processing, and environmental parameters, three-dimensional (3D) nanofibrous scaffolds can also be created using this method. In this work, the preparation and characterization of bioactive glass (BG) scaffolds based on the SiO 2 –CaO sol–gel system, a biomaterial with a highly reactive surface, is reported. The electrospinning technique was combined with sol–gel methods to obtain nanofibrous 3D cotton wool-like scaffolds. The addition of zinc and copper ions to the silica-calcia system was examined, and the influence of these ions on the material properties and characteristics was investigated by various characterization techniques, from morphological and chemical properties to antibacterial and wound closure capability, cell viability and ion release. Our findings show that the cotton wool-like ion-doped nanofibers are promising for wound healing applications.

Jovana Zvicer et al 2024 Biomed. Mater. 19 055044

In tissue engineering, collaboration among experts from different fields is needed to design appropriate cell scaffolds and the required three-dimensional environment. Osteochondral tissue engineering is particularly challenging due to the need to provide scaffolds that imitate structural and compositional differences between two neighboring tissues, articular cartilage and bone, and the required complex biophysical environments for cultivating such scaffolds. This work focuses on two key objectives: first, to develop bilayered osteochondral scaffolds based on gellan gum and bioactive glass and, second, to create a biomimetic environment for scaffold characterization by designing and utilizing novel dual-medium cultivation bioreactor chambers. Basic chemical engineering principles were utilized to help achieve both aims. First, a simple heat transport model based on one-dimensional conduction was applied as a guideline for bilayer scaffold preparation, leading to the formation of a gelatinous upper part and a macroporous lower part with a thin, well-integrated interfacial zone. Second, a novel cultivation chamber was developed to be used in a dynamic compression bioreactor to provide possibilities for flow of two different media, such as chondrogenic and osteogenic. These chambers were utilized for characterization of the novel scaffolds with regard to bioactivity and stability under dynamic compression and fluid perfusion over 14 d, while flow distribution under different conditions was analyzed by a tracer method and residence time distribution analysis.

Andrew Padalhin et al 2024 Biomed. Mater. 19 055042

During the healing process after intra-nasal surgery, the growth and repair of damaged tissues can result in the development of postoperative adhesions. Various techniques have been devised to minimize the occurrence of postoperative adhesions which include insertion of stents in the middle meatus, application of removable nasal packing, and utilizing biodegradable materials with antiadhesive properties. This study assesses the efficacy of two sodium hyaluronate (SH)–based freeze-dried hydrogel composites in preventing postoperative nasal adhesions, comparing them with commonly used biodegradable materials in nasal surgery. The freeze-dried hydrogels, sodium hyaluronate and collagen 1(SH-COL1) and sodium hyaluronate, carboxymethyl cellulose, and collagen 1 (SH-CMC-COL1), were evaluated for their ability to reduce bleeding time, promote wound healing, and minimize fibrous tissue formation. Results showed that SH-CMC-COL1 significantly reduced bleeding time compared to both biodegradable polyurethane foam and SH-COL1. Both SH-COL1 and SH-CMC-COL1 exhibited enhanced wound healing effects, as indicated by significantly greater wound size reduction after two weeks compared to the control. Histological analyses revealed significant differences in re-epithelialization and blood vessel count among all tested materials, suggesting variable initial wound tissue response. Although all treatment groups had more epithelial growth, with X-SCC having higher blood vessel count at 7 d post treatment, all treatment groups did not differ in all histomorphometric parameters by day 14. However, the long-term application of SH-COL1 demonstrated a notable advantage in reducing nasal adhesion formation compared to all other tested materials. This indicates the potential of SH–based hydrogels, particularly SH-COL1, in mitigating postoperative complications associated with nasal surgery. These findings underscore the versatility and efficacy of SH–based freeze-dried hydrogel composites for the management of short-term and long-term nasal bleeding with an anti-adhesion effect. Further research is warranted to optimize their clinical use, particularly in understanding the inflammatory factors influencing tissue adhesions and assessing material performance under conditions mimicking clinical settings. Such insights will be crucial for refining therapeutic approaches and optimizing biomaterial design, ultimately improving patient outcomes in nasal surgery.

Yan Liu et al 2024 Biomed. Mater. 19 052009

Curcumin is a natural polyphenolic compound extracted from turmeric with antibacterial, antioxidant, antitumor, preventive and therapeutic neurological disorders and a variety of bioactivities, which is widely used in the field of food and medicine. However, the drawbacks of curcumin such as poor aqueous solubility and stability have limited the practical application of curcumin. To overcome these defects and enhance its functional properties, various nanoscale systems (liposomes, polymer nanoparticles, protein nanoparticles, solid lipid nanoparticles, metal nanoparticles, etc) have been extensively employed for curcumin encapsulation and delivery. Despite the rapid development of curcumin nanoformulations, there is a lack of comprehensive reviews on their preparation and properties. This review provides an overview of the construction of curcumin nano-delivery systems, mechanisms of action, nanocarrier preparation methods and the applications of curcumin nanocarriers in the food and pharmaceutical fields to provide a theoretical basis and technological support for the efficient bio-utilization, product development and early clinical application of curcumin.

Review articles

Rosalie Bordett et al 2024 Biomed. Mater. 19 052008

Soft-tissue injuries affecting muscles, nerves, vasculature, tendons, and ligaments often diminish the quality of life due to pain, loss of function, and financial burdens. Both natural healing and surgical interventions can result in scarring, which potentially may impede functional recovery and lead to persistent pain. Scar tissue, characterized by a highly disorganized fibrotic extracellular matrix, may serve as a physical barrier to regeneration and drug delivery. While approaches such as drugs, biomaterials, cells, external stimulation, and other physical forces show promise in mitigating scarring and promoting regenerative healing, their implementation remains limited and challenging. Ultrasound, laser, electrical, and magnetic forms of external stimulation have been utilized to promote soft tissue as well as neural tissue regeneration. After stimulation, neural tissues experience increased proliferation of Schwann cells, secretion of neurotropic factors, production of myelin, and growth of vasculature, all aimed at supporting axon regeneration and innervation. Yet, the outcomes of healing vary depending on the pathophysiology of the damaged nerve, the timing of stimulation following injury, and the specific parameters of stimulation employed. Increased treatment intensity and duration have been noted to hinder the healing process by inducing tissue damage. These stimulation modalities, either alone or in combination with nerve guidance conduits and scaffolds, have been demonstrated to promote healing. However, the literature currently lacks a detailed understanding of the stimulation parameters used for nerve healing applications. In this article, we aim to address this gap by summarizing existing reports and providing an overview of stimulation parameters alongside their associated healing outcomes.

Yinan Wang et al 2024 Biomed. Mater. 19 052007

Melanoma is a common malignant tumor, with a five-year mortality rate as high as 62% in cases of metastatic melanoma according to cancer statistics (2024). In recent years, the focus of melanoma research has predominantly centered on immunotherapy and targeted therapy, grappling with challenges such as resistance and immunogenicity. The discovery of nanoparticles (NPs) has brought nano-delivery systems to the forefront of melanoma diagnosis and treatment. Although certain NPs, like liposomes, have gained clinical approval, utilizing most nano-delivery systems for melanoma diagnosis and treatment remains largely exploratory. The inherent limitations of NPs present a major obstacle to their clinical translation. By selecting suitable nanocarriers and functionalizing NPs to optimize nano-delivery systems, and combining these systems with other therapies, it is possible to reduce the systemic toxicity and resistance associated with conventional therapies and the NPs themselves. This optimization could significantly improve the effectiveness of nano-delivery systems in the early detection and timely treatment of melanoma. However, there have been few reviews on the optimization of NPs and the combined application of other therapies in the treatment and diagnostic application of melanoma in the past three years. This review summarizes the latest applications of nano-delivery systems in the diagnosis and treatment of melanoma over the past three years, including innovations and achievements in both preclinical and clinical studies, offering new perspectives on their potential and future application prospects. It integrates clinical data and patent information, highlights trends in nano-delivery system development, and offers new insights into their clinical translation. Additionally, it discusses the challenges and opportunities of nano-delivery systems in melanoma treatment, providing a foundation for advancing their application in diagnosis, treatment, and clinical translation.

Sheqing Chen et al 2024 Biomed. Mater. 19 052006

Osteonecrosis of the femoral head (ONFH) is a refractory disease affecting young adults, resulting in severe hip pain, femoral head collapse, and disabling dysfunction. By far, the underlying mechanism of its pathology is unclear, and still lack of a mature and effective treatment. Exosomes, a regulator of cell–cell communication, their cargos may vary in response to different physiological or pathological conditions. To date, many studies have demonstrated that exosomes have the potential to become a diagnostic marker and therapeutic agent in many human diseases including ONFH. As a cell-free therapeutic agent, exosomes are becoming a promising tool within this field due to their crucial role in osteogenesis and angiogenesis in recent decades. Usually, exosomes from ONFH tissues could promote ONFH damage, while stem cells derived exosomes could delay diseases and repair femoral head necrosis. Herein, we describe the properties of exosomes, discuss its effect on pathogenesis, diagnosis, and treatment potential in ONFH, and examine the involvement of different signaling pathways. We also propose our suggestions for the future research of exosomes in ONFH field and hope to provide a potential therapeutic strategy for patients with ONFH.

Muhammad Umar Aslam Khan et al 2024 Biomed. Mater. 19 052005

Bone is a dynamic tissue that can always regenerate itself through remodeling to maintain biofunctionality. This tissue performs several vital physiological functions. However, bone scaffolds are required for critical-size damages and fractures, and these can be addressed by bone tissue engineering. Bone tissue engineering (BTE) has the potential to develop scaffolds for repairing critical-size damaged bone. BTE is a multidisciplinary engineered scaffold with the desired properties for repairing damaged bone tissue. Herein, we have provided an overview of the common carbohydrate polymers, fundamental structural, physicochemical, and biological properties, and fabrication techniques for bone tissue engineering. We also discussed advanced biofabrication strategies and provided the limitations and prospects by highlighting significant issues in bone tissue engineering. There are several review articles available on bone tissue engineering. However, we have provided a state-of-the-art review article that discussed recent progress and trends within the last 3–5 years by emphasizing challenges and future perspectives.

Accepted manuscripts

Hanaei et al 

Osteosarcoma is the mostly commonly occurring primary bone cancer. Despite comprehensive treatment programs including neoadjuvant chemotherapy and tumour resection, survival rates have not improved significantly since the 1970s. Survival rates are dramatically reduced for patients who suffer a local recurrence. Furthermore, primary bone cancer patients are at increased risk of bone fractures. Consequently, there is an urgent need for alternative treatment options. In this paper we report the development of novel gallium doped bioactive glass that selectively kill bone cancer cells whilst simultaneously stimulating new bone growth. Here we show, using a combination of MTT, LIVE/DEAD assays and image analysis, that bioactive glasses containing gallium oxide are highly toxic and reduce both the proliferation and migration of bone cancer cells (Saos-2) in a dose dependant manner. Glasses containing 5 mol% gallium oxide reduced the viability of osteosarcoma cells by 99% without being cytotoxic to the non-cancerous normal human osteoblasts (NHOst) control cells. Furthermore, FTIR and Energy-dispersive X-ray spectroscopy results confirmed the formation of an amorphous calcium phosphate / hydroxy apatite layer on the surface of the bioactive glass particulates, after 7 days incubating in simulated body fluid, indicating the early stages of bone formation. These materials show significant potential for use in bone cancer applications as part of a multimodal treatment.&#xD;

Xing et al 

Traditional cell culture does not accurately simulate the cell microenvironment and demonstrate the specificity of individual cell growth. In this study, we created a 4D cell culture model. It is a precision instrument consisting of an electromagnet, a force transducer, and a cantilever bracket. A petri dish is placed above the magnet, gel beads encapsulated with magnetic nanoparticles and tongue cancer cells are placed in the petri dish. In this model, a magnetic force is generated on the magnetic nanoparticles in the culture medium to drive gel movement when the magnet is energized, and the gel exerts an external force on the cells. It can mimic the microenvironment inside the tongue squamous cell carcinoma cell CAL-27 when the tongue is moving. Electron microscopy and rheological analysis experiments were performed on hydrogels to characterize the alginate. Calcein-AM/PI staining was conducted to verify the biosafety of the hydrogel culture system. Four experimental groups were set up in this study. On this basis, we confirmed that proliferation of tongue squamous cell CAL-27 significantly increased after 5 days compared to cells cultured without mechanical stimulation by MTT. The cell diameter of different groups was measured on captured photographs, and it was concluded that the cell diameter in the dynamic culture environment was larger than that of the hydrogel culture alone in the same period. Furthermore, the cell morphology of tongue squamous cell carcinoma was better. In flow cytometry experiments, there were fewer cells in the G0/G1 phase while the proportion of cells in the G2/M phase was increased following mechanical stimulation. The changes in RNA levels between groups after 21 days of culture were detected using RT-PCR. This device, mimicking the microenvironment of tongue cancer cells in vivo, can enable better visualization of the cell growth in vivo. Therefore, this study provides a reliable basis for subsequent tongue biopsy, research on the pathogenesis of tongue cancer, and drug treatment.&#xD;

Ferrari et al 

Vascular tissue engineering endeavors to design, fabricate, and validate biodegradable and bioabsorbable small-diameter vascular scaffolds engineered with bioactive molecules, capable of meeting the challenges imposed by commercial vascular prostheses. A comprehensive investigation of these engineered scaffolds in bioreactor is deemed essential as a prerequisite before any in vivo experimentation in order to get information regarding their behavior under physiological conditions and predict the biological activities they will possess. This study focuses on an innovative electrospun scaffold made of poly(caprolactone) and poly(glycerol sebacate), integrating quercetin, able to modulate inflammation, and gelatin, necessary to reduce permeability. A custom-made bioreactor was used to assess the performances of the scaffolds maintained under different pressure regimes, covering the human physiological pressure range. As results, the 3D microfibrous architecture was notably influenced by the release of bioactives, maintaining the adequate properties needed for the in vivo regeneration and scaffolds showed mechanical properties similar to human native artery. Release of gelatin was adequate to avoid blood leakage and useful to make the material porous during the testing period, whereas the amount of released quercetin was useful to counteract the post-surgery inflammation. This study showcases the successful validation of an engineered scaffold in a bioreactor, enabling to consider it as a promising candidate for vascular substitutes in in vivo applications. Our approach represents a significant leap forward in the field of vascular tissue engineering, offering a multifaceted solution to the complex challenges associated with small-diameter vascular prostheses.&#xD;

Scaccini et al 

Regenerative medicine is continuously looking for new natural biocompatible and possibly biodegradable materials, but also mechanically compliant. Chitosan is emerging as a promising FDA-approved biopolymer for tissue engineering, however, its exploitation in regenerative devices is limited by its brittleness and can be further improved, for example, by blending it with other materials or by tuning its superficial microstructure.&#xD;Here, we developed membranes made of chitosan and glycerol, by solvent casting and micropatterned them with directional geometries with different levels of axial symmetry. These membranes were characterized by light microscopy and atomic force microscopy (AFM), thermal, mechanical, and degradation assays, and also tested in vitro as scaffolds with Schwann cells.&#xD;The glycerol-blended chitosan membranes are optimized in terms of mechanical properties, and present a physiological-grade Young's modulus (≈ 0.7 MPa). The directional topographies are effective in directing cell polarization and migration and in particular are highly performant substrates for collective cell migration.&#xD;Here, we demonstrate that a combination of a soft compliant biomaterial and topographical micropatterning can improve the integration of these scaffolds with Schwann cells, which is a fundamental step in the peripheral nerve regeneration process.&#xD;

Wang et al 

Medical dressings with multifunctional properties, including potent regeneration capability and good biocompatibility, are increasingly needed in clinical practice. In this study, we reported a novel hybrid wound dressing (PCL/SerMA/DMOG) that combines electrospun PCL membranes with DMOG-loaded methacrylated sericin (SerMA) hydrogel. In such a design, DMOG molecules are released from the hybrid dressing in a sustained manner in vitro. A series of in vitro assays demonstrated that DMOG-loaded hybrid dressing has multiple biological functions, including promotion of human umbilical vein endothelial cells (HUVECs) proliferation and migration, in vitro vascularization, and the generation of intracellular NO. When applied to the cutaneous wound, the PCL/SerMA/DMOG dressing significantly accelerated wound closure and tissue regeneration by promoting angiogenesis in the wound area, collagen deposition, and cell proliferation within the wound bed. These results highlight the potential clinical application of PCL/SerMA/DMOG hybrid dressings as promising alternatives for accelerating wound healing via improved biocompatibility and angiogenesis amelioration.

More Accepted manuscripts

Open access

Shirin B Hanaei et al 2024 Biomed. Mater.

Pier Francesco Ferrari et al 2024 Biomed. Mater.

Luca Scaccini et al 2024 Biomed. Mater.

Yu Yusa et al 2024 Biomed. Mater. 19 055043

The corrosion of magnesium (Mg)-based bioabsorbable implanting devices is influenced by implantation environment which dynamically changes by biological response including wound healing. Understanding the corrosion mechanisms along the healing process is essential for the development of Mg-based devices. In this study, a hematoma model was created in a rat femur to analyze Mg corrosion with hematoma in the early stage of implantation. Pure Mg specimen (99.9%, ϕ 1.2 × 6 mm) was implanted in rat femur under either hematoma or non-hematoma conditions. After a designated period of implantation, the specimens were collected and weighed. The insoluble salts formed on the specimen surfaces were analyzed using scanning electron microscopy, energy-dispersive x-ray spectroscopy, and Raman spectroscopy on days 1, 3, and 7. The results indicate that hematomas promote Mg corrosion and change the insoluble salt precipitation. The weight loss of the hematoma group (27.31 ± 5.91 µg mm −2 ) was significantly larger than that of the non-hematoma group (14.77 ± 3.28 µg mm −2 ) on day 7. In the non-hematoma group, carbonate and phosphate were detected even on day 1, but the only latter was detected on day 7. In the hematoma group, hydroxide was detected on day 1, followed by the formation of carbonate and phosphate on days 3 and 7. The obtained results suggest the hypoxic and acidic microenvironment in hematomas accelerates the Mg corrosion immediately after implantation, and the subsequent hematoma resorption process leads to the formation of phosphate and carbonate with organic molecules. This study revealed the risk of hematomas as an acceleration factor of the corrosion of Mg-based devices leading to the early implant failure. It is important to consider this risk in the design of Mg-based devices and to optimize surgical procedures controlling hemorrhage at implantation and reducing unexpected bleeding after surgery.

Hong Chen et al 2024 Biomed. Mater. 19 055039

Guided bone regeneration (GBR) membranes play an important role in oral bone regeneration. However, enhancing their bone regeneration potential and antibacterial properties is crucial. Herein, silk fibroin (SF)/polycaprolactone (PCL) core–shell nanofibers loaded with epigallocatechin gallate (EGCG) were prepared using emulsion electrospinning. The nanofibrous membranes were characterized via scanning electron microscopy, transmission electron microscopy, Fourier transform infrared spectroscopy, thermogravimetric analysis, water contact angle (CA) measurement, mechanical properties testing, drug release kinetics, and 1,1-diphenyl-2-picryl-hydrazyl radical (DPPH) free radical scavenging assay. Mouse pre-osteoblast MC3T3-E1 cells were used to assess the biological characteristics, cytocompatibility, and osteogenic differentiation potential of the nanofibrous membrane. Additionally, the antibacterial properties against Staphylococcus aureus (S. aureus) and Escherichia coli (E. coli) were evaluated. The nanofibers prepared by emulsion electrospinning exhibited a stable core–shell structure with a smooth and continuous surface. The tensile strength of the SF/PCL membrane loaded with EGCG was 3.88 ± 0.15 Mpa, the water CA was 50°, and the DPPH clearance rate at 24 h was 81.73% ± 0.07%. The EGCG release rate of membranes prepared by emulsion electrospinning was reduced by 12% within 72 h compared to that of membranes prepared via traditional electrospinning. In vitro experiments indicate that the core–shell membranes loaded with EGCG demonstrated good cell compatibility and promoted adhesion, proliferation, and osteogenic differentiation of MC3T3-E1 cells. Furthermore, the EGCG-loaded membranes exhibited inhibitory effects on E. coli and S. aureus . These findings indicate that core–shell nanofibrous membranes encapsulated with EGCG prepared using emulsion electrospinning possess good antioxidant, osteogenic, and antibacterial properties, making them potential candidates for research in GBR materials.

Changyuan Gu et al 2024 Biomed. Mater. 19 055038

Porous poly (lactic-co-glycolic acid)/ β -tricalcium phosphate/Icaritin (PLGA/ β -TCP/ICT, PTI) scaffold is a tissue engineering scaffold based on PLGA/ β -TCP (PT) containing Icaritin, the main active ingredient of the Chinese medicine Epimedium. Due to its excellent mechanical properties and osteogenic effect, PTI scaffold has the potential to promote bone defect repair. However, the release of ICT from the scaffolds is difficult to control. In this study, we constructed Ti 3 C 2 T x @PLGA/ICT microspheres (TIM) and evaluated their characterization as well as ICT release under near-infrared (NIR) irradiation. We utilized TIM to modify the PT scaffold and performed biological experiments. First, we cultured rat bone marrow mesenchymal stem cells on the scaffold to assess biocompatibility and osteogenic potential under on-demand NIR irradiation. Subsequently, to evaluate the osteogenic properties of TIM-modified scaffold in vivo , the scaffold was implanted into a femoral condyle defect model. TIM have excellent drug-loading capacity and encapsulation efficiency for ICT, and the incorporation of Ti 3 C 2 T x endows TIM with photothermal conversion capability. Under 0.90 W cm −2 NIR irradiation, the temperature of TIM maintained at 42.0 ± 0.5 °C and the release of ICT was accelerated. Furthermore, while retaining its original properties, the TIM-modified scaffold was biocompatible and could promote cell proliferation, osteogenic differentiation, and biomineralization in vitro , as well as the osteogenesis and osseointegration in vivo , and its effect was further enhanced through the modulation of ICT release under NIR irradiation. In summary, TIM-modified scaffold has the potential to be applied in bone defects repairing.

Victor I Garcia-Perez et al 2024 Biomed. Mater. 19 055037

This study delves into the potential of amorphous titanium oxide (aTiO 2 ) nano-coating to enhance various critical aspects of non-Ti-based metallic orthopedic implants. These implants, such as medical-grade stainless steel (SS), are widely used for orthopedic devices that demand high strength and durability. The aTiO 2 nano-coating, deposited via magnetron sputtering, is a unique attempt to improve the osteogenesis, the inflammatory response, and to reduce bacterial colonization on SS substrates. The study characterized the nanocoated surfaces (SS-a TiO 2 ) in topography, roughness, wettability, and chemical composition. Comparative samples included uncoated SS and sandblasted/acid-etched Ti substrates (Ti). The biological effects were assessed using human mesenchymal stem cells (MSCs) and primary murine macrophages. Bacterial tests were carried out with two aerobic pathogens ( S. aureus and S. epidermidis ) and an anaerobic bacterial consortium representing an oral dental biofilm. Results from this study provide strong evidence of the positive effects of the aTiO 2 nano-coating on SS surfaces. The coating enhanced MSC osteoblastic differentiation and exhibited a response similar to that observed on Ti surfaces. Macrophages cultured on aTiO 2 nano-coating and Ti surfaces showed comparable anti-inflammatory phenotypes. Most significantly, a reduction in bacterial colonization across tested species was observed compared to uncoated SS substrates, further supporting the potential of aTiO 2 nano-coating in biomedical applications. The findings underscore the potential of magnetron-sputtering deposition of aTiO 2 nano-coating on non-Ti metallic surfaces such as medical-grade SS as a viable strategy to enhance osteoinductive factors and decrease pathogenic bacterial adhesion. This could significantly improve the performance of metallic-based biomedical devices beyond titanium.

More Open Access articles

Journal links

  • Submit an article
  • About the journal
  • Editorial Board
  • Author guidelines
  • Review for this journal
  • Publication charges
  • Journal collections
  • Pricing and ordering

Journal information

  • 2006-present Biomedical Materials doi: 10.1088/issn.1748-605X Online ISSN: 1748-605X Print ISSN: 1748-6041

Information

  • Author Services

Initiatives

You are accessing a machine-readable page. In order to be human-readable, please install an RSS reader.

All articles published by MDPI are made immediately available worldwide under an open access license. No special permission is required to reuse all or part of the article published by MDPI, including figures and tables. For articles published under an open access Creative Common CC BY license, any part of the article may be reused without permission provided that the original article is clearly cited. For more information, please refer to https://www.mdpi.com/openaccess .

Feature papers represent the most advanced research with significant potential for high impact in the field. A Feature Paper should be a substantial original Article that involves several techniques or approaches, provides an outlook for future research directions and describes possible research applications.

Feature papers are submitted upon individual invitation or recommendation by the scientific editors and must receive positive feedback from the reviewers.

Editor’s Choice articles are based on recommendations by the scientific editors of MDPI journals from around the world. Editors select a small number of articles recently published in the journal that they believe will be particularly interesting to readers, or important in the respective research area. The aim is to provide a snapshot of some of the most exciting work published in the various research areas of the journal.

Original Submission Date Received: .

  • Active Journals
  • Find a Journal
  • Proceedings Series
  • For Authors
  • For Reviewers
  • For Editors
  • For Librarians
  • For Publishers
  • For Societies
  • For Conference Organizers
  • Open Access Policy
  • Institutional Open Access Program
  • Special Issues Guidelines
  • Editorial Process
  • Research and Publication Ethics
  • Article Processing Charges
  • Testimonials
  • Preprints.org
  • SciProfiles
  • Encyclopedia

polymers-logo

Article Menu

journal of biomedical materials research

  • Subscribe SciFeed
  • Recommended Articles
  • Google Scholar
  • on Google Scholar
  • Table of Contents

Find support for a specific problem in the support section of our website.

Please let us know what you think of our products and services.

Visit our dedicated information section to learn more about MDPI.

JSmol Viewer

Protein immobilization on bacterial cellulose for biomedical application.

journal of biomedical materials research

1. Introduction

2. structure of bacterial cellulose, 3. biomedical application of bacterial cellulose, 3.1. wound healing and antibacterial wound dressings, 3.2. tissue engineering, 3.3. artificial blood vessels, 3.4. cell culture system, 3.5. targeted drug delivery system, 3.6. enzyme immobilization, 4. methods of protein immobilization on bacterial cellulose, 5. ex situ and in situ modifications of bacterial cellulose for protein immobilization, 5.1. ex situ bacterial cellulose modification, 5.1.1. native bacterial cellulose, 5.1.2. bacterial cellulose nanoparticles, bacterial cellulose nanofibrils, bacterial cellulose nanocrystals, 5.1.3. crosslinking, chemical crosslinking.

  • Glutaraldehyde
  • Citric acid
  • Physical crosslinking

5.1.4. Modification of the Chemical Structure of Bacterial Cellulose

Polymer grafting on bc, 5.2. in situ bacterial cellulose modification, 6. conclusions, supplementary materials, author contributions, institutional review board statement, acknowledgments, conflicts of interest.

  • Kharisov, B.I.; Kharissova, O.V.; Ortiz-Mendez, U. Drug Delivery: LyoCell® Technology—A Lipidic Drug Delivery System Based on Reverse Cubic and Hexagonal Phase Lyotropic Liquid Crystalline Nanoparticles. In CRC Concise Encyclopedia of Nanotechnology ; CRC Press: Boca Raton, FL, USA, 2016; pp. 208–212. [ Google Scholar ]
  • Maghraby, Y.R.; El-Shabasy, R.M.; Ibrahim, A.H.; Azzazy, H.M.E.-S. Enzyme Immobilization Technologies and Industrial Applications. ACS Omega 2023 , 8 , 5184–5196. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Liang, J.F.; Li, Y.T.; Yang, V.C. Biomedical Application of Immobilized Enzymes. J. Pharm. Sci. 2000 , 89 , 979–990. [ Google Scholar ] [ CrossRef ]
  • Pinmanee, P.; Sompinit, K.; Jantimaporn, A.; Khongkow, M.; Haltrich, D.; Nimchua, T.; Sukyai, P. Purification and Immobilization of Superoxide Dismutase Obtained from Saccharomyces Cerevisiae TBRC657 on Bacterial Cellulose and Its Protective Effect against Oxidative Damage in Fibroblasts. Biomolecules 2023 , 13 , 1156. [ Google Scholar ] [ CrossRef ]
  • Jesionowski, T.; Zdarta, J.; Krajewska, B. Enzyme Immobilization by Adsorption: A Review. Adsorption 2014 , 20 , 801–821. [ Google Scholar ] [ CrossRef ]
  • Klimek, K.; Ginalska, G. Proteins and Peptides as Important Modifiers of the Polymer Scaffolds for Tissue Engineering Applications—A Review. Polymers 2020 , 12 , 844. [ Google Scholar ] [ CrossRef ]
  • Isobe, N.; Lee, D.-S.; Kwon, Y.-J.; Kimura, S.; Kuga, S.; Wada, M.; Kim, U.-J. Immobilization of Protein on Cellulose Hydrogel. Cellulose 2011 , 18 , 1251–1256. [ Google Scholar ] [ CrossRef ]
  • Lyu, X.; Gonzalez, R.; Horton, A.; Li, T. Immobilization of Enzymes by Polymeric Materials. Catalysts 2021 , 11 , 1211. [ Google Scholar ] [ CrossRef ]
  • Tang, S.; Chi, K.; Xu, H.; Yong, Q.; Yang, J.; Catchmark, J.M. A Covalently Cross-Linked Hyaluronic Acid/Bacterial Cellulose Composite Hydrogel for Potential Biological Applications. Carbohydr. Polym. 2021 , 252 , 117123. [ Google Scholar ] [ CrossRef ]
  • Liu, Y.; Chen, J.Y. Enzyme Immobilization on Cellulose Matrixes. J. Bioact. Compat. Polym. 2016 , 31 , 553–567. [ Google Scholar ] [ CrossRef ]
  • Ul-Islam, M.; Khan, T.; Park, J.K. Nanoreinforced Bacterial Cellulose–Montmorillonite Composites for Biomedical Applications. Carbohydr. Polym. 2012 , 89 , 1189–1197. [ Google Scholar ] [ CrossRef ]
  • Shah, N.; Ul-Islam, M.; Khattak, W.A.; Park, J.K. Overview of Bacterial Cellulose Composites: A Multipurpose Advanced Material. Carbohydr. Polym. 2013 , 98 , 1585–1598. [ Google Scholar ] [ CrossRef ]
  • He, W.; Xu, J.; Zheng, Y.; Chen, J.; Yin, Y.; Mosselhy, D.A.; Zou, F.; Ma, M.; Liu, X. Bacterial Cellulose/Soybean Protein Isolate Composites with Promoted Inflammation Inhibition, Angiogenesis and Hair Follicle Regeneration for Wound Healing. Int. J. Biol. Macromol. 2022 , 211 , 754–766. [ Google Scholar ] [ CrossRef ]
  • Wu, S.; Wu, S.; Su, F. Novel Process for Immobilizing an Enzyme on a Bacterial Cellulose Membrane through Repeated Absorption. J. Chem. Technol. Biotechnol. 2017 , 92 , 109–114. [ Google Scholar ] [ CrossRef ]
  • Klinthoopthamrong, N.; Chaikiawkeaw, D.; Phoolcharoen, W.; Rattanapisit, K.; Kaewpungsup, P.; Pavasant, P.; Hoven, V.P. Bacterial Cellulose Membrane Conjugated with Plant-Derived Osteopontin: Preparation and Its Potential for Bone Tissue Regeneration. Int. J. Biol. Macromol. 2020 , 149 , 51–59. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Qian, L.; Yang, Y.; Xu, T.; Zhang, S.; Nica, V.; Tang, R.; Song, W. Fabrication of Efficient Protein Imprinted Materials Based on Pearl Necklace-like MOFs Bacterial Cellulose Composites. Carbohydr. Polym. 2022 , 294 , 119835. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Fu, L.; Zhang, J.; Yang, G. Present Status and Applications of Bacterial Cellulose-Based Materials for Skin Tissue Repair. Carbohydr. Polym. 2013 , 92 , 1432–1442. [ Google Scholar ] [ CrossRef ]
  • Ullah, M.W.; Manan, S.; Kiprono, S.J.; Ul-Islam, M.; Yang, G. Synthesis, Structure, and Properties of Bacterial Cellulose. In Nanocellulose ; Wiley: Hoboken, NJ, USA, 2019; pp. 81–113. [ Google Scholar ]
  • Ullah, M.W.; Ul-Islam, M.; Khan, S.; Kim, Y.; Park, J.K. Innovative Production of Bio-Cellulose Using a Cell-Free System Derived from a Single Cell Line. Carbohydr. Polym. 2015 , 132 , 286–294. [ Google Scholar ] [ CrossRef ]
  • Silva, I.G.R.d.; Pantoja, B.T.d.S.; Almeida, G.H.D.R.; Carreira, A.C.O.; Miglino, M.A. Bacterial Cellulose and ECM Hydrogels: An Innovative Approach for Cardiovascular Regenerative Medicine. Int. J. Mol. Sci. 2022 , 23 , 3955. [ Google Scholar ] [ CrossRef ]
  • Helenius, G.; Bäckdahl, H.; Bodin, A.; Nannmark, U.; Gatenholm, P.; Risberg, B. In Vivo Biocompatibility of Bacterial Cellulose. J. Biomed. Mater. Res. Part A 2006 , 76A , 431–438. [ Google Scholar ] [ CrossRef ]
  • Arikibe, J.E.; Lata, R.; Kuboyama, K.; Ougizawa, T.; Rohindra, D. PH-Responsive Studies of Bacterial Cellulose /Chitosan Hydrogels Crosslinked with Genipin: Swelling and Drug Release Behaviour. ChemistrySelect 2019 , 4 , 9915–9926. [ Google Scholar ] [ CrossRef ]
  • Lupașcu, R.E.; Ghica, M.V.; Dinu-Pîrvu, C.-E.; Popa, L.; Velescu, B.Ș.; Arsene, A.L. An Overview Regarding Microbial Aspects of Production and Applications of Bacterial Cellulose. Materials 2022 , 15 , 676. [ Google Scholar ] [ CrossRef ]
  • Choi, S.M.; Rao, K.M.; Zo, S.M.; Shin, E.J.; Han, S.S. Bacterial Cellulose and Its Applications. Polymers 2022 , 14 , 1080. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Gregory, D.A.; Tripathi, L.; Fricker, A.T.R.; Asare, E.; Orlando, I.; Raghavendran, V.; Roy, I. Bacterial Cellulose: A Smart Biomaterial with Diverse Applications. Mater. Sci. Eng. R Rep. 2021 , 145 , 100623. [ Google Scholar ] [ CrossRef ]
  • Gao, M.; Li, J.; Bao, Z.; Hu, M.; Nian, R.; Feng, D.; An, D.; Li, X.; Xian, M.; Zhang, H. A Natural in Situ Fabrication Method of Functional Bacterial Cellulose Using a Microorganism. Nat. Commun. 2019 , 10 , 437. [ Google Scholar ] [ CrossRef ]
  • Saleh, A.K.; El-Gendi, H.; El-Fakharany, E.M.; Owda, M.E.; Awad, M.A.; Kamoun, E.A. Exploitation of Cantaloupe Peels for Bacterial Cellulose Production and Functionalization with Green Synthesized Copper Oxide Nanoparticles for Diverse Biological Applications. Sci. Rep. 2022 , 12 , 19241. [ Google Scholar ] [ CrossRef ]
  • Farooq, U.; Ullah, M.W.; Yang, Q.; Aziz, A.; Xu, J.; Zhou, L.; Wang, S. High-Density Phage Particles Immobilization in Surface-Modified Bacterial Cellulose for Ultra-Sensitive and Selective Electrochemical Detection of Staphylococcus Aureus. Biosens. Bioelectron. 2020 , 157 , 112163. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Jayani, T.; Sanjeev, B.; Marimuthu, S.; Uthandi, S. Bacterial Cellulose Nano Fiber (BCNF) as Carrier Support for the Immobilization of Probiotic, Lactobacillus Acidophilus 016. Carbohydr. Polym. 2020 , 250 , 116965. [ Google Scholar ] [ CrossRef ]
  • Żywicka, A.; Banach, A.; Junka, A.F.; Drozd, R.; Fijałkowski, K. Bacterial Cellulose as a Support for Yeast Immobilization—Correlation between Carrier Properties and Process Efficiency. J. Biotechnol. 2019 , 291 , 1–6. [ Google Scholar ] [ CrossRef ]
  • Swingler, S.; Gupta, A.; Gibson, H.; Kowalczuk, M.; Heaselgrave, W.; Radecka, I. Recent Advances and Applications of Bacterial Cellulose in Biomedicine. Polymers 2021 , 13 , 412. [ Google Scholar ] [ CrossRef ]
  • Chung, C.K.; Beekmann, U.; Kralisch, D.; Bierau, K.; Chan, A.; Ossendorp, F.; Cruz, L.J. Bacterial Cellulose as Drug Delivery System for Optimizing Release of Immune Checkpoint Blocking Antibodies. Pharmaceutics 2022 , 14 , 1351. [ Google Scholar ] [ CrossRef ]
  • Pavaloiu, R.-D.; Stoica, A.; Stroescu, M.; Dobre, T. Controlled Release of Amoxicillin from Bacterial Cellulose Membranes. Open Chem. 2014 , 12 , 962–967. [ Google Scholar ] [ CrossRef ]
  • Shao, W.; Liu, H.; Wang, S.; Wu, J.; Huang, M.; Min, H.; Liu, X. Controlled Release and Antibacterial Activity of Tetracycline Hydrochloride-Loaded Bacterial Cellulose Composite Membranes. Carbohydr. Polym. 2016 , 145 , 114–120. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Cacicedo, M.L.; Islan, G.A.; Drachemberg, M.F.; Alvarez, V.A.; Bartel, L.C.; Bolzán, A.D.; Castro, G.R. Hybrid Bacterial Cellulose–Pectin Films for Delivery of Bioactive Molecules. New J. Chem. 2018 , 42 , 7457–7467. [ Google Scholar ] [ CrossRef ]
  • Cacicedo, M.L.; Pacheco, G.; Islan, G.A.; Alvarez, V.A.; Barud, H.S.; Castro, G.R. Chitosan-Bacterial Cellulose Patch of Ciprofloxacin for Wound Dressing: Preparation and Characterization Studies. Int. J. Biol. Macromol. 2020 , 147 , 1136–1145. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Junka, A.; Bartoszewicz, M.; Dziadas, M.; Szymczyk, P.; Dydak, K.; Żywicka, A.; Owczarek, A.; Bil-Lula, I.; Czajkowska, J.; Fijałkowski, K. Application of Bacterial Cellulose Experimental Dressings Saturated with Gentamycin for Management of Bone Biofilm in Vitro and Ex Vivo. J. Biomed. Mater. Res. Part B Appl. Biomater. 2020 , 108 , 30–37. [ Google Scholar ] [ CrossRef ]
  • Trovatti, E.; Freire, C.S.R.; Pinto, P.C.; Almeida, I.F.; Costa, P.; Silvestre, A.J.D.; Neto, C.P.; Rosado, C. Bacterial Cellulose Membranes Applied in Topical and Transdermal Delivery of Lidocaine Hydrochloride and Ibuprofen: In Vitro Diffusion Studies. Int. J. Pharm. 2012 , 435 , 83–87. [ Google Scholar ] [ CrossRef ]
  • Cacicedo, M.L.; León, I.E.; Gonzalez, J.S.; Porto, L.M.; Alvarez, V.A.; Castro, G.R. Modified Bacterial Cellulose Scaffolds for Localized Doxorubicin Release in Human Colorectal HT-29 Cells. Colloids Surf. B Biointerfaces 2016 , 140 , 421–429. [ Google Scholar ] [ CrossRef ]
  • Subtaweesin, C.; Woraharn, W.; Taokaew, S.; Chiaoprakobkij, N.; Sereemaspun, A.; Phisalaphong, M. Characteristics of Curcumin-Loaded Bacterial Cellulose Films and Anticancer Properties against Malignant Melanoma Skin Cancer Cells. Appl. Sci. 2018 , 8 , 1188. [ Google Scholar ] [ CrossRef ]
  • Akagi, S.; Ando, H.; Fujita, K.; Shimizu, T.; Ishima, Y.; Tajima, K.; Matsushima, T.; Kusano, T.; Ishida, T. Therapeutic Efficacy of a Paclitaxel-Loaded Nanofibrillated Bacterial Cellulose (PTX/NFBC) Formulation in a Peritoneally Disseminated Gastric Cancer Xenograft Model. Int. J. Biol. Macromol. 2021 , 174 , 494–501. [ Google Scholar ] [ CrossRef ]
  • Hamimed, S.; Abdeljelil, N.; Landoulsi, A.; Chatti, A.; Aljabali, A.A.A.; Barhoum, A. Bacterial Cellulose Nanofibers. In Handbook of Nanocelluloses ; Springer International Publishing: Cham, Switzerland, 2022; pp. 297–334. [ Google Scholar ]
  • Wu, S.-C.; Lia, Y.-K. Application of Bacterial Cellulose Pellets in Enzyme Immobilization. J. Mol. Catal. B Enzym. 2008 , 54 , 103–108. [ Google Scholar ] [ CrossRef ]
  • Vasconcelos, N.F.; Andrade, F.K.; Vieira, L.d.A.P.; Vieira, R.S.; Vaz, J.M.; Chevallier, P.; Mantovani, D.; Borges, M.d.F.; Rosa, M. de F. Oxidized Bacterial Cellulose Membrane as Support for Enzyme Immobilization: Properties and Morphological Features. Cellulose 2020 , 27 , 3055–3083. [ Google Scholar ] [ CrossRef ]
  • Yao, W.; Wu, X.; Zhu, J.; Sun, B.; Miller, C. In Vitro Enzymatic Conversion of γ-Aminobutyric Acid Immobilization of Glutamate Decarboxylase with Bacterial Cellulose Membrane (BCM) and Non-Linear Model Establishment. Enzym. Microb. Technol. 2013 , 52 , 258–264. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Bacterial Cellulose Market, by Product Type (Industrial Grade and Technical Grade), by Application (Composites Materials, Nonwovens Absorbent Webs, Paper and Board, and Food Products), by Method, by End-Use Industry, and by Region Forecast to 2032. Available online: https://www.emergenresearch.com/industry-report/bacterial-cellulose-market (accessed on 7 August 2024).
  • Bacterial Cellulose Market Analysis & Forecast 2024–2030. Available online: https://www.profsharemarketresearch.com/bacterial-cellulose-market-report/ (accessed on 21 August 2024).
  • Jadczak, K.; Ochędzan-Siodłak, W. Bacterial Cellulose: Biopolymer with Novel Medical Applications. J. Biomater. Appl. 2023 , 38 , 51–63. [ Google Scholar ] [ CrossRef ]
  • Pötzinger, Y.; Kralisch, D.; Fischer, D. Bacterial Nanocellulose: The Future of Controlled Drug Delivery? Ther. Deliv. 2017 , 8 , 753–761. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Ho, Y.-S.; Fahad Halim, A.F.M.; Islam, M.T. The Trend of Bacterial Nanocellulose Research Published in the Science Citation Index Expanded From 2005 to 2020: A Bibliometric Analysis. Front. Bioeng. Biotechnol. 2022 , 9 , 795341. [ Google Scholar ] [ CrossRef ]
  • Moniri, M.; Boroumand Moghaddam, A.; Azizi, S.; Abdul Rahim, R.; Bin Ariff, A.; Zuhainis Saad, W.; Navaderi, M.; Mohamad, R. Production and Status of Bacterial Cellulose in Biomedical Engineering. Nanomaterials 2017 , 7 , 257. [ Google Scholar ] [ CrossRef ]
  • Adachi, O.; Moonmangmee, D.; Toyama, H.; Yamada, M.; Shinagawa, E.; Matsushita, K. New Developments in Oxidative Fermentation. Appl. Microbiol. Biotechnol. 2003 , 60 , 643–653. [ Google Scholar ] [ CrossRef ]
  • Vadanan, S.V.; Basu, A.; Lim, S. Bacterial Cellulose Production, Functionalization, and Development of Hybrid Materials Using Synthetic Biology. Polym. J. 2022 , 54 , 481–492. [ Google Scholar ] [ CrossRef ]
  • Prilepskii, A.; Nikolaev, V.; Klaving, A. Conductive Bacterial Cellulose: From Drug Delivery to Flexible Electronics. Carbohydr. Polym. 2023 , 313 , 120850. [ Google Scholar ] [ CrossRef ]
  • Fernandes, I.d.A.A.; Pedro, A.C.; Ribeiro, V.R.; Bortolini, D.G.; Ozaki, M.S.C.; Maciel, G.M.; Haminiuk, C.W.I. Bacterial Cellulose: From Production Optimization to New Applications. Int. J. Biol. Macromol. 2020 , 164 , 2598–2611. [ Google Scholar ] [ CrossRef ]
  • Fontana, J.D.; Koop, H.S.; Tiboni, M.; Grzybowski, A.; Pereira, A.; Kruger, C.D.; da Silva, M.G.R.; Wielewski, L.P. New Insights on Bacterial Cellulose. In Food Biosynthesis ; Elsevier: Amsterdam, The Netherlands, 2017; pp. 213–249. [ Google Scholar ]
  • Feng, X.; Ge, Z.; Wang, Y.; Xia, X.; Zhao, B.; Dong, M. Production and Characterization of Bacterial Cellulose from Kombucha-Fermented Soy Whey. Food Prod. Process. Nutr. 2024 , 6 , 20. [ Google Scholar ] [ CrossRef ]
  • Tang, W.; Jia, S.; Jia, Y.; Yang, H. The Influence of Fermentation Conditions and Post-Treatment Methods on Porosity of Bacterial Cellulose Membrane. World J. Microbiol. Biotechnol. 2010 , 26 , 125–131. [ Google Scholar ] [ CrossRef ]
  • Bäckdahl, H.; Esguerra, M.; Delbro, D.; Risberg, B.; Gatenholm, P. Engineering Microporosity in Bacterial Cellulose Scaffolds. J. Tissue Eng. Regen. Med. 2008 , 2 , 320–330. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Tang, K.Y.; Heng, J.Z.X.; Chai, C.H.T.; Chan, C.Y.; Low, B.Q.L.; Chong, S.M.E.; Loh, H.Y.; Li, Z.; Ye, E.; Loh, X.J. Modified Bacterial Cellulose for Biomedical Applications. Chem. Asian J. 2022 , 17 , e202200598. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Zhang, L.-K.; Du, S.; Wang, X.; Jiao, Y.; Yin, L.; Zhang, Y.; Guan, Y.-Q. Bacterial Cellulose Based Composites Enhanced Transdermal Drug Targeting for Breast Cancer Treatment. Chem. Eng. J. 2019 , 370 , 749–759. [ Google Scholar ] [ CrossRef ]
  • Hu, W.; Chen, S.; Yang, J.; Li, Z.; Wang, H. Functionalized Bacterial Cellulose Derivatives and Nanocomposites. Carbohydr. Polym. 2014 , 101 , 1043–1060. [ Google Scholar ] [ CrossRef ]
  • Ul-Islam, M.; Khan, S.; Fatima, A.; Ahmad, M.W.; Khan, M.S.; Ul Islam, S.; Manan, S.; Ullah, M.W. Production of Bio-Cellulose from Renewable Resources: Properties and Applications. In Renewable Polymers and Polymer-Metal Oxide Composites ; Elsevier: Amsterdam, The Netherlands, 2022; pp. 307–339. [ Google Scholar ]
  • Ciolacu, D.E.; Nicu, R.; Ciolacu, F. Cellulose-Based Hydrogels as Sustained Drug-Delivery Systems. Materials 2020 , 13 , 5270. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Jankau, J.; Błażyńska-Spychalska, A.; Kubiak, K.; Jędrzejczak-Krzepkowska, M.; Pankiewicz, T.; Ludwicka, K.; Dettlaff, A.; Pęksa, R. Bacterial Cellulose Properties Fulfilling Requirements for a Biomaterial of Choice in Reconstructive Surgery and Wound Healing. Front. Bioeng. Biotechnol. 2022 , 9 , 805053. [ Google Scholar ] [ CrossRef ]
  • Klemm, D.; Cranston, E.D.; Fischer, D.; Gama, M.; Kedzior, S.A.; Kralisch, D.; Kramer, F.; Kondo, T.; Lindström, T.; Nietzsche, S.; et al. Nanocellulose as a Natural Source for Groundbreaking Applications in Materials Science: Today’s State. Mater. Today 2018 , 21 , 720–748. [ Google Scholar ] [ CrossRef ]
  • Basu, A.; Strømme, M.; Ferraz, N. Towards Tunable Protein-Carrier Wound Dressings Based on Nanocellulose Hydrogels Crosslinked with Calcium Ions. Nanomaterials 2018 , 8 , 550. [ Google Scholar ] [ CrossRef ]
  • Huo, Y.; Liu, Y.; Xia, M.; Du, H.; Lin, Z.; Li, B.; Liu, H. Nanocellulose-Based Composite Materials Used in Drug Delivery Systems. Polymers 2022 , 14 , 2648. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Monika, P.; Chandraprabha, M.N.; Rangarajan, A.; Waiker, P.V.; Chidambara Murthy, K.N. Challenges in Healing Wound: Role of Complementary and Alternative Medicine. Front. Nutr. 2022 , 8 , 791899. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Lin, Y.-K.; Chen, K.-H.; Ou, K.-L.; Liu, M. Effects of Different Extracellular Matrices and Growth Factor Immobilization on Biodegradability and Biocompatibility of Macroporous Bacterial Cellulose. J. Bioact. Compat. Polym. 2011 , 26 , 508–518. [ Google Scholar ] [ CrossRef ]
  • Zhang, W.; Liu, Y.; Zhang, H. Extracellular Matrix: An Important Regulator of Cell Functions and Skeletal Muscle Development. Cell Biosci. 2021 , 11 , 65. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Sharda, D.; Kaur, P.; Choudhury, D. Protein-Modified Nanomaterials: Emerging Trends in Skin Wound Healing. Discov. Nano 2023 , 18 , 127. [ Google Scholar ] [ CrossRef ]
  • Ben Ayed, H.; Bardaa, S.; Moalla, D.; Jridi, M.; Maalej, H.; Sahnoun, Z.; Rebai, T.; Jacques, P.; Nasri, M.; Hmidet, N. Wound Healing and in Vitro Antioxidant Activities of Lipopeptides Mixture Produced by Bacillus Mojavensis A21. Process Biochem. 2015 , 50 , 1023–1030. [ Google Scholar ] [ CrossRef ]
  • Savitskaya, I.S.; Shokatayeva, D.H.; Kistaubayeva, A.S.; Ignatova, L.V.; Digel, I.E. Antimicrobial and Wound Healing Properties of a Bacterial Cellulose Based Material Containing B. Subtilis Cells. Heliyon 2019 , 5 , e02592. [ Google Scholar ] [ CrossRef ]
  • Tudoroiu, E.-E.; Dinu-Pîrvu, C.-E.; Albu Kaya, M.G.; Popa, L.; Anuța, V.; Prisada, R.M.; Ghica, M.V. An Overview of Cellulose Derivatives-Based Dressings for Wound-Healing Management. Pharmaceuticals 2021 , 14 , 1215. [ Google Scholar ] [ CrossRef ]
  • Zhong, S.P.; Zhang, Y.Z.; Lim, C.T. Tissue Scaffolds for Skin Wound Healing and Dermal Reconstruction. WIREs Nanomed. Nanobiotechnol. 2010 , 2 , 510–525. [ Google Scholar ] [ CrossRef ]
  • Li, M.; Xia, W.; Khoong, Y.M.; Huang, L.; Huang, X.; Liang, H.; Zhao, Y.; Mao, J.; Yu, H.; Zan, T. Smart and Versatile Biomaterials for Cutaneous Wound Healing. Biomater. Res. 2023 , 27 , 87. [ Google Scholar ] [ CrossRef ]
  • Orlando, I.; Basnett, P.; Nigmatullin, R.; Wang, W.; Knowles, J.C.; Roy, I. Chemical Modification of Bacterial Cellulose for the Development of an Antibacterial Wound Dressing. Front. Bioeng. Biotechnol. 2020 , 8 , 557885. [ Google Scholar ] [ CrossRef ]
  • Portela, R.; Leal, C.R.; Almeida, P.L.; Sobral, R.G. Bacterial Cellulose: A Versatile Biopolymer for Wound Dressing Applications. Microb. Biotechnol. 2019 , 12 , 586–610. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Lin, S.-P.; Loira Calvar, I.; Catchmark, J.M.; Liu, J.-R.; Demirci, A.; Cheng, K.-C. Biosynthesis, Production and Applications of Bacterial Cellulose. Cellulose 2013 , 20 , 2191–2219. [ Google Scholar ] [ CrossRef ]
  • de Amorim, J.D.P.; da Silva Junior, C.J.G.; de Medeiros, A.D.M.; do Nascimento, H.A.; Sarubbo, M.; de Medeiros, T.P.M.; Costa, A.F.d.S.; Sarubbo, L.A. Bacterial Cellulose as a Versatile Biomaterial for Wound Dressing Application. Molecules 2022 , 27 , 5580. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Raut, M.; Asare, E.; Syed Mohamed, S.; Amadi, E.; Roy, I. Bacterial Cellulose-Based Blends and Composites: Versatile Biomaterials for Tissue Engineering Applications. Int. J. Mol. Sci. 2023 , 24 , 986. [ Google Scholar ] [ CrossRef ]
  • He, W.; Wu, J.; Xu, J.; Mosselhy, D.A.; Zheng, Y.; Yang, S. Bacterial Cellulose: Functional Modification and Wound Healing Applications. Adv. Wound Care 2021 , 10 , 623–640. [ Google Scholar ] [ CrossRef ]
  • Kavitha, K.V. Choice of Wound Care in Diabetic Foot Ulcer: A Practical Approach. World J. Diabetes 2014 , 5 , 546. [ Google Scholar ] [ CrossRef ]
  • Abdelrahman, T.; Newton, H. Wound Dressings: Principles and Practice. Surgery 2011 , 29 , 491–495. [ Google Scholar ] [ CrossRef ]
  • Meftahi, A.; Khajavi, R.; Rashidi, A.; Sattari, M.; Yazdanshenas, M.E.; Torabi, M. The Effects of Cotton Gauze Coating with Microbial Cellulose. Cellulose 2010 , 17 , 199–204. [ Google Scholar ] [ CrossRef ]
  • Wiegand, C.; Moritz, S.; Hessler, N.; Kralisch, D.; Wesarg, F.; Müller, F.A.; Fischer, D.; Hipler, U.-C. Antimicrobial Functionalization of Bacterial Nanocellulose by Loading with Polihexanide and Povidone-Iodine. J. Mater. Sci. Mater. Med. 2015 , 26 , 245. [ Google Scholar ] [ CrossRef ]
  • Sosnik, A.; Seremeta, K. Polymeric Hydrogels as Technology Platform for Drug Delivery Applications. Gels 2017 , 3 , 25. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Li, Y.; Wang, S.; Huang, R.; Huang, Z.; Hu, B.; Zheng, W.; Yang, G.; Jiang, X. Evaluation of the Effect of the Structure of Bacterial Cellulose on Full Thickness Skin Wound Repair on a Microfluidic Chip. Biomacromolecules 2015 , 16 , 780–789. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Moradpoor, H.; Mohammadi, H.; Safaei, M.; Mozaffari, H.R.; Sharifi, R.; Gorji, P.; Sulong, A.B.; Muhamad, N.; Ebadi, M. Recent Advances on Bacterial Cellulose-Based Wound Management: Promises and Challenges. Int. J. Polym. Sci. 2022 , 2022 , 1–24. [ Google Scholar ] [ CrossRef ]
  • Meng, S.; Wu, H.; Xiao, D.; Lan, S.; Dong, A. Recent Advances in Bacterial Cellulose-Based Antibacterial Composites for Infected Wound Therapy. Carbohydr. Polym. 2023 , 316 , 121082. [ Google Scholar ] [ CrossRef ]
  • Jabbari, F.; Babaeipour, V. Bacterial Cellulose as an Ideal Potential Treatment for Burn Wounds: A Comprehensive Review. Wound Repair Regen. 2024 , 32 , 323–339. [ Google Scholar ] [ CrossRef ]
  • Rajwade, J.M.; Paknikar, K.M.; Kumbhar, J.V. Applications of Bacterial Cellulose and Its Composites in Biomedicine. Appl. Microbiol. Biotechnol. 2015 , 99 , 2491–2511. [ Google Scholar ] [ CrossRef ]
  • Taokaew, S.; Phisalaphong, M.; Newby, B.Z. Modification of Bacterial Cellulose with Organosilanes to Improve Attachment and Spreading of Human Fibroblasts. Cellulose 2015 , 22 , 2311–2324. [ Google Scholar ] [ CrossRef ]
  • Czaplewski, L.; Bax, R.; Clokie, M.; Dawson, M.; Fairhead, H.; Fischetti, V.A.; Foster, S.; Gilmore, B.F.; Hancock, R.E.W.; Harper, D.; et al. Alternatives to Antibiotics—A Pipeline Portfolio Review. Lancet Infect. Dis. 2016 , 16 , 239–251. [ Google Scholar ] [ CrossRef ]
  • Sulaeva, I.; Henniges, U.; Rosenau, T.; Potthast, A. Bacterial Cellulose as a Material for Wound Treatment: Properties and Modifications. A Review. Biotechnol. Adv. 2015 , 33 , 1547–1571. [ Google Scholar ] [ CrossRef ]
  • Zheng, L.; Li, S.; Luo, J.; Wang, X. Latest Advances on Bacterial Cellulose-Based Antibacterial Materials as Wound Dressings. Front. Bioeng. Biotechnol. 2020 , 8 , 593768. [ Google Scholar ] [ CrossRef ]
  • Horue, M.; Silva, J.M.; Berti, I.R.; Brandão, L.R.; Barud, H.d.S.; Castro, G.R. Bacterial Cellulose-Based Materials as Dressings for Wound Healing. Pharmaceutics 2023 , 15 , 424. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Czaja, W.K.; Young, D.J.; Kawecki, M.; Brown, R.M. The Future Prospects of Microbial Cellulose in Biomedical Applications. Biomacromolecules 2007 , 8 , 1–12. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Ullah, H.; Wahid, F.; Santos, H.A.; Khan, T. Advances in Biomedical and Pharmaceutical Applications of Functional Bacterial Cellulose-Based Nanocomposites. Carbohydr. Polym. 2016 , 150 , 330–352. [ Google Scholar ] [ CrossRef ]
  • Zhou, C.; Yang, Z.; Xun, X.; Ma, L.; Chen, Z.; Hu, X.; Wu, X.; Wan, Y.; Ao, H. De Novo Strategy with Engineering a Multifunctional Bacterial Cellulose-Based Dressing for Rapid Healing of Infected Wounds. Bioact. Mater. 2022 , 13 , 212–222. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Pasaribu, K.M.; Ilyas, S.; Tamrin, T.; Radecka, I.; Swingler, S.; Gupta, A.; Stamboulis, A.G.; Gea, S. Bioactive Bacterial Cellulose Wound Dressings for Burns with Collagen In-Situ and Chitosan Ex-Situ Impregnation. Int. J. Biol. Macromol. 2023 , 230 , 123118. [ Google Scholar ] [ CrossRef ]
  • Yuan, H.; Chen, L.; Hong, F.F. A Biodegradable Antibacterial Nanocomposite Based on Oxidized Bacterial Nanocellulose for Rapid Hemostasis and Wound Healing. ACS Appl. Mater. Interfaces 2020 , 12 , 3382–3392. [ Google Scholar ] [ CrossRef ]
  • Chiaoprakobkij, N.; Suwanmajo, T.; Sanchavanakit, N.; Phisalaphong, M. Curcumin-Loaded Bacterial Cellulose/Alginate/Gelatin as A Multifunctional Biopolymer Composite Film. Molecules 2020 , 25 , 3800. [ Google Scholar ] [ CrossRef ]
  • Lamboni, L.; Li, Y.; Liu, J.; Yang, G. Silk Sericin-Functionalized Bacterial Cellulose as a Potential Wound-Healing Biomaterial. Biomacromolecules 2016 , 17 , 3076–3084. [ Google Scholar ] [ CrossRef ]
  • Aditya, T.; Allain, J.P.; Jaramillo, C.; Restrepo, A.M. Surface Modification of Bacterial Cellulose for Biomedical Applications. Int. J. Mol. Sci. 2022 , 23 , 610. [ Google Scholar ] [ CrossRef ]
  • Radu, C.D.; Verestiuc, L.; Ulea, E.; Lipsa, F.D.; Vulpe, V.; Munteanu, C.; Bulgariu, L.; Pașca, S.; Tamas, C.; Ciuntu, B.M.; et al. Evaluation of Keratin/Bacterial Cellulose Based Scaffolds as Potential Burned Wound Dressing. Appl. Sci. 2021 , 11 , 1995. [ Google Scholar ] [ CrossRef ]
  • Sampaio, L.M.P.; Padrão, J.; Faria, J.; Silva, J.P.; Silva, C.J.; Dourado, F.; Zille, A. Laccase Immobilization on Bacterial Nanocellulose Membranes: Antimicrobial, Kinetic and Stability Properties. Carbohydr. Polym. 2016 , 145 , 1–12. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Kim, S.; Ko, J.; Choi, J.H.; Kang, J.Y.; Lim, C.; Shin, M.; Lee, D.W.; Kim, J.W. Antigen–Antibody Interaction-Derived Bioadhesion of Bacterial Cellulose Nanofibers to Promote Topical Wound Healing. Adv. Funct. Mater. 2022 , 32 , 2110557. [ Google Scholar ] [ CrossRef ]
  • Vasconcelos, N.F.; Cunha, A.P.; Ricardo, N.M.P.S.; Freire, R.S.; Vieira, L.d.A.P.; Brígida, A.I.S.; Borges, M.d.F.; Rosa, M.d.F.; Vieira, R.S.; Andrade, F.K. Papain Immobilization on Heterofunctional Membrane Bacterial Cellulose as a Potential Strategy for the Debridement of Skin Wounds. Int. J. Biol. Macromol. 2020 , 165 , 3065–3077. [ Google Scholar ] [ CrossRef ]
  • Bayazidi, P.; Almasi, H.; Asl, A.K. Immobilization of Lysozyme on Bacterial Cellulose Nanofibers: Characteristics, Antimicrobial Activity and Morphological Properties. Int. J. Biol. Macromol. 2018 , 107 , 2544–2551. [ Google Scholar ] [ CrossRef ]
  • Costa de Oliveira Souza, C.M.; de Souza, C.F.; Mogharbel, B.F.; Irioda, A.C.; Cavichiolo Franco, C.R.; Sierakowski, M.R.; Athayde Teixeira de Carvalho, K. Nanostructured Cellulose–Gellan–Xyloglucan–Lysozyme Dressing Seeded with Mesenchymal Stem Cells for Deep Second-Degree Burn Treatment. Int. J. Nanomed. 2021 , 16 , 833–850. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • dos Santos, C.A.; dos Santos, G.R.; Soeiro, V.S.; dos Santos, J.R.; Rebelo, M.d.A.; Chaud, M.V.; Gerenutti, M.; Grotto, D.; Pandit, R.; Rai, M.; et al. Bacterial Nanocellulose Membranes Combined with Nisin: A Strategy to Prevent Microbial Growth. Cellulose 2018 , 25 , 6681–6689. [ Google Scholar ] [ CrossRef ]
  • Rollini, M.; Musatti, A.; Cavicchioli, D.; Bussini, D.; Farris, S.; Rovera, C.; Romano, D.; De Benedetti, S.; Barbiroli, A. From Cheese Whey Permeate to Sakacin-A/Bacterial Cellulose Nanocrystal Conjugates for Antimicrobial Food Packaging Applications: A Circular Economy Case Study. Sci. Rep. 2020 , 10 , 21358. [ Google Scholar ] [ CrossRef ]
  • Malheiros, P.S.; Jozala, A.F.; Pessoa, A., Jr.; Vila, M.M.D.C.; Balcão, V.M.; Franco, B.D.G.M. Immobilization of Antimicrobial Peptides from Lactobacillus Sakei Subsp. Sakei 2a in Bacterial Cellulose: Structural and Functional Stabilization. Food Packag. Shelf Life 2018 , 17 , 25–29. [ Google Scholar ] [ CrossRef ]
  • Benevenuto, L.G.D.; da Silva Barud, H.; Cruz, S.A.; Caillier, B.; da Silva Paiva, R.; Achcar, J.A.; Montrezor, L.H. Bacterial Cellulose-Based Cell Culture Platform Modified by Oxygen Plasma for Tissue Engineering Applications. Cellulose 2023 , 30 , 9625–9634. [ Google Scholar ] [ CrossRef ]
  • Huang, Y.; Wang, J.; Yang, F.; Shao, Y.; Zhang, X.; Dai, K. Modification and Evaluation of Micro-Nano Structured Porous Bacterial Cellulose Scaffold for Bone Tissue Engineering. Mater. Sci. Eng. C 2017 , 75 , 1034–1041. [ Google Scholar ] [ CrossRef ]
  • Khan, S.; Ul-Islam, M.; Ikram, M.; Ullah, M.W.; Israr, M.; Subhan, F.; Kim, Y.; Jang, J.H.; Yoon, S.; Park, J.K. Three-Dimensionally Microporous and Highly Biocompatible Bacterial Cellulose–Gelatin Composite Scaffolds for Tissue Engineering Applications. RSC Adv. 2016 , 6 , 110840–110849. [ Google Scholar ] [ CrossRef ]
  • Nagaoka, M.; Jiang, H.-L.; Hoshiba, T.; Akaike, T.; Cho, C.-S. Application of Recombinant Fusion Proteins for Tissue Engineering. Ann. Biomed. Eng. 2010 , 38 , 683–693. [ Google Scholar ] [ CrossRef ]
  • Lee, S.-H.; An, S.-J.; Lim, Y.-M.; Huh, J.-B. The Efficacy of Electron Beam Irradiated Bacterial Cellulose Membranes as Compared with Collagen Membranes on Guided Bone Regeneration in Peri-Implant Bone Defects. Materials 2017 , 10 , 1018. [ Google Scholar ] [ CrossRef ]
  • Xiong, G.; Luo, H.; Zhu, Y.; Raman, S.; Wan, Y. Creation of Macropores in Three-Dimensional Bacterial Cellulose Scaffold for Potential Cancer Cell Culture. Carbohydr. Polym. 2014 , 114 , 553–557. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Picheth, G.F.; Pirich, C.L.; Sierakowski, M.R.; Woehl, M.A.; Sakakibara, C.N.; de Souza, C.F.; Martin, A.A.; da Silva, R.; de Freitas, R.A. Bacterial Cellulose in Biomedical Applications: A Review. Int. J. Biol. Macromol. 2017 , 104 , 97–106. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Yadav, V.; Sun, L.; Panilaitis, B.; Kaplan, D.L. In Vitro Chondrogenesis with Lysozyme Susceptible Bacterial Cellulose as a Scaffold. J. Tissue Eng. Regen. Med. 2015 , 9 , E276–E288. [ Google Scholar ] [ CrossRef ]
  • Girard, V.; Chaussé, J.; Vermette, P. Bacterial Cellulose: A Comprehensive Review. J. Appl. Polym. Sci. 2024 , 141 , e55163. [ Google Scholar ] [ CrossRef ]
  • Osorio, M.; Cañas, A.; Puerta, J.; Díaz, L.; Naranjo, T.; Ortiz, I.; Castro, C. Ex Vivo and In Vivo Biocompatibility Assessment (Blood and Tissue) of Three-Dimensional Bacterial Nanocellulose Biomaterials for Soft Tissue Implants. Sci. Rep. 2019 , 9 , 10553. [ Google Scholar ] [ CrossRef ]
  • Acasigua, G.; Olyveira, G.; Costa, L.; Braghirolli, D.; Fossati, A.; Guastaldi, A.; Pranke, P.; Daltro, G.; Basmaji, P. Novel Chemically Modified Bacterial Cellulose Nanocomposite as Potential Biomaterial for Stem Cell Therapy Applications. Curr. Stem Cell Res. Ther. 2014 , 9 , 117–123. [ Google Scholar ] [ CrossRef ]
  • Fu, L.; Zhou, P.; Zhang, S.; Yang, G. Evaluation of Bacterial Nanocellulose-Based Uniform Wound Dressing for Large Area Skin Transplantation. Mater. Sci. Eng. C 2013 , 33 , 2995–3000. [ Google Scholar ] [ CrossRef ]
  • Khan, S.; Ul-Islam, M.; Ullah, M.W.; Zhu, Y.; Narayanan, K.B.; Han, S.S.; Park, J.K. Fabrication Strategies and Biomedical Applications of Three-Dimensional Bacterial Cellulose-Based Scaffolds: A Review. Int. J. Biol. Macromol. 2022 , 209 , 9–30. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Pértile, R.A.N.; Moreira, S.; Gil da Costa, R.M.; Correia, A.; Guãrdao, L.; Gartner, F.; Vilanova, M.; Gama, M. Bacterial Cellulose: Long-Term Biocompatibility Studies. J. Biomater. Sci. Polym. Ed. 2012 , 23 , 1339–1354. [ Google Scholar ] [ CrossRef ]
  • Castner, D.G. Biomedical Surface Analysis: Evolution and Future Directions (Review). Biointerphases 2017 , 12 , 02C301. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Dugan, J.M.; Gough, J.E.; Eichhorn, S.J. Bacterial Cellulose Scaffolds and Cellulose Nanowhiskers for Tissue Engineering. Nanomedicine 2013 , 8 , 287–298. [ Google Scholar ] [ CrossRef ]
  • Zhijiang, C.; Guang, Y. Bacterial Cellulose/Collagen Composite: Characterization and First Evaluation of Cytocompatibility. J. Appl. Polym. Sci. 2011 , 120 , 2938–2944. [ Google Scholar ] [ CrossRef ]
  • Albu, M.; Vuluga, Z.; Panaitescu, D.; Vuluga, D.; Căşărică, A.; Ghiurea, M. Morphology and Thermal Stability of Bacterial Cellulose/Collagen Composites. Open Chem. 2014 , 12 , 968–975. [ Google Scholar ] [ CrossRef ]
  • Adhikari, J.; Dasgupta, S.; Barui, A.; Ghosh, M.; Saha, P. Collagen Incorporated Functionalized Bacterial Cellulose Composite: A Macromolecular Approach for Successful Tissue Engineering Applications. Cellulose 2023 , 30 , 9079–9111. [ Google Scholar ] [ CrossRef ]
  • Noh, Y.K.; Dos Santos Da Costa, A.; Park, Y.S.; Du, P.; Kim, I.-H.; Park, K. Fabrication of Bacterial Cellulose-Collagen Composite Scaffolds and Their Osteogenic Effect on Human Mesenchymal Stem Cells. Carbohydr. Polym. 2019 , 219 , 210–218. [ Google Scholar ] [ CrossRef ]
  • Saska, S.; Teixeira, L.N.; Tambasco de Oliveira, P.; Minarelli Gaspar, A.M.; Lima Ribeiro, S.J.; Messaddeq, Y.; Marchetto, R. Bacterial Cellulose-Collagen Nanocomposite for Bone Tissue Engineering. J. Mater. Chem. 2012 , 22 , 22102. [ Google Scholar ] [ CrossRef ]
  • Svensson, A.; Nicklasson, E.; Harrah, T.; Panilaitis, B.; Kaplan, D.L.; Brittberg, M.; Gatenholm, P. Bacterial Cellulose as a Potential Scaffold for Tissue Engineering of Cartilage. Biomaterials 2005 , 26 , 419–431. [ Google Scholar ] [ CrossRef ]
  • Koike, T.; Sha, J.; Bai, Y.; Matsuda, Y.; Hideshima, K.; Yamada, T.; Kanno, T. Efficacy of Bacterial Cellulose as a Carrier of BMP-2 for Bone Regeneration in a Rabbit Frontal Sinus Model. Materials 2019 , 12 , 2489. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Gorgieva, S.; Hribernik, S. Microstructured and Degradable Bacterial Cellulose–Gelatin Composite Membranes: Mineralization Aspects and Biomedical Relevance. Nanomaterials 2019 , 9 , 303. [ Google Scholar ] [ CrossRef ]
  • Cakmak, A.M.; Unal, S.; Sahin, A.; Oktar, F.N.; Sengor, M.; Ekren, N.; Gunduz, O.; Kalaskar, D.M. 3D Printed Polycaprolactone/Gelatin/Bacterial Cellulose/Hydroxyapatite Composite Scaffold for Bone Tissue Engineering. Polymers 2020 , 12 , 1962. [ Google Scholar ] [ CrossRef ]
  • Keskin, Z.; Sendemir Urkmez, A.; Hames, E.E. Novel Keratin Modified Bacterial Cellulose Nanocomposite Production and Characterization for Skin Tissue Engineering. Mater. Sci. Eng. C 2017 , 75 , 1144–1153. [ Google Scholar ] [ CrossRef ]
  • Pértile, R.; Moreira, S.; Andrade, F.; Domingues, L.; Gama, M. Bacterial Cellulose Modified Using Recombinant Proteins to Improve Neuronal and Mesenchymal Cell Adhesion. Biotechnol. Prog. 2012 , 28 , 526–532. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Yang, J.; Zhu, Z.; Liu, Y.; Zheng, Y.; Xie, Y.; Lin, J.; Cai, T. Double-Modified Bacterial Cellulose/Soy Protein Isolate Composites by Laser Hole Forming and Selective Oxidation Used for Urethral Repair. Biomacromolecules 2022 , 23 , 291–302. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Zang, S.; Zhang, R.; Chen, H.; Lu, Y.; Zhou, J.; Chang, X.; Qiu, G.; Wu, Z.; Yang, G. Investigation on Artificial Blood Vessels Prepared from Bacterial Cellulose. Mater. Sci. Eng. C 2015 , 46 , 111–117. [ Google Scholar ] [ CrossRef ]
  • Fink, H.; Faxälv, L.; Molnár, G.F.; Drotz, K.; Risberg, B.; Lindahl, T.L.; Sellborn, A. Real-Time Measurements of Coagulation on Bacterial Cellulose and Conventional Vascular Graft Materials. Acta Biomater. 2010 , 6 , 1125–1130. [ Google Scholar ] [ CrossRef ]
  • Roberts, E.L.; Abdollahi, S.; Oustadi, F.; Stephens, E.D.; Badv, M. Bacterial-Nanocellulose-Based Biointerfaces and Biomimetic Constructs for Blood-Contacting Medical Applications. ACS Mater. Au 2023 , 3 , 418–441. [ Google Scholar ] [ CrossRef ]
  • Klemm, D.; Schumann, D.; Udhardt, U.; Marsch, S. Bacterial Synthesized Cellulose—Artificial Blood Vessels for Microsurgery. Prog. Polym. Sci. 2001 , 26 , 1561–1603. [ Google Scholar ] [ CrossRef ]
  • Lin, N.; Dufresne, A. Nanocellulose in Biomedicine: Current Status and Future Prospect. Eur. Polym. J. 2014 , 59 , 302–325. [ Google Scholar ] [ CrossRef ]
  • Klemm, D.; Heublein, B.; Fink, H.; Bohn, A. Cellulose: Fascinating Biopolymer and Sustainable Raw Material. Angew. Chem. Int. Ed. 2005 , 44 , 3358–3393. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Brown, E.E.; Zhang, J.; Laborie, M.-P.G. Never-Dried Bacterial Cellulose/Fibrin Composites: Preparation, Morphology and Mechanical Properties. Cellulose 2011 , 18 , 631–641. [ Google Scholar ] [ CrossRef ]
  • Andrade, F.K.; Silva, J.P.; Carvalho, M.; Castanheira, E.M.S.; Soares, R.; Gama, M. Studies on the Hemocompatibility of Bacterial Cellulose. J. Biomed. Mater. Res. Part A 2011 , 98A , 554–566. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Tibbitt, M.W.; Anseth, K.S. Hydrogels as Extracellular Matrix Mimics for 3D Cell Culture. Biotechnol. Bioeng. 2009 , 103 , 655–663. [ Google Scholar ] [ CrossRef ]
  • Leitão, A.F.; Faria, M.A.; Faustino, A.M.R.; Moreira, R.; Mela, P.; Loureiro, L.; Silva, I.; Gama, M. A Novel Small-Caliber Bacterial Cellulose Vascular Prosthesis: Production, Characterization, and Preliminary In Vivo Testing. Macromol. Biosci. 2016 , 16 , 139–150. [ Google Scholar ] [ CrossRef ]
  • Chaicharoenaudomrung, N.; Kunhorm, P.; Noisa, P. Three-Dimensional Cell Culture Systems as an in Vitro Platform for Cancer and Stem Cell Modeling. World J. Stem Cells 2019 , 11 , 1065–1083. [ Google Scholar ] [ CrossRef ]
  • Unal, S.; Arslan, S.; Yilmaz, B.K.; Oktar, F.N.; Sengil, A.Z.; Gunduz, O. Production and Characterization of Bacterial Cellulose Scaffold and Its Modification with Hyaluronic Acid and Gelatin for Glioblastoma Cell Culture. Cellulose 2021 , 28 , 117–132. [ Google Scholar ] [ CrossRef ]
  • Opwis, K.; Gutmann, J.S. Surface Modification of Textile Materials with Hydrophobins. Text. Res. J. 2011 , 81 , 1594–1602. [ Google Scholar ] [ CrossRef ]
  • Wan, Z.; Wang, L.; Ma, L.; Sun, Y.; Yang, X. Controlled Hydrophobic Biosurface of Bacterial Cellulose Nanofibers through Self-Assembly of Natural Zein Protein. ACS Biomater. Sci. Eng. 2017 , 3 , 1595–1604. [ Google Scholar ] [ CrossRef ]
  • Rijal, G.; Li, W. 3D Scaffolds in Breast Cancer Research. Biomaterials 2016 , 81 , 135–156. [ Google Scholar ] [ CrossRef ]
  • Reis, E.M.d.; Berti, F.V.; Colla, G.; Porto, L.M. Bacterial Nanocellulose-IKVAV Hydrogel Matrix Modulates Melanoma Tumor Cell Adhesion and Proliferation and Induces Vasculogenic Mimicry in Vitro. J. Biomed. Mater. Res. Part B Appl. Biomater. 2018 , 106 , 2741–2749. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Xiong, G.; Luo, H.; Gu, F.; Zhang, J.; Hu, D.; Wan, Y. A Novel in Vitro Three-Dimensional Macroporous Scaffolds from Bacterial Cellulose for Culture of Breast Cancer Cells. J. Biomater. Nanobiotechnol. 2013 , 4 , 316–326. [ Google Scholar ] [ CrossRef ]
  • Luo, H.; Gu, P.; Xiong, G.; Hu, D. A Multichanneled Bacterial Cellulose Scaffold for 3D in Vitro Cancer Culture. Cellul. Chem. Technol. 2016 , 50 , 49–56. [ Google Scholar ]
  • Islam, S.U.; Ul-Islam, M.; Ahsan, H.; Ahmed, M.B.; Shehzad, A.; Fatima, A.; Sonn, J.K.; Lee, Y.S. Potential Applications of Bacterial Cellulose and Its Composites for Cancer Treatment. Int. J. Biol. Macromol. 2021 , 168 , 301–309. [ Google Scholar ] [ CrossRef ]
  • Gao, C.; Wan, Y.; Yang, C.; Dai, K.; Tang, T.; Luo, H.; Wang, J. Preparation and Characterization of Bacterial Cellulose Sponge with Hierarchical Pore Structure as Tissue Engineering Scaffold. J. Porous Mater. 2011 , 18 , 139–145. [ Google Scholar ] [ CrossRef ]
  • Ul-Islam, M.; Subhan, F.; Islam, S.U.; Khan, S.; Shah, N.; Manan, S.; Ullah, M.W.; Yang, G. Development of Three-Dimensional Bacterial Cellulose/Chitosan Scaffolds: Analysis of Cell-Scaffold Interaction for Potential Application in the Diagnosis of Ovarian Cancer. Int. J. Biol. Macromol. 2019 , 137 , 1050–1059. [ Google Scholar ] [ CrossRef ]
  • Chen, Y.; Zhou, X.; Lin, Q.; Jiang, D. Bacterial Cellulose/Gelatin Composites: In Situ Preparation and Glutaraldehyde Treatment. Cellulose 2014 , 21 , 2679–2693. [ Google Scholar ] [ CrossRef ]
  • Lopes, T.D.; Riegel-Vidotti, I.C.; Grein, A.; Tischer, C.A.; Faria-Tischer, P.C. de S. Bacterial Cellulose and Hyaluronic Acid Hybrid Membranes: Production and Characterization. Int. J. Biol. Macromol. 2014 , 67 , 401–408. [ Google Scholar ] [ CrossRef ]
  • de Oliveira, S.A.; da Silva, B.C.; Riegel-Vidotti, I.C.; Urbano, A.; de Sousa Faria-Tischer, P.C.; Tischer, C.A. Production and Characterization of Bacterial Cellulose Membranes with Hyaluronic Acid from Chicken Comb. Int. J. Biol. Macromol. 2017 , 97 , 642–653. [ Google Scholar ] [ CrossRef ]
  • Caro-Astorga, J.; Walker, K.T.; Herrera, N.; Lee, K.-Y.; Ellis, T. Bacterial Cellulose Spheroids as Building Blocks for 3D and Patterned Living Materials and for Regeneration. Nat. Commun. 2021 , 12 , 5027. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Innala, M.; Riebe, I.; Kuzmenko, V.; Sundberg, J.; Gatenholm, P.; Hanse, E.; Johannesson, S. 3D Culturing and Differentiation of SH-SY5Y Neuroblastoma Cells on Bacterial Nanocellulose Scaffolds. Artif. Cells Nanomed. Biotechnol. 2014 , 42 , 302–308. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Wang, B.; Ji, P.; Ma, Y.; Song, J.; You, Z.; Chen, S. Bacterial Cellulose Nanofiber Reinforced Poly(Glycerol-Sebacate) Biomimetic Matrix for 3D Cell Culture. Cellulose 2021 , 28 , 8483–8492. [ Google Scholar ] [ CrossRef ]
  • Wang, J.; Zhao, L.; Zhang, A.; Huang, Y.; Tavakoli, J.; Tang, Y. Novel Bacterial Cellulose/Gelatin Hydrogels as 3D Scaffolds for Tumor Cell Culture. Polymers 2018 , 10 , 581. [ Google Scholar ] [ CrossRef ]
  • Robbins, M.; Pisupati, V.; Azzarelli, R.; Nehme, S.I.; Barker, R.A.; Fruk, L.; Schierle, G.S.K. Biofunctionalised Bacterial Cellulose Scaffold Supports the Patterning and Expansion of Human Embryonic Stem Cell-Derived Dopaminergic Progenitor Cells. Stem Cell Res. Ther. 2021 , 12 , 574. [ Google Scholar ] [ CrossRef ]
  • He, M.; Yan, Y.; Liu, X.; Li, L.; Yang, B.; Liu, M.; Yu, Q.; Wang, E.; Li, P.; Liu, T.; et al. A Nanobody-Mediated Drug System against Largemouth Bass Virus Delivered by Bacterial Nanocellulose in Micropterus Salmoides. Int. J. Biol. Macromol. 2024 , 266 , 131146. [ Google Scholar ] [ CrossRef ]
  • Liang, S. Advances in Drug Delivery Applications of Modified Bacterial Cellulose-Based Materials. Front. Bioeng. Biotechnol. 2023 , 11 , 1252706. [ Google Scholar ] [ CrossRef ]
  • Zhang, W.; Wang, X.; Wang, J.; Zhang, L. Drugs Adsorption and Release Behavior of Collagen/Bacterial Cellulose Porous Microspheres. Int. J. Biol. Macromol. 2019 , 140 , 196–205. [ Google Scholar ] [ CrossRef ]
  • Qi, Z.; Pei, P.; Zhang, Y.; Chen, H.; Yang, S.; Liu, T.; Zhang, Y.; Yang, K. 131I-APD-L1 Immobilized by Bacterial Cellulose for Enhanced Radio-Immunotherapy of Cancer. J. Control. Release 2022 , 346 , 240–249. [ Google Scholar ] [ CrossRef ]
  • Munawaroh, H.S.H.; Anwar, B.; Yuliani, G.; Murni, I.C.; Arindita, N.P.Y.; Maulidah, G.S.; Martha, L.; Hidayati, N.A.; Chew, K.W.; Show, P.-L. Bacterial Cellulose Nanocrystal as Drug Delivery System for Overcoming the Biological Barrier of Cyano-Phycocyanin: A Biomedical Application of Microbial Product. Bioengineered 2023 , 14 , 2252226. [ Google Scholar ] [ CrossRef ]
  • Estevinho, B.N.; Samaniego, N.; Talens-Perales, D.; Fabra, M.J.; López-Rubio, A.; Polaina, J.; Marín-Navarro, J. Development of Enzymatically-Active Bacterial Cellulose Membranes through Stable Immobilization of an Engineered β-Galactosidase. Int. J. Biol. Macromol. 2018 , 115 , 476–482. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Cacicedo, M.L.; Castro, M.C.; Servetas, I.; Bosnea, L.; Boura, K.; Tsafrakidou, P.; Dima, A.; Terpou, A.; Koutinas, A.; Castro, G.R. Progress in Bacterial Cellulose Matrices for Biotechnological Applications. Bioresour. Technol. 2016 , 213 , 172–180. [ Google Scholar ] [ CrossRef ]
  • Fernandez-Lorente, G.; Lopez-Gallego, F.; Bolivar, J.M.; Rocha-Martin, J.; Moreno-Perez, S.; Guisan, J.M. Immobilization of Proteins on Glyoxyl Activated Supports: Dramatic Stabilization of Enzymes by Multipoint Covalent Attachment on Pre-Existing Supports. Curr. Org. Chem. 2015 , 19 , 1–13. [ Google Scholar ] [ CrossRef ]
  • Drozd, R.; Szymańska, M.; Rakoczy, R.; Junka, A.; Szymczyk, P.; Fijałkowski, K. Functionalized Magnetic Bacterial Cellulose Beads as Carrier for Lecitase® Ultra Immobilization. Appl. Biochem. Biotechnol. 2019 , 187 , 176–193. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Brisola, J.; Andrade, G.J.S.; de Oliveira, S.A.; Viana, R.M.R.; Tischer, P.C.d.S.F.; Tischer, C.A. Covalent Immobilization of Lipase on Bacterial Cellulose Membrane and Nanocellulose. Mater. Res. 2022 , 25 , e20210350. [ Google Scholar ] [ CrossRef ]
  • Kim, H.J.; Jin, J.N.; Kan, E.; Kim, K.J.; Lee, S.H. Bacterial Cellulose-Chitosan Composite Hydrogel Beads for Enzyme Immobilization. Biotechnol. Bioprocess Eng. 2017 , 22 , 89–94. [ Google Scholar ] [ CrossRef ]
  • Cai, Q.; Hu, C.; Yang, N.; Wang, Q.; Wang, J.; Pan, H.; Hu, Y.; Ruan, C. Enhanced Activity and Stability of Industrial Lipases Immobilized onto Spherelike Bacterial Cellulose. Int. J. Biol. Macromol. 2018 , 109 , 1174–1181. [ Google Scholar ] [ CrossRef ]
  • Yu, B.; Cheng, H.; Zhuang, W.; Zhu, C.; Wu, J.; Niu, H.; Liu, D.; Chen, Y.; Ying, H. Stability and Repeatability Improvement of Horseradish Peroxidase by Immobilization on Amino-Functionalized Bacterial Cellulose. Process Biochem. 2019 , 79 , 40–48. [ Google Scholar ] [ CrossRef ]
  • Li, G.; Nandgaonkar, A.G.; Wang, Q.; Zhang, J.; Krause, W.E.; Wei, Q.; Lucia, L.A. Laccase-Immobilized Bacterial Cellulose/TiO 2 Functionalized Composite Membranes: Evaluation for Photo- and Bio-Catalytic Dye Degradation. J. Membr. Sci. 2017 , 525 , 89–98. [ Google Scholar ] [ CrossRef ]
  • Pesaran, M.; Amoabediny, G.; Yazdian, F. Effect of Cultivation Time and Medium Condition in Production of Bacterial Cellulose Nanofiber for Urease Immobilization. Int. J. Polym. Sci. 2015 , 2015 , 1–8. [ Google Scholar ] [ CrossRef ]
  • Shishparenok, A.N.; Koroleva, S.A.; Dobryakova, N.V.; Gladilina, Y.A.; Gromovykh, T.I.; Solopov, A.B.; Kudryashova, E.V.; Zhdanov, D.D. Bacterial Cellulose Films for L-Asparaginase Delivery to Melanoma Cells. Int. J. Biol. Macromol. 2024 , 276 , 133932. [ Google Scholar ] [ CrossRef ]
  • Gupta, D.M.G. Polymer-Based Strategies for Enzyme Immobilization: A Comprehensive Review. Tuijin Jishu/J. Propuls. Technol. 2023 , 44 , 792–802. [ Google Scholar ] [ CrossRef ]
  • Akduman, B.; Uygun, M.; Çoban, E.P.; Uygun, D.A.; Bıyık, H.; Akgöl, S. Reversible Immobilization of Urease by Using Bacterial Cellulose Nanofibers. Appl. Biochem. Biotechnol. 2013 , 171 , 2285–2294. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Drozd, R.; Szymańska, M.; Przygrodzka, K.; Hoppe, J.; Leniec, G.; Kowalska, U. The Simple Method of Preparation of Highly Carboxylated Bacterial Cellulose with Ni- and Mg-Ferrite-Based Versatile Magnetic Carrier for Enzyme Immobilization. Int. J. Mol. Sci. 2021 , 22 , 8563. [ Google Scholar ] [ CrossRef ]
  • Pandey, D.; Daverey, A.; Arunachalam, K. Biochar: Production, Properties and Emerging Role as a Support for Enzyme Immobilization. J. Clean. Prod. 2020 , 255 , 120267. [ Google Scholar ] [ CrossRef ]
  • Overview of Crosslinking and Protein Modification. Available online: https://www.thermofisher.com/ru/ru/home/life-science/protein-biology/protein-biology-learning-center/protein-biology-resource-library/pierce-protein-methods/overview-crosslinking-protein-modification.html#3 (accessed on 7 June 2024).
  • Jackson, E.; Correa, S.; Betancor, L. Cellulose-Based Nanosupports for Enzyme Immobilization. In Cellulose-Based Superabsorbent Hydrogels. Polymers and Polymeric Composites: A Reference Series ; Springer: Cham, Switzerland, 2019; pp. 1235–1253. [ Google Scholar ]
  • Khozeymeh Nezhad, M.; Aghaei, H. Tosylated Cloisite as a New Heterofunctional Carrier for Covalent Immobilization of Lipase and Its Utilization for Production of Biodiesel from Waste Frying Oil. Renew. Energy 2021 , 164 , 876–888. [ Google Scholar ] [ CrossRef ]
  • Dutta, K.; Hu, D.; Zhao, B.; Ribbe, A.E.; Zhuang, J.; Thayumanavan, S. Templated Self-Assembly of a Covalent Polymer Network for Intracellular Protein Delivery and Traceless Release. J. Am. Chem. Soc. 2017 , 139 , 5676–5679. [ Google Scholar ] [ CrossRef ]
  • Bora, U.; Kannan, K.; Nahar, P. A Simple Method for Functionalization of Cellulose Membrane for Covalent Immobilization of Biomolecules. J. Membr. Sci. 2005 , 250 , 215–222. [ Google Scholar ] [ CrossRef ]
  • Schmidt, M.; Abdul Latif, A.; Prager, A.; Gläser, R.; Schulze, A. Highly Efficient One-Step Protein Immobilization on Polymer Membranes Supported by Response Surface Methodology. Front. Chem. 2022 , 9 , 804698. [ Google Scholar ] [ CrossRef ]
  • Birkheur, S.; Laureto, E.; Fernandes, R.V.; Tischer, C.; Butera, A.P.; Ribeiro-Viana, R.M. Diphenyltetrazole Modified Bacterial Cellulose Film: Considerations on Heterogeneous Modification and Bioconjugation. Mater. Res. 2020 , 23 , e20200197. [ Google Scholar ] [ CrossRef ]
  • Vandamme, E.J.; De Baets, S.; Vanbaelen, A.; Joris, K.; De Wulf, P. Improved Production of Bacterial Cellulose and Its Application Potential. Polym. Degrad. Stab. 1998 , 59 , 93–99. [ Google Scholar ] [ CrossRef ]
  • Chiaoprakobkij, N.; Seetabhawang, S.; Okhawilai, M.; Uyama, H.; Phisalaphong, M. Multifunctional Bacterial Cellulose-Gelatin Containing Mangosteen Extract Films with Improved Antibacterial and Anticancer Properties. Cellulose 2022 , 29 , 6811–6830. [ Google Scholar ] [ CrossRef ]
  • Stumpf, T.R.; Yang, X.; Zhang, J.; Cao, X. In Situ and Ex Situ Modifications of Bacterial Cellulose for Applications in Tissue Engineering. Mater. Sci. Eng. C 2018 , 82 , 372–383. [ Google Scholar ] [ CrossRef ]
  • Pang, M.; Huang, Y.; Meng, F.; Zhuang, Y.; Liu, H.; Du, M.; Ma, Q.; Wang, Q.; Chen, Z.; Chen, L.; et al. Application of Bacterial Cellulose in Skin and Bone Tissue Engineering. Eur. Polym. J. 2020 , 122 , 109365. [ Google Scholar ] [ CrossRef ]
  • Katepetch, C.; Rujiravanit, R. Synthesis of Magnetic Nanoparticle into Bacterial Cellulose Matrix by Ammonia Gas-Enhancing in Situ Co-Precipitation Method. Carbohydr. Polym. 2011 , 86 , 162–170. [ Google Scholar ] [ CrossRef ]
  • Yano, S.; Maeda, H.; Nakajima, M.; Hagiwara, T.; Sawaguchi, T. Preparation and Mechanical Properties of Bacterial Cellulose Nanocomposites Loaded with Silica Nanoparticles. Cellulose 2008 , 15 , 111–120. [ Google Scholar ] [ CrossRef ]
  • Novogrodsky, A. Induction of Lymphocyte Cytotoxicity by Modification of the Effector or Target Cells with Periodate or with Neuraminidase and Galactose Oxidase. J. Immunol. 1975 , 114 , 1089–1093. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Badshah, M.; Ullah, H.; Khan, A.R.; Khan, S.; Park, J.K.; Khan, T. Surface Modification and Evaluation of Bacterial Cellulose for Drug Delivery. Int. J. Biol. Macromol. 2018 , 113 , 526–533. [ Google Scholar ] [ CrossRef ]
  • Rouabhia, M.; Asselin, J.; Tazi, N.; Messaddeq, Y.; Levinson, D.; Zhang, Z. Production of Biocompatible and Antimicrobial Bacterial Cellulose Polymers Functionalized by RGDC Grafting Groups and Gentamicin. ACS Appl. Mater. Interfaces 2014 , 6 , 1439–1446. [ Google Scholar ] [ CrossRef ]
  • Lin, Q.; Zheng, Y.; Wang, G.; Shi, X.; Zhang, T.; Yu, J.; Sun, J. Protein Adsorption Behaviors of Carboxymethylated Bacterial Cellulose Membranes. Int. J. Biol. Macromol. 2015 , 73 , 264–269. [ Google Scholar ] [ CrossRef ]
  • Oshima, T.; Kondo, K.; Ohto, K.; Inoue, K.; Baba, Y. Preparation of Phosphorylated Bacterial Cellulose as an Adsorbent for Metal Ions. React. Funct. Polym. 2008 , 68 , 376–383. [ Google Scholar ] [ CrossRef ]
  • Azarniya, A.; Tamjid, E.; Eslahi, N.; Simchi, A. Modification of Bacterial Cellulose/Keratin Nanofibrous Mats by a Tragacanth Gum-Conjugated Hydrogel for Wound Healing. Int. J. Biol. Macromol. 2019 , 134 , 280–289. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Kim, D.-Y.; Nishiyama, Y.; Kuga, S. Surface Acetylation of Bacterial Cellulose. Cellulose 2002 , 9 , 361–367. [ Google Scholar ] [ CrossRef ]
  • Shen, H.; Liao, S.; Jiang, C.; Zhang, J.; Wei, Q.; Ghiladi, R.A.; Wang, Q. In Situ Grown Bacterial Cellulose/MoS2 Composites for Multi-Contaminant Wastewater Treatment and Bacteria Inactivation. Carbohydr. Polym. 2022 , 277 , 118853. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Potočnik, V.; Gorgieva, S.; Trček, J. From Nature to Lab: Sustainable Bacterial Cellulose Production and Modification with Synthetic Biology. Polymers 2023 , 15 , 3466. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Serafica, G.; Mormino, R.; Bungay, H. Inclusion of Solid Particles in Bacterial Cellulose. Appl. Microbiol. Biotechnol. 2002 , 58 , 756–760. [ Google Scholar ] [ CrossRef ]
  • Cheng, K.-C.; Catchmark, J.M.; Demirci, A. Enhanced Production of Bacterial Cellulose by Using a Biofilm Reactor and Its Material Property Analysis. J. Biol. Eng. 2009 , 3 , 12. [ Google Scholar ] [ CrossRef ]
  • Ul-Islam, M.; Khan, T.; Park, J.K. Water Holding and Release Properties of Bacterial Cellulose Obtained by in Situ and Ex Situ Modification. Carbohydr. Polym. 2012 , 88 , 596–603. [ Google Scholar ] [ CrossRef ]
  • Seifert, M.; Hesse, S.; Kabrelian, V.; Klemm, D. Controlling the Water Content of Never Dried and Reswollen Bacterial Cellulose by the Addition of Water-soluble Polymers to the Culture Medium. J. Polym. Sci. Part A Polym. Chem. 2004 , 42 , 463–470. [ Google Scholar ] [ CrossRef ]
  • Gorgieva, S. Bacterial Cellulose as a Versatile Platform for Research and Development of Biomedical Materials. Processes 2020 , 8 , 624. [ Google Scholar ] [ CrossRef ]
  • Kim, G.-D.; Yang, H.; Park, H.R.; Park, C.-S.; Park, Y.S.; Lee, S.E. Evaluation of Immunoreactivity of in Vitro and in Vivo Models against Bacterial Synthesized Cellulose to Be Used as a Prosthetic Biomaterial. BioChip J. 2013 , 7 , 201–209. [ Google Scholar ] [ CrossRef ]
  • Czaja, W.; Krystynowicz, A.; Bielecki, S.; Brownjr, R. Microbial Cellulose—The Natural Power to Heal Wounds. Biomaterials 2006 , 27 , 145–151. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Fürsatz, M.; Skog, M.; Sivlér, P.; Palm, E.; Aronsson, C.; Skallberg, A.; Greczynski, G.; Khalaf, H.; Bengtsson, T.; Aili, D. Functionalization of Bacterial Cellulose Wound Dressings with the Antimicrobial Peptide ε -Poly-L-Lysine. Biomed. Mater. 2018 , 13 , 025014. [ Google Scholar ] [ CrossRef ]
  • Shah, J.L.; Li, G.; Shaffer, J.L.; Azoulay, M.I.; Gibbs, I.C.; Nagpal, S.; Soltys, S.G. Stereotactic Radiosurgery and Hypofractionated Radiotherapy for Glioblastoma. Neurosurgery 2018 , 82 , 24–34. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Autier, L.; Clavreul, A.; Cacicedo, M.L.; Franconi, F.; Sindji, L.; Rousseau, A.; Perrot, R.; Montero-Menei, C.N.; Castro, G.R.; Menei, P. A New Glioblastoma Cell Trap for Implantation after Surgical Resection. Acta Biomater. 2019 , 84 , 268–279. [ Google Scholar ] [ CrossRef ]
  • Andrade, F.K.; Morais, J.P.S.; Muniz, C.R.; Nascimento, J.H.O.; Vieira, R.S.; Gama, F.M.P.; Rosa, M.F. Stable Microfluidized Bacterial Cellulose Suspension. Cellulose 2019 , 26 , 5851–5864. [ Google Scholar ] [ CrossRef ]
  • Liyaskina, E.; Revin, V.; Paramonova, E.; Nazarkina, M.; Pestov, N.; Revina, N.; Kolesnikova, S. Nanomaterials from Bacterial Cellulose for Antimicrobial Wound Dressing. J. Phys. Conf. Ser. 2017 , 784 , 012034. [ Google Scholar ] [ CrossRef ]
  • Zhang, X.; Wang, D.; Liu, S.; Tang, J. Bacterial Cellulose Nanofibril-Based Pickering Emulsions: Recent Trends and Applications in the Food Industry. Foods 2022 , 11 , 4064. [ Google Scholar ] [ CrossRef ]
  • Singhsa, P.; Narain, R.; Manuspiya, H. Bacterial Cellulose Nanocrystals (BCNC) Preparation and Characterization from Three Bacterial Cellulose Sources and Development of Functionalized BCNCs as Nucleic Acid Delivery Systems. ACS Appl. Nano Mater. 2018 , 1 , 209–221. [ Google Scholar ] [ CrossRef ]
  • Li, Q.; Wang, Y.; Wu, Y.; He, K.; Li, Y.; Luo, X.; Li, B.; Wang, C.; Liu, S. Flexible Cellulose Nanofibrils as Novel Pickering Stabilizers: The Emulsifying Property and Packing Behavior. Food Hydrocoll. 2019 , 88 , 180–189. [ Google Scholar ] [ CrossRef ]
  • Liu, W.; Du, H.; Zhang, M.; Liu, K.; Liu, H.; Xie, H.; Zhang, X.; Si, C. Bacterial Cellulose-Based Composite Scaffolds for Biomedical Applications: A Review. ACS Sustain. Chem. Eng. 2020 , 8 , 7536–7562. [ Google Scholar ] [ CrossRef ]
  • Balea, A.; Fuente, E.; Monte, M.C.; Merayo, N.; Campano, C.; Negro, C.; Blanco, A. Industrial Application of Nanocelluloses in Papermaking: A Review of Challenges, Technical Solutions, and Market Perspectives. Molecules 2020 , 25 , 526. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Rol, F.; Belgacem, M.N.; Gandini, A.; Bras, J. Recent Advances in Surface-Modified Cellulose Nanofibrils. Prog. Polym. Sci. 2019 , 88 , 241–264. [ Google Scholar ] [ CrossRef ]
  • Li, Q.; Chen, P.; Li, Y.; Li, B.; Liu, S. Construction of Cellulose-Based Pickering Stabilizer as a Novel Interfacial Antioxidant: A Bioinspired Oxygen Protection Strategy. Carbohydr. Polym. 2020 , 229 , 115395. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Vasconcelos, N.F.; Feitosa, J.P.A.; da Gama, F.M.P.; Morais, J.P.S.; Andrade, F.K.; de Souza Filho, M.d.S.M.; Rosa, M.d.F. Bacterial Cellulose Nanocrystals Produced under Different Hydrolysis Conditions: Properties and Morphological Features. Carbohydr. Polym. 2017 , 155 , 425–431. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Ribeiro-Viana, R.M.; Faria-Tischer, P.C.S.; Tischer, C.A. Preparation of Succinylated Cellulose Membranes for Functionalization Purposes. Carbohydr. Polym. 2016 , 148 , 21–28. [ Google Scholar ] [ CrossRef ]
  • Oryan, A.; Kamali, A.; Moshiri, A.; Baharvand, H.; Daemi, H. Chemical Crosslinking of Biopolymeric Scaffolds: Current Knowledge and Future Directions of Crosslinked Engineered Bone Scaffolds. Int. J. Biol. Macromol. 2018 , 107 , 678–688. [ Google Scholar ] [ CrossRef ]
  • Heck, T.; Faccio, G.; Richter, M.; Thöny-Meyer, L. Enzyme-Catalyzed Protein Crosslinking. Appl. Microbiol. Biotechnol. 2013 , 97 , 461–475. [ Google Scholar ] [ CrossRef ]
  • Chanabodeechalermrung, B.; Chaiwarit, T.; Sommano, S.R.; Rachtanapun, P.; Kantrong, N.; Chittasupho, C.; Jantrawut, P. Dual Crosslinked Ion-Based Bacterial Cellulose Composite Hydrogel Containing Polyhexamethylene Biguanide. Membranes 2022 , 12 , 825. [ Google Scholar ] [ CrossRef ]
  • Shen, X.; Shamshina, J.L.; Berton, P.; Gurau, G.; Rogers, R.D. Hydrogels Based on Cellulose and Chitin: Fabrication, Properties, and Applications. Green Chem. 2016 , 18 , 53–75. [ Google Scholar ] [ CrossRef ]
  • Wang, J.; Wan, Y.Z.; Luo, H.L.; Gao, C.; Huang, Y. Immobilization of Gelatin on Bacterial Cellulose Nanofibers Surface via Crosslinking Technique. Mater. Sci. Eng. C 2012 , 32 , 536–541. [ Google Scholar ] [ CrossRef ]
  • Castro, C.; Cordeiro, N.; Faria, M.; Zuluaga, R.; Putaux, J.-L.; Filpponen, I.; Velez, L.; Rojas, O.J.; Gañán, P. In-Situ Glyoxalization during Biosynthesis of Bacterial Cellulose. Carbohydr. Polym. 2015 , 126 , 32–39. [ Google Scholar ] [ CrossRef ]
  • Chaiyasat, A.; Jearanai, S.; Moonmangmee, S.; Moonmangmee, D.; P Christopher, L.; Nur Alam, M.; Chaiyasat, P. Novel Green Hydrogel Material Using Bacterial Cellulose. Orient. J. Chem. 2018 , 34 , 1735–1740. [ Google Scholar ] [ CrossRef ]
  • Meftahi, A.; Khajavi, R.; Rashidi, A.; Rahimi, M.K.; Bahador, A. Preventing the Collapse of 3D Bacterial Cellulose Network via Citric Acid. J. Nanostruct. Chem. 2018 , 8 , 311–320. [ Google Scholar ] [ CrossRef ]
  • Sommer, A.; Dederko-Kantowicz, P.; Staroszczyk, H.; Sommer, S.; Michalec, M. Enzymatic and Chemical Cross-Linking of Bacterial Cellulose/Fish Collagen Composites—A Comparative Study. Int. J. Mol. Sci. 2021 , 22 , 3346. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Hubbe, M.A. Review of the Mechanistic Roles of Nanocellulose, Cellulosic Fibers, and Hydrophilic Cellulose Derivatives in Cellulose-Based Absorbents. In Cellulose-Based Superabsorbent Hydrogels ; Springer: Cham, Switzerland, 2019; pp. 123–153. [ Google Scholar ]
  • Pal, K.; Paulson, A.T.; Rousseau, D. Biopolymers in Controlled-Release Delivery Systems. In Handbook of Biopolymers and Biodegradable Plastics ; Elsevier: Amsterdam, The Netherlands, 2013; pp. 329–363. [ Google Scholar ]
  • Bao, L.; Li, C.; Tang, M.; Chen, L.; Hong, F.F. Potential of a Composite Conduit with Bacterial Nanocellulose and Fish Gelatin for Application as Small-Diameter Artificial Blood Vessel. Polymers 2022 , 14 , 4367. [ Google Scholar ] [ CrossRef ]
  • Delgado, L.M.; Bayon, Y.; Pandit, A.; Zeugolis, D.I. To Cross-Link or Not to Cross-Link? Cross-Linking Associated Foreign Body Response of Collagen-Based Devices. Tissue Eng. Part B Rev. 2015 , 21 , 298–313. [ Google Scholar ] [ CrossRef ]
  • Treesuppharat, W.; Rojanapanthu, P.; Siangsanoh, C.; Manuspiya, H.; Ummartyotin, S. Synthesis and Characterization of Bacterial Cellulose and Gelatin-Based Hydrogel Composites for Drug-Delivery Systems. Biotechnol. Rep. 2017 , 15 , 84–91. [ Google Scholar ] [ CrossRef ]
  • Brown, E.E.; Laborie, M.-P.G.; Zhang, J. Glutaraldehyde Treatment of Bacterial Cellulose/Fibrin Composites: Impact on Morphology, Tensile and Viscoelastic Properties. Cellulose 2012 , 19 , 127–137. [ Google Scholar ] [ CrossRef ]
  • Lees, A.; Nelson, B.L.; Mond, J.J. Activation of Soluble Polysaccharides with 1-Cyano-4-Dimethylaminopyridinium Tetrafluoroborate for Use in Protein—Polysaccharide Conjugate Vaccines and Immunological Reagents. Vaccine 1996 , 14 , 190–198. [ Google Scholar ] [ CrossRef ]
  • Mond, J.J.; Lees, A.; Snapper, C.M. T Cell-Independent Antigens Type 2. Annu. Rev. Immunol. 1995 , 13 , 655–692. [ Google Scholar ] [ CrossRef ]
  • Shafer, D.E.; Toll, B.; Schuman, R.F.; Nelson, B.L.; Mond, J.J.; Lees, A. Activation of Soluble Polysaccharides with 1-Cyano-4-Dimethylaminopyridinium Tetrafluoroborate (CDAP) for Use in Protein-Polysaccharide Conjugate Vaccines and Immunological Reagents. II. Selective Crosslinking of Proteins to CDAP-Activated Polysaccharides. Vaccine 2000 , 18 , 1273–1281. [ Google Scholar ] [ CrossRef ]
  • Wacker, M.; Riedel, J.; Walles, H.; Scherner, M.; Awad, G.; Varghese, S.; Schürlein, S.; Garke, B.; Veluswamy, P.; Wippermann, J.; et al. Comparative Evaluation on Impacts of Fibronectin, Heparin–Chitosan, and Albumin Coating of Bacterial Nanocellulose Small-Diameter Vascular Grafts on Endothelialization In Vitro. Nanomaterials 2021 , 11 , 1952. [ Google Scholar ] [ CrossRef ]
  • Akhtar, M.F.; Hanif, M.; Ranjha, N.M. Methods of Synthesis of Hydrogels … A Review. Saudi Pharm. J. 2016 , 24 , 554–559. [ Google Scholar ] [ CrossRef ]
  • Mohite, P.B.; Adhav, S.S. A Hydrogels: Methods of Preparation and Applications. Int. J. Adv. Pharm. 2017 , 06 , 79–85. [ Google Scholar ]
  • Alven, S.; Aderibigbe, B.A. Chitosan and Cellulose-Based Hydrogels for Wound Management. Int. J. Mol. Sci. 2020 , 21 , 9656. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Mohd Amin, M.C.I.; Ahmad, N.; Halib, N.; Ahmad, I. Synthesis and Characterization of Thermo- and PH-Responsive Bacterial Cellulose/Acrylic Acid Hydrogels for Drug Delivery. Carbohydr. Polym. 2012 , 88 , 465–473. [ Google Scholar ] [ CrossRef ]
  • Jing, W.; Chunxi, Y.; Yizao, W.; Honglin, L.; Fang, H.; Kerong, D.; Yuan, H. Laser Patterning of Bacterial Cellulose Hydrogel and Its Modification With Gelatin and Hydroxyapatite for Bone Tissue Engineering. Soft Mater. 2013 , 11 , 173–180. [ Google Scholar ] [ CrossRef ]
  • Mohamad, N.; Mohd Amin, M.C.I.; Pandey, M.; Ahmad, N.; Rajab, N.F. Bacterial Cellulose/Acrylic Acid Hydrogel Synthesized via Electron Beam Irradiation: Accelerated Burn Wound Healing in an Animal Model. Carbohydr. Polym. 2014 , 114 , 312–320. [ Google Scholar ] [ CrossRef ]
  • Ahmad, N.; Amin, M.C.I.M.; Mahali, S.M.; Ismail, I.; Chuang, V.T.G. Biocompatible and Mucoadhesive Bacterial Cellulose -g- Poly(Acrylic Acid) Hydrogels for Oral Protein Delivery. Mol. Pharm. 2014 , 11 , 4130–4142. [ Google Scholar ] [ CrossRef ]
  • Lai, C.; Zhang, S.; Chen, X.; Sheng, L. Nanocomposite Films Based on TEMPO-Mediated Oxidized Bacterial Cellulose and Chitosan. Cellulose 2014 , 21 , 2757–2772. [ Google Scholar ] [ CrossRef ]
  • Credou, J.; Berthelot, T. Cellulose: From Biocompatible to Bioactive Material. J. Mater. Chem. B 2014 , 2 , 4767–4788. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Cui, Q.; Zheng, Y.; Lin, Q.; Song, W.; Qiao, K.; Liu, S. Selective Oxidation of Bacterial Cellulose by NO 2 –HNO 3 . RSC Adv. 2014 , 4 , 1630–1639. [ Google Scholar ] [ CrossRef ]
  • Zheng, Y.; Wen, X.; Wu, J.; Wang, L.-N.; Yuan, Z.; Peng, J.; Meng, H. Immobilization of Collagen Peptide on Dialdehyde Bacterial Cellulose Nanofibers via Covalent Bonds for Tissue Engineering and Regeneration. Int. J. Nanomed. 2015 , 10 , 4623–4637. [ Google Scholar ] [ CrossRef ]
  • Luz, E.P.C.G.; Chaves, P.H.S.; Vieira, L.d.A.P.; Ribeiro, S.F.; Borges, M.d.F.; Andrade, F.K.; Muniz, C.R.; Infantes-Molina, A.; Rodríguez-Castellón, E.; Rosa, M.d.F.; et al. In Vitro Degradability and Bioactivity of Oxidized Bacterial Cellulose-Hydroxyapatite Composites. Carbohydr. Polym. 2020 , 237 , 116174. [ Google Scholar ] [ CrossRef ]
  • Li, J.; Wan, Y.; Li, L.; Liang, H.; Wang, J. Preparation and Characterization of 2,3-Dialdehyde Bacterial Cellulose for Potential Biodegradable Tissue Engineering Scaffolds. Mater. Sci. Eng. C 2009 , 29 , 1635–1642. [ Google Scholar ] [ CrossRef ]
  • Zhang, W.; Wang, X.; Li, X.; Zhang, L.; Jiang, F. A 3D Porous Microsphere with Multistage Structure and Component Based on Bacterial Cellulose and Collagen for Bone Tissue Engineering. Carbohydr. Polym. 2020 , 236 , 116043. [ Google Scholar ] [ CrossRef ]
  • Zeng, N.; Chen, H.; Wu, Y.; Liu, Z. Adipose Stem Cell-Based Treatments for Wound Healing. Front. Cell Dev. Biol. 2022 , 9 , 821652. [ Google Scholar ] [ CrossRef ]
  • Bian, D.; Wu, Y.; Song, G.; Azizi, R.; Zamani, A. The Application of Mesenchymal Stromal Cells (MSCs) and Their Derivative Exosome in Skin Wound Healing: A Comprehensive Review. Stem Cell Res. Ther. 2022 , 13 , 24. [ Google Scholar ] [ CrossRef ]
  • Zhu, Z.; Yang, J.; Ji, X.; Wang, Z.; Dai, C.; Li, S.; Li, X.; Xie, Y.; Zheng, Y.; Lin, J.; et al. Clinical Application of a Double-Modified Sulfated Bacterial Cellulose Scaffold Material Loaded with FGFR2-Modified Adipose-Derived Stem Cells in Urethral Reconstruction. Stem Cell Res. Ther. 2022 , 13 , 463. [ Google Scholar ] [ CrossRef ]
  • Jia, Y.; Zhai, X.; Fu, W.; Liu, Y.; Li, F.; Zhong, C. Surfactant-Free Emulsions Stabilized by Tempo-Oxidized Bacterial Cellulose. Carbohydr. Polym. 2016 , 151 , 907–915. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • do Nascimento, E.S.; Pereira, A.L.S.; Barros, M.d.O.; Barroso, M.K.d.A.; Lima, H.L.S.; Borges, M.d.F.; Feitosa, J.P.d.A.; de Azeredo, H.M.C.; Rosa, M.d.F. TEMPO Oxidation and High-Speed Blending as a Combined Approach to Disassemble Bacterial Cellulose. Cellulose 2019 , 26 , 2291–2302. [ Google Scholar ] [ CrossRef ]
  • Khattak, S.; Qin, X.-T.; Wahid, F.; Huang, L.-H.; Xie, Y.-Y.; Jia, S.-R.; Zhong, C. Permeation of Silver Sulfadiazine Into TEMPO-Oxidized Bacterial Cellulose as an Antibacterial Agent. Front. Bioeng. Biotechnol. 2021 , 8 , 616467. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Koshy, T.M.; Gowda, D.V.; Tom, S.; Karunakar, G.; Srivastava, A.; Moin, A. Polymer Grafting-An Overview. Am. J. PharmTech Res. 2016 , 6 , 1–13. [ Google Scholar ]
  • Stepanova, M.; Korzhikova-Vlakh, E. Modification of Cellulose Micro- and Nanomaterials to Improve Properties of Aliphatic Polyesters/Cellulose Composites: A Review. Polymers 2022 , 14 , 1477. [ Google Scholar ] [ CrossRef ]
  • Hamedi, S.; Shojaosadati, S.A.; Najafi, V.; Alizadeh, V. A Novel Double-Network Antibacterial Hydrogel Based on Aminated Bacterial Cellulose and Schizophyllan. Carbohydr. Polym. 2020 , 229 , 115383. [ Google Scholar ] [ CrossRef ]
  • Hettegger, H.; Sumerskii, I.; Sortino, S.; Potthast, A.; Rosenau, T. Silane Meets Click Chemistry: Towards the Functionalization of Wet Bacterial Cellulose Sheets. ChemSusChem 2015 , 8 , 680–687. [ Google Scholar ] [ CrossRef ]
  • Patoary, M.K.; Islam, S.R.; Farooq, A.; Rashid, M.A.; Sarker, S.; Hossain, M.Y.; Rakib, M.A.N.; Al-Amin, M.; Liu, L. Phosphorylation of Nanocellulose: State of the Art and Prospects. Ind. Crops Prod. 2023 , 201 , 116965. [ Google Scholar ] [ CrossRef ]
  • Oshima, T.; Taguchi, S.; Ohe, K.; Baba, Y. Phosphorylated Bacterial Cellulose for Adsorption of Proteins. Carbohydr. Polym. 2011 , 83 , 953–958. [ Google Scholar ] [ CrossRef ]
  • Müller, D.; Mandelli, J.S.; Marins, J.A.; Soares, B.G.; Porto, L.M.; Rambo, C.R.; Barra, G.M.O. Electrically Conducting Nanocomposites: Preparation and Properties of Polyaniline (PAni)-Coated Bacterial Cellulose Nanofibers (BC). Cellulose 2012 , 19 , 1645–1654. [ Google Scholar ] [ CrossRef ]
  • Luo, H.; Xiong, G.; Huang, Y.; He, F.; Wang, Y.; Wan, Y. Preparation and Characterization of a Novel COL/BC Composite for Potential Tissue Engineering Scaffolds. Mater. Chem. Phys. 2008 , 110 , 193–196. [ Google Scholar ] [ CrossRef ]
  • Gao, G.; Cao, Y.; Zhang, Y.; Wu, M.; Ma, T.; Li, G. In Situ Production of Bacterial Cellulose/Xanthan Gum Nanocomposites with Enhanced Productivity and Properties Using Enterobacter Sp. FY-07. Carbohydr. Polym. 2020 , 248 , 116788. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Cheng, K.-C.; Catchmark, J.M.; Demirci, A. Effect of Different Additives on Bacterial Cellulose Production by Acetobacter Xylinum and Analysis of Material Property. Cellulose 2009 , 16 , 1033–1045. [ Google Scholar ] [ CrossRef ]
  • de Lima Fontes, M.; Meneguin, A.B.; Tercjak, A.; Gutierrez, J.; Cury, B.S.F.; dos Santos, A.M.; Ribeiro, S.J.L.; Barud, H.S. Effect of in Situ Modification of Bacterial Cellulose with Carboxymethylcellulose on Its Nano/Microstructure and Methotrexate Release Properties. Carbohydr. Polym. 2018 , 179 , 126–134. [ Google Scholar ] [ CrossRef ]
  • Orelma, H.; Morales, L.O.; Johansson, L.-S.; Hoeger, I.C.; Filpponen, I.; Castro, C.; Rojas, O.J.; Laine, J. Affibody Conjugation onto Bacterial Cellulose Tubes and Bioseparation of Human Serum Albumin. RSC Adv. 2014 , 4 , 51440–51450. [ Google Scholar ] [ CrossRef ]
  • Jiang, Y.; Yu, G.; Zhou, Y.; Liu, Y.; Feng, Y.; Li, J. Effects of Sodium Alginate on Microstructural and Properties of Bacterial Cellulose Nanocrystal Stabilized Emulsions. Colloids Surf. A Physicochem. Eng. Asp. 2020 , 607 , 125474. [ Google Scholar ] [ CrossRef ]
  • Zhou, L.L.; Sun, D.P.; Hu, L.Y.; Li, Y.W.; Yang, J.Z. Effect of Addition of Sodium Alginate on Bacterial Cellulose Production by Acetobacter xylinum . J. Ind. Microbiol. Biotechnol. 2007 , 34 , 483–489. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Gilmour, K.A.; Aljannat, M.; Markwell, C.; James, P.; Scott, J.; Jiang, Y.; Torun, H.; Dade-Robertson, M.; Zhang, M. Biofilm Inspired Fabrication of Functional Bacterial Cellulose through Ex-Situ and in-Situ Approaches. Carbohydr. Polym. 2023 , 304 , 120482. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Morris, R.J.; Schor, M.; Gillespie, R.M.C.; Ferreira, A.S.; Baldauf, L.; Earl, C.; Ostrowski, A.; Hobley, L.; Bromley, K.M.; Sukhodub, T.; et al. Natural Variations in the Biofilm-Associated Protein BslA from the Genus Bacillus. Sci. Rep. 2017 , 7 , 6730. [ Google Scholar ] [ CrossRef ]
  • Pandey, A.; Singh, M.K.; Singh, A. Bacterial Cellulose: A Smart Biomaterial for Biomedical Applications. J. Mater. Res. 2023 , 145 , 100623. [ Google Scholar ] [ CrossRef ]
  • Shabanpour, B.; Kazemi, M.; Ojagh, S.M.; Pourashouri, P. Bacterial Cellulose Nanofibers as Reinforce in Edible Fish Myofibrillar Protein Nanocomposite Films. Int. J. Biol. Macromol. 2018 , 117 , 742–751. [ Google Scholar ] [ CrossRef ]
  • Ullah, H.; Santos, H.A.; Khan, T. Applications of Bacterial Cellulose in Food, Cosmetics and Drug Delivery. Cellulose 2016 , 23 , 2291–2314. [ Google Scholar ] [ CrossRef ]
  • Müller, A.; Ni, Z.; Hessler, N.; Wesarg, F.; Müller, F.A.; Kralisch, D.; Fischer, D. The Biopolymer Bacterial Nanocellulose as Drug Delivery System: Investigation of Drug Loading and Release Using the Model Protein Albumin. J. Pharm. Sci. 2013 , 102 , 579–592. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Gorgieva, S.; Trček, J. Bacterial Cellulose: Production, Modification and Perspectives in Biomedical Applications. Nanomaterials 2019 , 9 , 1352. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Qiao, N.; Fan, X.; Hu, S.; Zhang, X.; Wang, L.; Du, Y.; Wang, L.; Zhang, X.; Yu, D. Bacterial Cellulose as an Oleaginous Yeast Cell Carrier for Soybean Oil Refinery Effluent Treatment and Pyrolysis Oil Production. Bioprocess Biosyst. Eng. 2021 , 44 , 661–671. [ Google Scholar ] [ CrossRef ]
  • Islam, M.U.; Ullah, M.W.; Khan, S.; Shah, N.; Park, J.K. Strategies for Cost-Effective and Enhanced Production of Bacterial Cellulose. Int. J. Biol. Macromol. 2017 , 102 , 1166–1173. [ Google Scholar ] [ CrossRef ]
  • Hussain, Z.; Sajjad, W.; Khan, T.; Wahid, F. Production of Bacterial Cellulose from Industrial Wastes: A Review. Cellulose 2019 , 26 , 2895–2911. [ Google Scholar ] [ CrossRef ]
  • Avcioglu, N.H. Bacterial Cellulose: Recent Progress in Production and Industrial Applications. World J. Microbiol. Biotechnol. 2022 , 38 , 86. [ Google Scholar ] [ CrossRef ]
  • Aragão, J.V.S.; Costa, A.F.S.; Silva, G.L.; Silva, S.M.; Macêdo, J.S.; Galdino, C.J.S., Jr.; Milanez, V.F.A.; Sarubbo, L.A. Analysis of the Environmental Life Cycle of Bacterial Cellulose Production. Ital. Assoc. Chem. Eng. 2020 , 79 , 445–450. [ Google Scholar ]
  • Weiss, M.; Haufe, J.; Carus, M.; Brandão, M.; Bringezu, S.; Hermann, B.; Patel, M.K. A Review of the Environmental Impacts of Biobased Materials. J. Ind. Ecol. 2012 , 16 , 169–181. [ Google Scholar ] [ CrossRef ]
  • Kamal, T.; Ul-Islam, M.; Fatima, A.; Ullah, M.W.; Manan, S. Cost-Effective Synthesis of Bacterial Cellulose and Its Applications in the Food and Environmental Sectors. Gels 2022 , 8 , 552. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Moradi, M.; Jacek, P.; Farhangfar, A.; Guimarães, J.T.; Forough, M. The Role of Genetic Manipulation and in Situ Modifications on Production of Bacterial Nanocellulose: A Review. Int. J. Biol. Macromol. 2021 , 183 , 635–650. [ Google Scholar ] [ CrossRef ] [ PubMed ]
  • Liu, K.; Catchmark, J.M. Bacterial Cellulose/Hyaluronic Acid Nanocomposites Production through Co-Culturing Gluconacetobacter Hansenii and Lactococcus Lactis in a Two-Vessel Circulating System. Bioresour. Technol. 2019 , 290 , 121715. [ Google Scholar ] [ CrossRef ]
  • Hu, H.; Catchmark, J.M.; Demirci, A. Co-Culture Fermentation on the Production of Bacterial Cellulose Nanocomposite Produced by Komagataeibacter Hansenii. Carbohydr. Polym. Technol. Appl. 2021 , 2 , 100028. [ Google Scholar ] [ CrossRef ]
  • Eslahi, N.; Mahmoodi, A.; Mahmoudi, N.; Zandi, N.; Simchi, A. Processing and Properties of Nanofibrous Bacterial Cellulose-Containing Polymer Composites: A Review of Recent Advances for Biomedical Applications. Polym. Rev. 2020 , 60 , 144–170. [ Google Scholar ] [ CrossRef ]
  • Torres, F.; Commeaux, S.; Troncoso, O. Biocompatibility of Bacterial Cellulose Based Biomaterials. J. Funct. Biomater. 2012 , 3 , 864–878. [ Google Scholar ] [ CrossRef ]
  • Zhong, C. Industrial-Scale Production and Applications of Bacterial Cellulose. Front. Bioeng. Biotechnol. 2020 , 8 , 605374. [ Google Scholar ] [ CrossRef ]

Click here to enlarge figure

The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

Shishparenok, A.N.; Furman, V.V.; Dobryakova, N.V.; Zhdanov, D.D. Protein Immobilization on Bacterial Cellulose for Biomedical Application. Polymers 2024 , 16 , 2468. https://doi.org/10.3390/polym16172468

Shishparenok AN, Furman VV, Dobryakova NV, Zhdanov DD. Protein Immobilization on Bacterial Cellulose for Biomedical Application. Polymers . 2024; 16(17):2468. https://doi.org/10.3390/polym16172468

Shishparenok, Anastasia N., Vitalina V. Furman, Natalia V. Dobryakova, and Dmitry D. Zhdanov. 2024. "Protein Immobilization on Bacterial Cellulose for Biomedical Application" Polymers 16, no. 17: 2468. https://doi.org/10.3390/polym16172468

Article Metrics

Article access statistics, supplementary material.

ZIP-Document (ZIP, 265 KiB)

Further Information

Mdpi initiatives, follow mdpi.

MDPI

Subscribe to receive issue release notifications and newsletters from MDPI journals

Thank you for visiting nature.com. You are using a browser version with limited support for CSS. To obtain the best experience, we recommend you use a more up to date browser (or turn off compatibility mode in Internet Explorer). In the meantime, to ensure continued support, we are displaying the site without styles and JavaScript.

  • View all journals
  • Explore content
  • About the journal
  • Publish with us
  • Sign up for alerts
  • Open access
  • Published: 03 September 2024

Self-assembly of sustainable plant protein protofilaments into a hydrogel for ultra-low friction across length scales

  • Olivia Pabois 1   na1 ,
  • Yihui Dong 2   na1 ,
  • Nir Kampf   ORCID: orcid.org/0000-0002-6713-6979 2 ,
  • Christian D. Lorenz   ORCID: orcid.org/0000-0003-1028-4804 3 ,
  • James Doutch 4 ,
  • Alejandro Avila-Sierra 5 ,
  • Marco Ramaioli 5 ,
  • Mingduo Mu 1 ,
  • Yasmin Message 1 ,
  • Evangelos Liamas 1 , 6 ,
  • Arwen I. I. Tyler   ORCID: orcid.org/0000-0003-2116-1084 1 ,
  • Jacob Klein   ORCID: orcid.org/0000-0001-6602-0694 2 &
  • Anwesha Sarkar   ORCID: orcid.org/0000-0003-1742-2122 1  

Communications Materials volume  5 , Article number:  158 ( 2024 ) Cite this article

23 Altmetric

Metrics details

  • Nanoscale biophysics
  • Self-assembly

Designing plant protein-based aqueous lubricants can be of great potential to achieve sustainability objectives by capitalising on inherent functional groups without using synthetic chemicals; however, such a concept remains in its infancy. Here, we engineer a class of self-assembled sustainable materials by using plant-based protofilaments and their assembly within a biopolymeric hydrogel giving rise to a distinct patchy architecture. By leveraging physical interactions, this material offers superlubricity with friction coefficients of 0.004-to-0.00007 achieved under moderate-to-high (10 2 -to-10 3 kPa) contact pressures. Multiscale experimental measurements combined with molecular dynamics simulations reveal an intriguing synergistic mechanism behind such ultra-low friction - where the uncoated areas of the protofilaments glue to the surface by hydrophobic interactions, whilst the hydrogel offers the hydration lubrication. The current approach establishes a robust platform towards unlocking an untapped potential of using plant protein-based building blocks across diverse applications where achieving superlubricity and environmental sustainability are key performance indicators.

Similar content being viewed by others

journal of biomedical materials research

Controlled self-assembly of plant proteins into high-performance multifunctional nanostructured films

journal of biomedical materials research

Hierarchically-structured metalloprotein composite coatings biofabricated from co-existing condensed liquid phases

journal of biomedical materials research

DRY & WET: meniscus splitting from a mixture of polysaccharides and water

Introduction.

Achieving superlubricity or near-zero friction with ultra-low sliding friction coefficients (<0.01) is a revolutionary engineering paradigm for energy saving 1 , 2 and biomedical applications 3 . Developing functional materials by exploiting the untapped potential of hydration lubrication achieving oil-free superlubricity at moderate-to-high contact pressures mimicking those found in biology (such as mucinous self-assemblies in healthy oral epithelial surfaces 4 or synovial fluids in articulating cartilage surfaces in healthy human joints 5 ) seems to be an obvious alternative for achieving a sustainable future. Although literature on hydration lubrication has surfaced in the past decade, engineering of eco-friendly, efficient, functional aqueous lubricants remains far from realisation. Many, if not most, aqueous lubricants showing superlubricity employ unilamellar vesicles 6 , solid lubricants 7 , polyzwitterionic brushes 8 , or amphiphilic surfactants 9 that are nearly exclusively derived from synthetic chemistry. Although the recent use of hydrogels offers a unique route for achieving ultra-low friction whilst promoting industrial sustainability objectives, often these hydrogels are prepared following synthetic routes and using lipids, polyacrylamides, and other zwitterionic synthetic polymers 10 , 11 , 12 , 13 , 14 . Therefore, seeking an environmentally friendly source of advanced functional lubricants showing superlubricity remains a critical bottleneck.

Using plant-derived proteins as alternative building blocks to synthetic aqueous lubricants could be an ideal way to address this sustainability issue as they can be naturally sourced in abundance with lower carbon footprints 15 . Although plant proteins suffer from functionality issues owing to their complex quaternary structure and limited solubility, there has been an increased momentum to convert them into functional materials such as microgels, nanostructured films, and amyloid fibrils (fabricated by us 16 and other groups 17 , 18 ) exploiting hydrophobic and coulombic interactions manipulating the physical ordering of their naturally occurring functional amino acids. Nevertheless, converting these sustainable plant proteins into sustainable, functional aqueous lubricant materials achieving superlubricity remains to be demonstrated.

In light of these considerations, the present study reports the first design of a self-assembly of plant protein-derived protofilaments within a biopolymeric hydrogel network as an advanced functional aqueous lubricant material offering superlubricity across length scales. Particularly, we use potato protein, which is a highly biocompatible, non-allergenic plant protein derived from the starch industry as a by-product 19 , to create such protofilaments using a highly facile physical crosslinking method, and then electrostatically assemble them with highly hydrating biopolymeric hydrogels. Striking results from this study demonstrate that a robust liquid-vanishing friction with friction coefficients of ca. 4 × 10 −3 to 7 × 10 −5 ) can be achieved under moderate-to-high contact pressures of 300 kPa to 3 MPa due to the synergistic action of the protofilaments offering the surface-anchoring properties, and the hydrogel providing the water mesh-induced hydration, thus behaving like a polymer brush, but made using ecofriendly materials. Such outstanding lubrication properties are due to the ‘patchy architecture’ of this highly viscoelastic and extensible self-assembled structure, where the areas of the protofilaments uncoated by the hydrogel are available to glue to the surface.

The molecular structure of the bulk and interfacial tribofilms formed by this unique self-assembly were characterised across different length scales combining a wide array of techniques, including imaging (transmission electron (TEM) and atomic force (AFM) microscopy), and scattering (light and small-angle neutron scattering (SANS)). Molecular dynamics (MD) simulations were performed both in the bulk phase and at the surface to gain further insights into the structural mechanism governing the interaction between both components and with the surface. We also thoroughly assessed the lubricant friction-reducing capability at multiscale, using both mini traction machine (macroscale study) and surface force balance (SFB) (nanoscale study). Complementary experiments aiming to get further insight into the adsorption behaviour and viscoelasticity (upon shear and extensional deformation) of the engineered self-assembled structures were additionally conducted with a quartz-crystal microbalance with dissipation monitoring (QCM-D), as well as with rotational and extensional rheometers, in order to subsequently define the key parameters influencing hydration performance. This biocompatible ecofriendly aqueous lubricant showing superlubricity at multiple length scales is a key milestone towards creating highly sustainable, plant-based aqueous lubricant materials in the energy sector 20 , as well as the next generation of engineered biomedical materials, such as artificial synovial fluid, tear and saliva for lifetime lubrication of natural biological contacts.

Results and discussion

Molecular structure of the self-assembled protofilament-hydrogel.

To comprehensively explore the protofilament-hydrogel assembling mechanism, the molecular architecture of each individual component (i.e., the potato protein protofilament (PoPF) and the hydrogel made using naturally occurring biopolymers (i.e., xanthan gum (XGH) or κ -carrageenan (KCH) hydrogels) and their combination (i.e., 2 XGH/PoPF and 2 KCH/PoPF) was resolved at both the nano- and atomistic scales, by employing a very wide range of structural techniques, including dynamic light scattering (DLS) (Fig.  1a ), transmission electron microscopy (TEM) (Fig.  1b ), atomic force microscopy (AFM) (Fig.  1c, e and Supplementary Fig.  1 ), small-angle neutron scattering (SANS) (Fig.  1d ), and molecular dynamics (MD) simulations (Fig.  2 and Supplementary Figs.  2 – 4 ). To this end, we designed a comprehensive protofilament fabrication process using potato protein and induced their self-assembly within a polysaccharide hydrogel (either XGH or KCH)) (see ‘Experimental’ Section). The characterisation of both the bulk (DLS, SANS, and MD simulations) and interfacial film (TEM, AFM, and MD simulations) structures formed by the protofilament-hydrogel systems ultimately enabled shedding light on the molecular process governing their superlubricity. Whilst we mainly focused on using XGH as the hydrogel component, we also outline the importance of the polymer type by pinpointing the specific differences observed using another polysaccharide hydrogel (KCH) where appropriate (Supplementary Fig.  1 ).

figure 1

a Particle size distribution of potato protein protofilament (PoPF) obtained from dynamic light scattering (DLS) measurements. PoPF displays a hydrodynamic diameter of ca. d H  = 100 nm and a surface charge of ζ = +30 mV. PoPF nanostructure observed by ( b ) transmission electron microscopy (TEM), following negative staining, and ( c ) atomic force microscopy (AFM), following sample deposition onto a negatively charged, hydrophilic (mica) surface and immersion into citrate buffer (pH 3.0). The TEM scale bar is 50 nm. Patatin (main potato protein unit 21 ) monomers seem to aggregate with each other, ultimately forming a nanofibril-like protofilament structure. d Scattered intensity ( I ) as a function of the scattering vector ( q ) for the potato protein protofilament/xanthan gum hydrogel (2 XGH/PoPF) and each individual component (1.2 wt% XGH, and 0.6 wt% PoPF), at 25 °C, measured by small-angle neutron scattering (SANS). Solid lines correspond to fits to the data obtained using the unified model 29 , 30 . e Nanostructure of the self-assembled protofilament/-hydrogel (2 XGH/PoPF) observed by AFM, following sample deposition onto a negatively charged, hydrophilic (mica) surface and immersion into citrate buffer (pH 3.0). A continuously connected protofilament-containing hydrogel network seems to form. Both XGH and the mica surface being negative charge, no adsorption was expected, therefore preventing any structure from being visible once measuring XGH on its own (Supplementary Fig.  1 ). Each measurement was reproduced at least three times; the average measurement is shown.

figure 2

a Snapshot of the MD simulation of a single potato protein (patatin)-based protofilament (PoPF) in an aqueous solution (radius of gyration ( R g ) = 46.1 ± 0.1 Å, eccentricity = 0.06 ± 0.01). b Snapshot of the MD simulation of xanthan gum (XG) partially coating PoPF in an aqueous solution ( R g  = 44.8 ± 0.1 Å, eccentricity = 0.15 ± 0.01). The XG molecules are represented by the chains of spheres, which are coloured by their element: cyan (carbon), red (oxygen) and white (hydrogen). PoPF is represented by a cartoon that shows the secondary structure of the potato protein (patatin). The cartoon of each patatin in PoPF is coloured differently. c , d Snapshots from the MD simulation of PoPF interacting with XGH in the presence of a PDMS surface, where two PoPF are connected by XGH, whereas the naked (i.e., uncoated by XGH) part of PoPF interacts with PDMS. The PDMS is in the form of a slab, and its molecules are shown as spheres where cyan (carbon), silicon (yellow), red (oxygen) and white (hydrogen) are used to denote the different atoms in the polymer. Contact maps for ( e ) PoPF and PDMS and ( f ) PoPF and XGH.

DLS measurements reveal that PoPF exhibits a monomodal size distribution (Fig.  1a ). Both TEM (Fig.  1b ) and AFM (Fig.  1c and Supplementary Fig.  1 ) images reveal that the 65 °C heat treatment of the 6 wt% potato protein isolate solution fabricated at pH 3.0 results in the formation of long, curly nanofibrillar structures via the aggregation of sphere-shaped patatin (main potato protein unit 21 ) particles, with diameter and height sizes of ca. 20 nm and 3–5 nm, respectively.

Both images have the same scale length (ca. 500 nm), but the sphere-shaped patatin particles are about 10 times larger in the AFM micrograph compared to TEM. The fibrillar length could not be determined given the high level of entanglement. The topographical image (Fig.  1c ) is consistent with observations made with a range of globular, plant-based proteins that are physically hydrolysed at low pH and high temperature 19 , 22 , 23 , 24 , 25 , 26 , 27 , including potato protein for which similar morphologies and dimensions were obtained 19 . Under these conditions, fibrillisation is well known to occur following a ‘polypeptide model’, whereby proteins first hydrolyse into shorter, lower molecular weight peptide fragments, which subsequently undergo intermolecular interactions due to denaturation-induced unfolding, and oligomerise into protofilaments, ultimately assembling into amyloid fibrils 17 , 19 , 22 , 23 , 24 , 25 , 26 , 27 .

Herein, the fibrillar structure formed by PoPF seems to comprise only one protofilament (i.e., no assembled and twisted morphology), therefore not reaching the amyloid-like fibril stage (hence the naming protofilaments). XGH unlike PoPF shows no prominent topographical features (Supplementary Fig.  1 ). This was expected as XGH is negatively charged (ζ = −40 mV 28 and thereby cannot adhere to the negatively charged mica surface, contrary to the positively charged PoPF (ζ = +30 mV). SANS data analysis using the unified model 29 , 30 (Fig.  1d ) reveals that PoPF exhibits a radius of gyration ( R g ) of 75.4 ± 1.5 Å with a fractal dimension of ca 3.3, thus suggesting the formation of structures with a quite rough surface—in agreement with observations made through microscopy (Fig.  1b, c ), while with its power of 1.1, XGH seems to be composed of rod-like objects with a R g of 156.4 ± 4.1 Å forming a quite branched, mass fractal network. Although PoPF seems to form fibril-like structures, the system could not be fitted to a cylinder-type model very likely because of its too high polydispersity and branching. Surprisingly, the combination of PoPF and XGH leads to the formation of convoluted, continuous interconnected filaments that seem to cover the mica surface (Fig.  1e and Supplementary Fig.  1 ). This suggests that PoPF is most likely coated in a ‘patchy way’ by XGH, and catalyses the surface binding of this self-assembly by virtue of its cationic surface charge, whereas XGH glues one protofilament to another, ultimately evolving into a dense, electrospun fibre-like, highly ordered filamentous network. The analysis of data corresponding to their combination suggests objects with a rough surface ( R g  = 74.2 ± 2.5 Å, and fractal dimension of ca. 2.9), which may self-assemble in the shape of large aggregates (Fig.  1d ). Noteworthy, this long, stranded architecture occurring due to XG rigid rods is a signature of 2 XGH/PoPF system; instead, the other polysaccharide (KC) seems to form fractal aggregates when mixed with PoPF (Supplementary Fig.  1 ).

MD simulations both in the bulk and at the interface

To further support the proposed patchy model of PoPF coated by XGH, MD simulations were carried out both in the bulk phase and at the interface, to gain an atomistic description of the interactions taking place between PoPF and the XGH aqueous medium, as well as with a highly hydrophobic, polydimethylsiloxane (PDMS) surface. To the best of our knowledge, this is the first MD simulation study performed on a self-assembled system made up of protofilaments and hydrogels. Using 8 patatin molecules to model a protofilament (Fig.  2a , see ‘Experimental’ section), the snapshots obtained from the simulations clearly show that the XG molecules coat only parts of the PoPF moieties, with the patatin in the PoPF moiety interacting with the XG molecules via its C-terminal end (Fig.  2b ). Remarkably, PoPF is adsorbed onto the surface of the PDMS substrates via its ‘naked’, uncoated area (Fig.  2 c) and XG molecules are able to form bridges between neighbouring PoPF entities (Fig.  2d ). We then identified any pair of residues (which are defined as a monomer in PDMS, an amino acid in the potato protein, and a saccharide in XG) in the different molecules to be in contact if their centre of mass was less than 1.5 nm away from a neighbouring molecule.

The contact map in Fig.  2e shows that PoPF binds to the PDMS surface specifically through residues 72 and 90, and through residues 180 and 200. Both of these regions of the protein contain high densities of hydrophobic residues which may shield themselves from the aqueous environment by interacting with the polymeric interface. The interaction of PoPF with the PDMS substrate causes the PoPF moieties to deform and become both more spherical (eccentricity of 0.13 ± 0.01 at the PDMS surface vs. 0.15 ± 0.01 in solution) and more compact ( R g of 43.6 ± 0.1 Å at the PDMS surface vs. 44.8 ± 0.1 Å in solution). Also, the interactions between the patatins in PoPF and XGH (Fig.  2f ) change, as residues 72–90 in the patatins, which had been interacting with XG in the aqueous phase (primarily through the positively charged residues ARG 70 and LYS 76), preferentially interact with the PDMS polymers (Fig.  2e ). When PoPF is adsorbed to the PDMS substrate, XG seems to bind primarily to the patatins through interactions with the C-terminal end (residues 327–359) of the potato protein, which has a particularly high density of positively charged residues (LYS 327, LYS 328, LYS 343, ARG 344, LYS 347, ARG 352, LYS 353, LYS 354, ARG 356, and LYS 359). When XG is bound to PoPF before it adsorbs to the polymer interface, we also observe interactions between residues 113–137 (there are three LYS residues and several polar residues within this region) and the XG molecules which are not available as a result of PoPF adsorption to the PDMS interface. The patatins are generally found to interact via the same regions, which have a high density of hydrophobic residues interacting with the PDMS surface, whether they are in an isolated protofilament (Supplementary Fig.  2a ), interacting with xanthan gum (Supplementary Fig.  2b ), or interacting with both XGH and PDMS (Supplementary Fig.  2c ). Additionally, XG molecules were shown to interact in a head-to-tail arrangement that facilitates fibre-like formations, such that the one end of a given XG molecule interacts with the opposite end of another molecule (Supplementary Fig.  3 ). In summary, the MD simulations (Fig.  2 ) provide insight into the molecular mechanisms that drive the patchy architecture of XGH on PoPF, which was also observed in the topographical images. This patchy architecture seems to be primarily driven by significant binding to the domains of the proteins that have a high density of cationic residues, while the binding of PoPF to the PDMS surface is as expected driven by hydrophobic interactions between domains of the protein with high densities of hydrophobic residues. It is now worth asking whether such a unique patchy architecture of the self-assembled structure offers synergistic lubricity benefits; this will be discussed later.

Rheological performance

Macromolecular hydrodynamic effects have been considered as a key contributor to lubrication mechanism in biological contact surfaces 4 , 31 , 32 , 33 . In order to understand the role of viscous lubrication in the overall lubricity performance, the resistance to shearing and stretching of the self-assembled protofilament/hydrogel was measured using rotational (Fig.  3a and Supplementary Fig.  5a ) and extensional (Fig.  3b and Supplementary Fig.  5b ) rheometry, respectively.

figure 3

Evolution of ( a ) the shear viscosity ( η shear ) as a function of the shear rate, obtained from stress-controlled rotational rheometry measurements, and ( b ) the apparent extensional viscosity ( η extensional ) and Trouton ratio ( T r ) as a function of the extensional rate, obtained from extensional rheometry measurements, performed on the potato protein protofilament/xanthan gum hydrogel (2 XGH/PoPF) and each individual component (1.2 wt% XGH, and 0.6 wt% PoPF), at 37 °C. The self-assembled protofilament/hydrogel exhibits a high-viscosity, shear-thinning behaviour, very similar to that of XGH on its own; in comparison, PoPF displays much lower shear viscosity values. While PoPF was found to break at an extremely short time (ca. t b  = 0.014 s), thereby preventing the assessment of its stretchability, both the self-assembled protofilament/hydrogel (2 XGH/PoPF) and XGH show a rise in both the apparent extensional viscosity and   T r  upon increasing extensional rates, more notable so for the latter; for both, particularly the self-assembled system, elastic forces predominate over viscous ones (as T r  > 3). Each experiment was reproduced at least three times; the average and a representative measurement are shown for shear and extensional rheology, respectively.

The shear viscosity evolution was monitored over a broad range of shear rates (0.01–1000 s −1 ) (Fig.  3a ). Irrespective of the type of polymer (XG or KC) used to prepare the hydrogel (Fig.  3a and Supplementary Fig.  5a ), both the self-assembled system (2 XGH/PoPF or 2 KCH/PoPF) and the individual components (PoPF, XGH, and KCH) display a shear-induced viscosity decrease over the whole range of shear rates studied, thereby exhibiting a shear-thinning behaviour. Nonetheless, compared to PoPF, whose shear viscosity does not exceed η shear  = 1.3 ± 0.1 Pa.s at 0.01 s −1 , and diminishes of only two orders of magnitude upon reaching 1000 s −1 , stabilising at a near-plateau from ca. 10 s −1 , XGH and 2 XGH/PoPF were found to display a strikingly sharper (four orders of magnitude), and similar, shear rate-dependent decrease (from η shear  = 1006 ± 122 Pa.s for XGH, and 757 ± 23 Pa.s for 2 XGH/PoPF, at 0.01 s −1 , to 0.052 ± 0.003 Pa.s for XGH, and 0.050 ± 0.002 Pa.s for 2 XGH/PoPF, at 1000 s −1 ). Compared to 2 XGH/PoPF (Fig.  3a ), 2 KCH/PoPF displays an order of magnitude lower viscosity, particularly at 0.01 s −1 (Supplementary Fig.  5a ); such an observation might be associated with the structural differences of the continuous, compact filaments offering more viscosity in contrast to the fractal aggregates (Supplementary Fig.  1 ). Of more importance, the viscosity of the self-assembled protofilament/hydrogel overlaps with that of the hydrogel alone, independently of the polymer types or shear rates tested (Fig.  3a and Supplementary Fig.  5a ). This result highlights that the lubricity difference (if any) between the individual components and the self-assembled system might be predominantly linked to the surface behaviour rather than the viscous behaviour, the latter, being largely dominated by the hydrogel component.

Changes in capillary thread shape (Supplementary Fig.  6 ) and diameter (Supplementary Fig.  7 ) upon extensional deformation were recorded over time, and the evolution of the apparent extensional viscosity ( η extensional ) and T r , which are characteristics of the lubricant viscoelasticity) (Fig.  3b and Supplementary Fig.  5b ) were plotted against extensional rates.

While the extensional rheology of PoPF could not be assessed due to its too short breaking time (ca. t b  = 0.014 s), XGH and the 2 XGH/PoPF systems were found to form very long-lived and slender filaments (Supplementary Figs.  6 and 7 ). Contrary to rotational rheology measurements, which show similar shear viscosity evolutions upon increasing shear rates for the XGH-based samples, extensional rheology measurements demonstrate better resistance to thinning ( t b  = 0.67 ± 0.37 s for 2 XGH/PoPF vs. t b  = 0.51 ± 0.22 s for XGH), lower extensional viscosity values (ca. η extensional  = 24.0 ± 9.1 Pa.s for 2 XGH/PoPF vs. ca. η extensional  = 44.2 ± 27.3 Pa.s for XGH), and lower Trouton ratio values (ca. T r  = 21.0 ± 2.8 for 2 XGH/PoPF vs. ca. T r  = 30.6 ± 6.8 for XGH) for the self-assembled system (Fig.  3b ). Similar trends, but lower in magnitude, were observed in terms of both shear and extensional rheology when changing the polysaccharide type from XGH to KCH (Supplementary Fig.  5 ).

Superlubricity

Usually, friction force measurements are carried out at only one scale, even though lubricity and the value of friction coefficients are known to be highly scale dependent 34 . In order to paint a full picture on the lubrication property of these new self-assembled protofilament/hydrogel systems, friction forces were measured at the macroscale using a PDMS ball-on-disk tribometer, with a contact pressure of 300 kPa 35 , which is representative of lower bound pressures found in biological conditions (Fig.  4 and Supplementary Fig.  9 ), and surface force balance (SFB) measurements were conducted at the nanoscale using mica surfaces, with much higher contact pressures (ca. 3 MPa 36 ), which are more representative of articular joint conditions (Fig.  5 and Supplementary Figs.  11 and 12 ) (see ‘Experimental’ section).

figure 4

a Speed-dependent evolution of the friction coefficient, obtained from tribology measurements performed with non-charged, hydrophobic (PDMS) surfaces (at the macroscale), on the self-assembled potato protein protofilament/xanthan gum hydrogel (2 XGH/PoPF) and each individual component (1.2 wt% XGH, and 0.6 wt% PoPF), at 37 °C. The lubrication properties of citrate buffer are also shown for comparison purposes. 2 XGH/PoPF shows an outstanding lubrication performance both in the boundary and hydrodynamic regions, exhibiting ultra-low friction coefficients contrary to both PoPF and XGH on their own. b Influence of the pH on the lubrication performance of the self-assembly system (2 XGH/PoPF) assessed by macroscale tribology. pH was changed from 3.0 to 5.0, upon addition of 1.0 M NaOH, and was increased from pH 3.0 to 7.0, before being decreased back to 3.0 with 1.0 M HCl. The superlubricity of 2 XGH/PoPF was retained at a pH below neutral pH (i.e., below the potato protein isolate isoelectric point (pI of ca. 5)). c Time-dependent evolution of the resonance frequency (Δ f ) measured using QCM-D, upon adsorption of the self-assembled potato protein protofilament/xanthan gum hydrogel (2 XGH/PoPF) and each individual component (1.2 wt% XGH, and 0.6 wt% PoPF), in the presence of a PDMS-coated surface. Each sample was injected into the chamber, which was then rinsed with buffer to assess the sample ability to remain adsorbed to the surface following buffer rinsing. For readability purposes, resonance frequencies are only shown for the 5th overtone. Each sample was diluted 20 times prior to any measurements. Both the self-assembled protofilament/hydrogel system and its individual components readily adsorb onto the PDMS surface, with 2 XGH/PoPF showing a much higher extent of adsorption. Each measurement was reproduced at least three times; the average measurement is shown.

figure 5

Normal force ( F n ), normalised by the radius of curvature ( R  = ca. 1 cm), plotted against the surface separation distance between curved mica surfaces ( D ), interacting across: a potato protein protofilaments (0.6 wt% PoPF), b xanthan gum hydrogel ((0.15 wt% XGH), and c the self-assembled protofilament/hydrogel (2XGH/PoPF); the yellow-shaded region summarizes the data from a . The insets show the force-distance profiles close to the hard-wall separations, on a magnified scale. Empty and filled symbols correspond to first and subsequent approaches. Typical traces of shear force vs. time for the sliding mica surfaces across: d PoPF, e XGH, and f 2XGH/PoPF. The uppermost trace in each graph shows the back-and-forth lateral motion applied to the upper surface, whereas the other traces show the forces transmitted to the lower surface at different loads ( F n ) and friction forces ( F s ). Each set of traces was recorded during the same approach profile. Shear force ( F s ) as a function of the normal force ( F n ) measured between mica surfaces across: g PoPF, h XGH, and i 2XGH/PoPF. Red lines in the main (lin–lin) figure and (lin–log) insets are constant friction coefficients ( μ  = ∂ F s /∂ F n ). Open and solid symbols represent first and second approaches, respectively. Each measurement was reproduced at least two times, and at least two contact positions were done in each pair of mica sheets; a representative measurement is shown.

Unlike plant protein-based microgels 16 , PoPF displays high friction coefficient values over the whole entrainment speed range, reaching up to 1.18 ± 0.05 at 0.004 m s −1 , similarly to citrate buffer at pH 3.0; instead, XGH seems to induce better lubrication, particularly in the hydrodynamic regime, where friction coefficients as low as 0.009 ± 0.002 are obtained at 0.1 m s −1 (Fig.  4a ). Surprisingly, the self-assembly of the protofilaments with the hydrogel synergistically results in an unusual, nearly speed-independent friction curve, with friction coefficient values as low as 0.004 ± 0.005 in the boundary regime (at 0.004 m s −1 ) (Fig.  4a ). When plotted against minimum hydrodynamic film thickness ( h min ) (Supplementary Fig.  8 ), the friction curves of the constituents overlapped with that of the buffer. On the contrary, the self-assembly (2 XGH/PoPF system) showed ultra-low friction independent of the film thickness, showing effective boundary films capable of maintaining low friction forces at low surface separations. Of more interest, pH does not seem to affect the lubrication properties of this self-assembled system, an increase in friction being observed only at a neutral pH of 7.0, where both PoPF and XGH are negatively charged and thus unable to interact with each other (Fig.  4b ). Such an ultra-low macroscopic lubrication performance is also observed when decreasing the polysaccharide/protein ratio from 2 to 0.5 (Supplementary Fig.  9a ) and changing the polysaccharide type from XGH to KCH (Fig.  S9b ). Based upon the complementary interfacial data obtained via QCM-D measurements with PDMS-coated silicon surfaces, the remarkably better lubrication properties exhibited by the self-assembled protofilament/hydrogel may be explained by its 3 times stronger ability to adsorb onto the surface and to remain attached following rinsing, compared to that of the polysaccharide hydrogel and protein protofilaments on their own (Fig.  4c ). Such a high adsorption capacity is rather specific to the use of XGH rather than KCH (Supplementary Fig.  10 ), thus highlighting the importance of the filamentous structure formed in the self-assembled 2 XGH/PoPF system (Fig.  1 e) and of its patchiness visualised through MD simulations (Fig.  2 ). Such an efficient friction-reducing mechanism can be attributed to the synergy between: PoPF almost behaving like a colloidal glue attaching to the surface, and XGH network sporadically interacting with the protofilaments forming a water-encapsulating macromolecular brush, ultimately providing hydration. To further investigate the hydration lubrication provided by this system, and clarify whether such ultra-low friction values were maintained in a high contact pressure situation, SFB measurements were conducted with the 2 XGH/PoPF system.

The normal ( F n (D) ) and shear ( F s ) force profiles, as well as the absolute surface separation between two mica surfaces ( D ) across three different systems: PoPF, XGH, and their self-assembly (2 XGH/PoPF), were measured with SFB. The normal force ( F n ) vs. surface separation ( D ) profiles between the mica surfaces bearing PoPF, XGH, and the self-assembled 2 XGH/PoPF system, across each dispersion in citrate buffer were recorded by SFB (see ‘Experimental’ section) 37 , 38 , and plotted in Fig.  5 as F n ( D )/R in accordance with the Derjaguin approximation 37 , 38 , to normalise for slightly different radii of curvature ( R ) of the mica sheets.

In the case of PoPF alone on both surfaces (Fig.  5 a), a monotonic long-ranged repulsion commencing at ca. D  = 600 nm is observed and may be attributed to steric interactions due to loose multilayer PoPF adsorbed on the surface, as indicated by the AFM micrographs (Fig.  1c ) and the broad DLS size distribution (Fig.  1a ). A sharper rise in F n starts at D  < ca. 300 nm, likely due to the compaction of the initially looser adsorbed PoPF. At the strongest compressions attained in our measurements, the surfaces reach a ‘hard-wall’ separation, D HW  = 300 ± 1  nm for both the first (open symbols) and subsequent (filled symbols) approach. This is equivalent to ca. 150 nm of compacted PoPF on each surface, likely in the form of multiple compacted protofilaments of broad size distribution, as suggested by the broad DLS distribution (Fig.  1a ) and AFM micrographs (Fig.  1c ).

In the case of XGH alone on both surfaces (Fig.  5b ), both the magnitude of the exponentially decaying, long-ranged repulsion, as well as the range of the strong steric repulsion, were much shorter compared to PoPF, with steric forces commencing at separations D  ≤ 50 nm. This suggests a compact and/or very weakly adsorbed layer of XGH on each surface. This is expected as the negatively charged XGH 28 would not adsorb to the negatively charged mica surfaces, but weak adsorption due to short-ranged van der Waals attraction at high buffer salt concentration may be possible. The surface separation ( D HW ) under progressively high compression is then reduced from ca. 25 nm to 0 nm. This indicates that XGH is fully squeezed out from the two mica surfaces under compression, in line with the expected very weak (if any) adsorption between XGH and the similarly (negatively) charged mica surfaces. Such a behaviour resembles the macroscale behaviour observed with high friction coefficient values (Fig.  4a ) supported by the limited adsorption observation with both AFM (Supplementary Fig.  1 ) in the presence of mica surfaces and QCM-D (Fig.  4c ) in the presence of PDMS surfaces.

In the case of the self-assembled 2 XGH/PoPF system (Fig.  5c ), monotonically repulsive interaction commences at ca. 500 nm, similarly to PoPF alone (Fig.  5a , see the shaded region in Fig.  5c ). This likely occurs for similar reasons, i.e., large PoPF adsorbed on the mica surface, though now complexed with XGH. It is of interest that the surface separation due to steric repulsion at the highest applied loads is significantly lower for the 2 XGH/PoPF combination (ca. D HW  = 200 nm) than for PoPF alone (for which ca. D HW  = 300 nm). This indicates that the complexation with XGH may result in lower PoPF adsorption on the negatively charged mica, due to the negative charge on XGH, as well as possibly due to changes in the aggregate structures. Importantly, no squeezing out was seen up to the highest loads applied—which correspond roughly to ca. 30 atm (i.e., 3 MPa) of contact pressure between the mica surfaces, suggesting that XGH complexes strongly with PoPF, leaving nonetheless some non-coated areas on PoPF, which in turn adhere strongly to the mica surfaces, as observed in the MD simulations (Fig.  2c, d ).

Figure  5d–f illustrates typical shear force ( F s ) vs. time ( t ) traces at different loads F n between the mica surfaces across dispersions of PoPF, XGH, and their self-assembly (2 XGH/PoPF) in citrate buffer, in response to the lateral back-and-forth motion applied to the upper surface. We note that typical shear traces across the buffer solution alone show very low frictional dissipation (attributed to hydration lubrication) down to mica-mica contact at high loads (Supplementary Fig.  11 ), demonstrating that the buffer carrier does not influence the frictional results across the macromolecular dispersions. The shear traces (Fig.  5d–f ) reveal clearly the following: frictional forces ( F s ) increase strongly already at quite low loads ( F n ) for PoPF (Fig.  5d ) and XGH (Fig.  5e ), but remain consistently low up to high loads for their self-assembly 2 XGH/PoPF (Fig.  5f ), thus showing excellent lubrication. The corresponding variation of F s vs. F n is summarised in Fig.  5g–i .

The friction coefficient ( µ ) values across both (single-component) PoPF (Fig.  5g ) and XGH (Fig.  5i ) dispersions are on the order of 10 −1 , though the origin of the high friction coefficient ( µ) values is likely different in the two cases. In PoPF case, the protein is positively charged, implying strong adsorption onto the mica surface (in the form of aggregates, as earlier discussed and seen in the AFM micrograph (Fig.  1c )); the poor lubrication very likely results from high-energy dissipation as the PoPF layers slide past each other, suggesting poor hydration of the proteins, which is in line with the macroscale behaviour (Fig.  4a ). We believe that it is less likely that bridging of the PoPF aggregates between the mica surfaces occurs, which would also result in high friction, as seen with other bridging macromolecules 39 , 40 . In the case of the negatively charged XGH (Fig.  5h ), molecules may be weakly adsorbed, as noted earlier, to the negatively charged mica surface via Van der Waals interactions effective at the high salt concentration of the buffer carrier, which strongly screens out the electrostatic interactions. In that case the relatively high friction may be attributed to bridging across the inter-surface gap by the long, linear polysaccharide, as observed in earlier studies of adsorbing polymers 39 . At the highest loads, the weakly-adsorbing XGH, also observed in QCM-D studies (Fig.  4b ), is largely if not entirely squeezed out, leaving at most a thin trapped layer bridging the gap, with high frictional dissipation as the surfaces slide past each other; or if the XGH is entirely squeezed out (see inset to Fig.  5b ), the friction may then be high due to mica-mica contact (Supplementary Fig.  11 ). The sliding velocity-independence of the friction across XGH solution (Supplementary Fig.  12 ) also indicates that the high friction observed with XGH alone is likely to be due to a bridging effect rather than a hydrodynamic one, in line with the earlier observed bridging behaviour 39 . The increased friction force with increasing load forces reflects a higher extent of bridging as XGH is squeezed out.

In strong contrast, for the shear across the 2 XGH/PoPF (Fig.  5i ), the friction, after an initial sharp rise at the lowest loads, is seen to be very low, reaching friction coefficient ( µ ) values in the order of 10 −4 to 10 −5 at the highest loads (at which the corresponding mean contact pressures are of ca. 30 atm). We attribute this as follows. Based on the AFM micrographs of the 2 XGH/PoPF system on mica (Fig.  1e ), the complexation between XGH and PoPF leads to densely packed, thread-like aggregates of pearl-like units (whose size appears roughly half that of the pearl-like units of the aggregates of PoPF alone, Fig.  1c ). This suggests the following scenario: the 2 XGH/PoPF system attaches strongly to the mica surface thanks to the naked (uncoated) areas of PoPF (for which ζ-potential of +30 mV) via a counterion-release mechanism. These partially surface-attached PoPF parts, however, expose the XGH component at their outer surface, and when these highly hydrated XGH moieties slide past each other, hydration lubrication reduces the frictional dissipation to the extremely low values observed (ca. µ  = 10 −4 –10 −5 ). Thus, the role of PoPF (via its positive charge) is to anchor the 2 XGH/PoPF aggregates to the negatively charged substrates, while exposing the outer hydrated XGH component allowing significant friction reduction. The initial sharp rise in friction at the lowest loads (Fig.  5i ), noted above and seen in similar systems previously 41 , is due to viscoelastic dissipation, as the loose PoPF are initially sheared, prior to their compression and compaction at higher loads, on approach and sliding of the surfaces.

In summary, the SFB measurements describing above the interactions between mica surfaces across dispersions of PoPF, XGH, and the 2 XGH/PoPF mixture, together with the AFM micrographs of the corresponding surface structures on mica in these dispersions, shed strong light on the nature of the frictional interactions in these three cases. The high friction conditions are then largely due to molecular interactions (e.g. bond breaking/reforming) as the boundary layers slide past each other, while the low friction arises from much-reduced dissipation due to shear of the sub-nanometre hydration layers in the hydration lubrication mechanism, together with passage over energy barriers as described previously 36 (where phonons are generated leading to the weak dissipation observed). In particular, these studies reveal that the very low friction achieved across the 2 XGH/PoPF system arises from the combined roles of the potato protein protofilaments and the highly hydrated polysaccharide hydrogel: the former acts largely to anchor the surface protofilaments to the (negatively-charged) mica substrate, while the latter provides efficient hydration lubrication by being exposed by the surface-attached protofilaments 42 , 43 , 44 . This underpins and illuminates the molecular mechanisms underlying the findings from the macroscopic tribology phenomena, where a similar synergistic behaviour of PoPF and XGH was apparent.

Summarising all these complementary suite of techniques, we demonstrate that an electrostatic self-assembly of PoPF and XGH (Fig.  6 ) provides efficient hydration lubrication fulfilling the high-pressure, low-friction requirements of biological conditions. This unique self-assembly provides robust boundary lubrication via (1) uncovered PoPF anchoring with surfaces effectively; (2) PoPF glueing to the XGH bringing XGH closer to the surface; whilst (3) the exposed, highly hydrated XGH complexed with PoPF providing the hydration lubrication. A key limitation of the study is that conventional, smooth, highly hydrophobic PDMS surfaces were employed in the macrotribology, QCM-D, and molecular dynamics (MD) simulations measurements, whilst negatively charged, hydrophilic (mica) surfaces were used for the atomic force microscopy (AFM) and surface force balance (SFB) experiments. However, it is worth noting that the superlubricity by these surface-attached protofilament-hydrogel self-assembly can be attributed to the molecular interactions between these boundary layers controlling the frictional dissipation that persisted irrespective of the surfaces used. In addition, the diversity of measurements allowed elucidating the molecular mechanism responsible for such outstanding lubrication properties of this potato protein protofilaments/polysaccharide hydrogel—not only at multiple length scales, but also under varying contact pressure conditions. With these range of measurements and surfaces, we are therefore tackling a wide range of biological applications where pressure many vary from few hundreds to thousands of kilopascals such as oral lubrication to those found in articular joints (MPa), showing the high potential of this aqueous lubricant.

figure 6

The green mesh represents the potato protein-based protofilaments (PoPF) partially coated by orange-coloured filaments connected to each other representing the xanthan based hydrogels (XGH) where the naked part of PoPF (i.e., uncoated by XGH) interacts with the tribo-contact surfaces shown as grey-coloured ball and the rectangular slab. The hydration lubrication mechanism is shown schematically by the transparent water-like spheres attached to the XGH. Zoomed image in the right shows the intact self-assembly outside of the triboshear condition.

Conclusions

In summary, we have shown that potato protein, a by-product from starch industry, can be converted into a high functional aqueous lubricant material offering superlubricity. The synergistic association of potato protein-based protofilaments with polysaccharide hydrogels in a patchy architecture leading to the formation of continuous, convoluted filaments, offer unmatched, ultra-low friction coefficient values (of 10 −2 to 10 −5 ), at moderate-to-extremely high pressures, resembling the remarkable lubrication behaviour of natural biological systems, such as in the oral mucosa, ocular and articular joints 41 , 45 , 46 . The superior sustainability of potato proteins further makes them an ideal candidate for the rational design of a next generation of sustainable, hydrogel-based aqueous lubricants. Future studies are focusing on extending such protofilament-hydrogel self-assembly based aqueous lubrication to other plant protein systems besides potato protein and also investigating the lubrication performance in surfaces with various degrees of hydrophobicity.

Experimental section

Potato protein isolate (91% protein content) was purchased from Sosa Ingredients (Barcelona, Spain), citric acid monohydrate (P > 99.5%) from Alfa Aesar (Thermo Fisher Scientific, Lancashire, UK), Decon 90 from Decon Lab Ltd (Hove, UK), ammonia solution (25 wt%) and toluene from Fisher Scientific (Thermo Fisher Scientific Inc, Loughborough, UK), isopropanol (P99.8%) from MB Fibreglass (Newtownabbey, UK), and xanthan gum (XG), κ -carrageenan (KC), trisodium citrate dihydrate, hydrochloric acid (HCl, 1 M), sodium azide (NaN 3 , P > 99.5%), silicon oil, sulfuric acid (P95.0–98.0%), and hydrogen peroxide solution (30 wt%) from Sigma-Aldrich (Gillingham, UK). The SYLGARD TM 184 silicone elastomer kit employed to coat the silicon sensors with polydimethylsiloxane (PDMS) for the quartz-crystal microbalance with dissipation monitoring (QCM-D) experiments was obtained from Dow Chemical Company Ltd (Cheadle, UK), and the silicon monomer and cross-linking agent were mixed at a 10:1 w/w ratio. Both ultrapure water, or Milli-Q grade water (18.2 MΩ cm, Merck Millipore, Bedford, MA, USA), and deuterium oxide (D 2 O, P99.9%), provided by Sigma-Aldrich (Gillingham, UK), were used in the experiments. Citrate buffer (pH 3.0) was prepared by mixing 10 mM citric acid monohydrate and 10 mM trisodium citrate dihydrate in adequate proportions so as to reach the appropriate acidic pH. NaN 3 (0.02 wt%) was added to all solutions as a preservative. All reagents were used as supplied without any further purification.

Synthesis of the self-assembled protofilament/hydrogel

Unlike protein-based microgels 16 or microgel-reinforced hydrogels 47 , 48 , protofilaments were first created by acidic hydrolysis of the plant protein, and then assembled with a polysaccharide hydrogel network. The plant protein solution (6.0 wt%) was prepared by adding powdered potato protein isolate in 10 mM citrate buffer at pH 3.0 and stirring for ca. 1.5 h to ensure complete solubilisation. Then, the pH of the solution was adjusted to 3.0 by adding 1 M HCl, the solution was centrifuged at 25,000 ×  g for 30 min, and finally the supernatant was heated at 65 °C for 30 min to form potato protein protofilaments (PoPF). Xanthan gum (XG)-based hydrogel (XGH, 1.5 wt%) was prepared by dissolving powdered XG in 10 mM citrate buffer at pH 3.0 at 21 ± 2 °C and shearing the solution for 24 h under constant stirring for complete hydration. On the other hand, κ -carrageenan (KC)-based hydrogel (KCH, 1.5 wt%) was prepared via the dissolution of powdered KC in 10 mM citrate buffer at pH 3.0 by heating at 95 °C while being sheared for 30 min to ensure complete solubilisation. PoPF was added to XGH or KCH dropwise at 21 ± 2 °C, under gentle stirring, to form the self-assembled protofilament/hydrogel at a 2:1 w/w XGH/PoPF or KCH/PoPF ratio, corresponding to a mixture of 1.2 wt% XGH or KCH and 0.6 wt% PoPF.

Particle size and ζ -potential measurements

Hydrodynamic diameter and surface charge ( ζ -potential) measurements were conducted by dynamic light scattering (DLS) on a Zetasizer Ultra instrument (Nano ZS series, Malvern Instruments Ltd, Malvern, UK), at 25 °C, using folded electrophoretic cells (DTS1070, Malvern Instruments Ltd, Malvern, UK) and 100-fold diluted samples.

Transmission electron microscopy (TEM)

The structure of the self-assembled 2XGH/PoPF protofilament/hydrogel was characterised using TEM, with a FEI Tecnai G2 Spirit-T12 microscope (Thermo Fisher Scientific, Waltham, MA, USA), whose electron gun voltage was fixed at 120 kV. Prior to visualisation, the electron contrast was increased by sample negative staining, using the following protocol: (1) electrostatic cleaning of the homemade, 300 mesh, carbon-coated copper grid with a PELCO easiGlow TM glow discharge cleaning system (Ted Pella Inc, Redding, CA, USA); (2) sample deposition onto the grid, followed by excess liquid blotting and ultrapure water washing; (3) grid staining upon addition of 2.0 wt% uranyl acetate; and (4) air-drying. Nanostructure images were captured using a Gantan CCD camera.

Small-angle neutron scattering (SANS)

SANS measurements were performed on the SANS2D time-of-flight instrument, at ISIS pulsed neutron source (STFC Rutherford Appleton Laboratory, Didcot, UK) 49 , using 1 mm path-length quartz cells (Hellma Analytics, Müllheim, Germany) thermostated at 25 °C with a circulating water bath. SANS2D data were acquired with a polychromatic incident beam of wavelength ( λ ) ranging between 1.75 and 16.5 Å, and with a fixed instrument setup of L 1 =  L 2 = 4 m (where L 1 is the collimation length and L 2 the sample-to-detector distance); a simultaneous scattering vector ( q )-range of 0.004 to 1 Å −1 was achieved using two 1 m 2 detectors at 2.4 and 4 m from the sample, with a q -resolution varying from ca. 2% at the highest q -values to ca. 19% with decreasing q -values, calculated using the Mildner Carpenter equation 50 . Detector images were radially averaged and corrected from the scattering of the empty cell and D 2 O background. Detector efficiency corrections and data normalisation to an appropriate standard were done using MANTID 51 .

Atomic force microscopy (AFM)

The topographic images of the self-assembled 2 XGH/PoPF and 2 KCH/PoPF protofilament/hydrogels, as well as of the individual components (XGH, KCH, and PoPF), were obtained using an atomic force microscope ((MFP-3D SA, Oxford Instruments Asylum Research, Santa Barbara, CA, USA). A silicon tip on silicon nitride, V-shaped cantilever (length: 115 μm; nominal spring constant: 0.35 N/m) (SNL, Bruker, Camarillo, CA, USA) was used as the probe, in non-contact mode under liquid. Samples were coated onto freshly cleaved mica surfaces by deposition of the dispersion onto the surface and incubation for ca. 30 min, and were then covered with citrate buffer at pH 3.0.

Molecular dynamics (MD) simulations

The structure of the patatin (main potato protein unit 21 ) used in these simulations was found in the protein database (1oxw) , and the CHARMM-GUI website was employed to build a model of the resulting PoPF in the MARTINI 3.0.0 forcefield 52 . Using this coarse-grain model, a simulation box was created containing 8 patatin molecules, 66,344 water beads, 720 Na + , and 896 Cl − ions (which corresponds to 0.15 M of NaCl solution, the quantity of ions required to neutralise the charge of the system). This system was simulated using the protocol prescribed by the CHARMM-GUI website 53 , which consists of a minimisation simulation using a soft potential, followed by one using the standard Lennard-Jones potential, and then an equilibration step using the NPT ensemble where a target temperature of 338 K and a target pressure of 1 bar were used. Finally, a production simulation was performed using the NPT ensemble with a target temperature of 338 K and a target pressure of 1 bar. The equilibration simulation was run for 1 ns, and the production NPT simulation for 1 μs. In both simulations, a 20 fs timestep was used. The velocity rescale thermostat was employed in both simulations, while the Berendsen barostat and the Parrinello-Rahman barostat were used in the equilibration simulation and the production simulation, respectively.

The self-assembled PoPF resulting from the coarse-grain simulation was then converted to an all-atom model described by the AMBER forcefield using the CHARMM-GUI 54 implementation of backward.py 55 . This all-atom representation of the system contained 692,689 atoms. This system was then simulated using the simulation protocol for proteins in solution prescribed by CHARMM-GUI 56 , which includes a steepest descent energy minimisation simulation, an equilibration simulation using the NVT ensemble in which the Nose-Hoover thermostat is used to control the temperature at 338 K, and then a production simulation using the NPT ensemble with the Nose-Hoover thermostat and the Parrinello Rahman barostat to control the temperature at 338 K and the pressure at 1 bar, respectively. The equilibration simulation was run for 125 ps with a 1 fs timestep, and the production simulation for 100 ns with a 2 fs timestep. The final 50 ns step of this production simulation was used to carry out the analysis of the PoPF in an aqueous solution.

The final configuration from the all-atom simulation of the PoPF in solution was then used as a starting configuration to investigate the interactions of XG molecules with the potato protein nanoparticle. A model of a XG molecule with six repeat units was built using doGlycans 57 and the GLYCAM forcefield 58 , which is based on AMBER. Twenty of these model XG molecules were then randomly inserted into the aqueous environment forming XGH around the PoPF, resulting in a system that contains 784,545 atoms. Then the same all-atom simulation protocol was used for this system as was used for the final all-atom simulations used to study the PoPF. However, for this system, the production simulation was run for 200 ns. Again, the final 50 ns step of the production simulation was used to carry out the analysis of the PoPF interacting with the XG molecules in an aqueous solution.

Finally, we used the final configuration from the production simulation of the PoPF interacting with the XGH and placed it such that it was within 1.5 nm of the interface of a slab of polydimethylsiloxane (PDMS). In order to create the slab of PDMS, we first built a single PDMS molecule with 13 monomers using PySoftK 59 , and then employed PACKMOL to generate a simulation box with 435 polymers. Using the PolyParGen webserver 60 , the AMBER forcefield description of the PDMS polymer was generated. Then, a series of simulations was carried out using the same protocol to generate an amorphous bulk of PDMS as had been previously used for generating an amorphous substrate of another polymer 61 . First, an energy minimisation with a steepest descent algorithm was performed to remove any high-energy steric clashes from the initial configuration. Then the temperature was equilibrated to 800 K in the NVT ensemble using the Berendsen thermostat for 100 ps. Subsequently, the pressure of the simulation box was equilibrated for 5 ns using the Berendsen thermostat and the Berendsen barostat (target pressure of 1 bar). The production simulation was then conducted using the Nosé-Hoover thermostat and the Parrinello-Rahman barostat, to sample from the true NPT ensemble 62 , 63 . The pressure was kept constant at 1 bar. The simulation began at a temperature of 800 K for 50 ns. Each PDMS molecule was able to diffuse and change its conformation at the high temperature. Following this, a first cooling stage was performed at a constant rate of 10 K ns −1 to a temperature of 600 K. Following 50 ns of simulation time, the system was cooled at a constant rate of 10 K ns −1 to the final target temperature of 350 K, at which point we observed no structural evolution of the polymer chains. Thus, only a short simulation at the final temperature was required to allow the local conformational changes of the polymer chains to occur. Finally, a slab of PDMS polymers was created by extending the z-dimension in the simulation box and conducting another energy minimisation in order to create two interfaces of the polymer slab.

After inserting PoPF and XGH near the top interface of the PDMS, water molecules and the necessary ions were added to maintain a neutral system and a 150 mM NaCl solution in the simulation box. The simulation box was large enough in the z-dimension such that it would allow the nanoparticle to be more than 3 nm from the other interface of PDMS through the periodic boundary conditions in the z-dimension. This resulted in a system containing a total of 878,345 atoms. We used a similar simulation protocol as was employed for the PoPF and XGH/PoPF simulations. For this simulation, the production simulation was run for 250 ns. Again, the final 50 ns step of the simulation was used for the analysis. In all simulations, the TIP3P water model was employed to describe the water molecules interactions 64 .

Rotational rheometry

The resistance to shearing of the self-assembled 2 XGH/PoPF and 2 KCH/PoPF protofilament/hydrogel, as well as of the individual components (XGH, KCH, PoPF), was assessed with a stress-controlled rheometer (Kinexus ultra + rotational rheometer, Malvern Instruments, Malvern, UK), fitted with a stainless-steel cone/plate geometry (2° angle cone/60 mm diameter (CP2/60) combined with a 65 mm diameter plate (PL65)) and equipped with a temperature-controlled Peltier system (with a ±0.1 °C temperature stability at thermal equilibrium). Apparent shear viscosity ( \({{\eta }}_{{{{\rm{shear}}}}}\) ) was recorded over a shear rate ranging from 0.1 to 1000 s −1 , at biologically relevant temperature of 37 °C. Each test was repeated at least three times on triplicate samples; the average measurement is shown.

Extensional rheometry

Resistance to stretching was measured using a HAAKE capillary breakup extensional rheometer (CaBER) 1 (Thermo Electron, Karlsruhe, Germany). The thinning of the midpoint diameter of the capillary bridge generated by the rapid separation of two circular plates (6 mm diameter (D o )) that axially constrained the sample was recorded using a laser micrometre, with a beam thickness of 1 mm and a resolution of 20 μm 47 . The initial separation ( h o ) between the two circular plates was set at 3 mm, leading to an initial aspect ratio ( h 0 / D 0 ) of 0.5. The final axial displacement ( h f ) was set at 10 mm in 50 ms to allow filament thinning. Each sample (ca. 0.1 mL) was injected between the plates using a 1 mL syringe. The experiment was triggered 60 s after loading the sample, to limit shear and temperature preconditioning effects. At least five repetitions were performed at 37 °C. High-speed videos of the experiments were also taken at 1000 frames/s, using a PhantomV1612 high-speed camera (Vision Research, Wayne, NJ, USA), to record the shape evolution of the capillary thread. Due to the displacement of the midpoint of the filament, the images acquired were processed using the ImageJ software to detect the filament interface, and compared to the data acquired with the laser micrometre. For an upper convected Maxwell model, the elastocapillary force balance predicts an exponential diameter decay in time with a characteristic relaxation time ( λ c ) 65 :

where D min is the instantaneous filament midpoint diameter. Surface tension measurements were performed in triplicates using the Wilhelmy plate method (Kruss ST10, KRÜSS GmbH, Hamburg, Germany) (Supplementary Fig.  13 ), at 37 °C and minimum speed (0.5 mm min −1 ) to limit the influence of the shear generated between the sample and measuring plate; the average measurement is shown.

The cylindrical elements of the self-assembled 2 XGH/PoPF and 2 KCH/PoPF protofilament/hydrogel, as well as of the individual components (XGH, KCH, PoPF), at the axial mid-plane plate were subject to a strain ( ε ) expressed as:

The instantaneous strain rate ( \(\dot{\in }\) ) for a cylindrical element of fluid is given by:

The apparent extensional viscosity of the liquid ( \({{\eta }}_{{{{\rm{extensional}}}}}\) ) is therefore expressed as:

For coherence with a recent study 47 , 48 ,the transient Trouton ratio ( T r ) was computed as the ratio between the apparent extensional and shear viscosity:

where the dependence of shear rate on the strain rate has been considered at the denominator.

Macroscale tribology

The lubrication performance was evaluated by tribology, using a conventional mini-traction machine (MTM2, PCS Instruments, London, UK) in combination with smooth hydrophobic elastomeric surfaces, i.e., a PDMS ball (19 mm diameter) and disc (46 mm diameter) in a sliding/rolling motion, displaying a 50 nm surface roughness and 2.0 MPa Young’s modulus 47 , 66 . A constant normal force of 2.0 N, corresponding to a Hertzian contact pressure of ca. 300 kPa 35 , and a temperature of 37 °C were applied. The relative motion of the rolling and sliding surfaces is represented by the entrainment speed, which is the average of the ball and disc linear speeds at the contact point, the sliding/rolling ratio being fixed at 50% (i.e., the contribution of both rolling and sliding to motion being defined as equal). The evolution of the friction coefficient was recorded over an entrainment speed range of 0.0035–1.0 m s −1 . The thickness ( h min ) of the hydrodynamic fluid-film for the samples was estimated using the following equation 35 , 67 , 68 , 69 :

where U is the dimensionless speed parameter ( \(\frac{u{\eta}_{\infty }}{E'r'}\) ), W is the dimensionless load parameter ( \(\frac{{F}_{N}}{{E}^{{\prime} }{{r}^{{\prime} }}^{2}}\) ), η ∞ is the viscosity ( η ) of the fluid-film at the tribologically relevant high-shear rates ( \(\dot{\gamma }\) ) at 1000 s −1 and u is the entrainment speed. The η ∞ often is taken as the limiting high-shear viscosity of the fluid in the rheological measurements (the second plateau in the η – \(\dot{\gamma }\) graphs) and is a measure of the hydrodynamic forces generated by the fluid-film during tribo-contacts 67 , 68 . Each experiment was repeated at least three times on triplicate samples; the average measurement is shown.

Surface force balance (SFB) measurements

The SFB setup (Supplementary Fig.  14 ) and the measurement procedures have been described previously in detail 37 , 38 . Briefly, the cleaved molecularly smooth mica sheets (each ca. 2.5 μm thick) were back-silvered and then glued on hemicylindrical quartz lenses of radius of 10 mm, which were oppositely mounted in a crossed-cylinder orientation in the SFB apparatus. The closest separation between two mica surfaces ( D ) was measured from the wavelength of fringes of equal chromatic order (FECO) (accuracy ca. 2–3 Å).

The normal forces ( F n ) were determined from the bending (Δ D 0  − Δ D ) of the horizontal spring, as follows:

where Δ D 0 is the applied normal motion, Δ D the change of surface separation, and K n the horizontal (normal force) spring constant.

The shear forces ( F s ) were determined through measuring the bending Δ X of the vertical spring, as follows:

where K s is the vertical (shear force) spring constant.

We estimate an uncertainty of ±10 −5 in the SFB measurements arising from thermal drift and optical fringe errors 41 .

Quartz-crystal microbalance with dissipation monitoring (QCM-D) measurements

The capacity to adsorb onto a hydrophobic surface and to remain attached following rinsing was evaluated by using a QCM-D (Q-Sense E4 system, Biolin Scientific AB, Västra Frölunda, Sweden), equipped with PDMS-coated sensors 4 .

Silica-coated QCM-D sensors (QSX-303, Q-Sense, Biolin Scientific AB, Västra Frölunda, Sweden) were first treated by UV/Ozone for 15 min to generate hydrophilic surfaces, and then immersed into sulfuric acid for 1 h, before being sonicated twice in ultrapure water for 10 min and dried under nitrogen. The substrates were further cleaned by immersing them into an RCA silicon wafer cleaning solution (made up of 5:1:1 ultrapure water/ammonia/30% hydrogen peroxide) at 80 °C, for 15 min, to remove any remaining organic/insoluble impurities, and by subjecting them to three cycles of 10-min sonication in ultrapure water, before drying them again under nitrogen. Cleaned surfaces were spin-coated with PDMS (prepared in toluene at a concentration 0.5 wt%) at 5000 RPM (with a 2500 RPM/s acceleration), for 60 s, and finally placed under vacuum overnight, at 80 °C, to ensure efficient PDMS cross-linking.

Prior to any measurement, PDMS-coated silicon substrates were thoroughly cleaned through sequential immersion in toluene (30 s), isopropanol (30 s), and ultrapure water (5 min), before rinsing them extensively with ultrapure water, and drying them under nitrogen. Once assembled, the QCM-D flow cells were prefilled with citrate buffer until reaching a stable baseline. Each sample was diluted (i.e., 20-fold dilution in citrate buffer) before being injected into the PDMS-coated silica sensor-containing cell. Once frequency and dissipation reached a plateau, characteristic of adsorption saturation, buffer was flushed into the cell, to study the desorption behaviour of the surface-adsorbed lubricant. Solutions were injected at a flow rate of 100 μL/min, and measurements conducted at a temperature of 25 ± 2 °C. Changes in resonance frequency (Δ f ) were recorded simultaneously over time. Each experiment was reproduced at least three times; a representative curve is shown for the 5th overtone.

Data availability

The raw data in support of most of the quantitative figures reported in this work are available from the corresponding author upon reasonable request.

Zhai, W. & Zhou, K. Nanomaterials in superlubricity. Adv. Funct. Mater. 29 , 1806395 (2019).

Article   Google Scholar  

Zhang, Z. et al. Macroscale superlubricity enabled by graphene-coated surfaces. Adv. Sci. 7 , 1903239 (2020).

Article   CAS   Google Scholar  

Lin, W. & Klein, J. Hydration lubrication in biomedical applications: from cartilage to hydrogels. Acc. Mater. Res. 3 , 213–223 (2022).

Article   CAS   PubMed   PubMed Central   Google Scholar  

Xu, F. et al. A self-assembled binary protein model explains high-performance salivary lubrication from macro to nanoscale. Adv. Mater. Interfaces 7 , 1901549 (2020).

Lin, W. & Klein, J. Recent progress in cartilage lubrication. Adv. Mater. 33 , 2005513 (2021).

Goldberg, R. et al. Boundary lubricants with exceptionally low friction coefficients based on 2d close-packed phosphatidylcholine liposomes. Adv. Mater. 23 , 3517–3521 (2011).

Article   CAS   PubMed   Google Scholar  

Berman, D., Erdemir, A. & Sumant, A. V. Graphene: a new emerging lubricant. Mater. Today 17 , 31–42 (2014).

Chen, M., Briscoe, W. H., Armes, S. P. & Klein, J. Lubrication at physiological pressures by polyzwitterionic brushes. Science 323 , 1698–1701 (2009).

Briscoe, W. H. et al. Boundary lubrication under water. Nature 444 , 191–194 (2006).

Lin, W. et al. Cartilage-inspired, lipid-based boundary-lubricated hydrogels. Science 370 , 335–338 (2020).

Chau, A. L., Edwards, C. E. R., Helgeson, M. E. & Pitenis, A. A. Designing superlubricious hydrogels from spontaneous peroxidation gradients. ACS Appl. Mater. Interfaces 15 , 43075–43086 (2023).

Dunn, A. C., Sawyer, W. G. & Angelini, T. E. Gemini interfaces in aqueous lubrication with hydrogels. Tribology Lett. 54 , 59–66 (2014).

Liu, W., Simič, R., Liu, Y. & Spencer, N. D. Effect of contact geometry on the friction of acrylamide hydrogels with different surface structures. Friction 10 , 360–373 (2022).

Wang, Z., Li, J., Liu, Y. & Luo, J. Macroscale superlubricity achieved between zwitterionic copolymer hydrogel and sapphire in water. Mater. Des. 188 , 108441 (2020).

Poore, J. & Nemecek, T. Reducing food’s environmental impacts through producers and consumers. Science 360 , 987–992 (2018).

Kew, B. et al. Transforming sustainable plant proteins into high performance lubricating microgels. Nat. Commun. 14 , 4743 (2023).

Zhou, J. et al. Oat plant amyloids for sustainable functional materials. Adv. Sci. 9 , 2104445 (2022).

Kamada, A. et al. Controlled self-assembly of plant proteins into high-performance multifunctional nanostructured films. Nat. Commun. 12 , 3529 (2021).

Josefsson, L. et al. Potato protein nanofibrils produced from a starch industry sidestream. ACS Sustain. Chem. Eng. 8 , 1058–1067 (2020).

Holmberg, K. & Erdemir, A. Influence of tribology on global energy consumption, costs and emissions. Friction 5 , 263–284 (2017).

Kew, B., Holmes, M., Stieger, M. & Sarkar, A. Oral tribology, adsorption and rheology of alternative food proteins. Food Hydrocoll. 116 , 106636 (2021).

Akkermans, C. et al. Micrometer-sized fibrillar protein aggregates from soy glycinin and soy protein isolate. J. Agric. Food Chem. 55 , 9877–9882 (2007).

Josefsson, L. et al. Structural basis for the formation of soy protein nanofibrils. RSC Adv. 9 , 6310–6319 (2019).

Munialo, C. D., Martin, A. H., van der Linden, E. & de Jongh, H. H. J. Fibril formation from pea protein and subsequent gel formation. J. Agric. Food Chem. 62 , 2418–2427 (2014).

Li, T. et al. Assembly behavior, structural characterization and rheological properties of legume proteins based amyloid fibrils. Food Hydrocoll. 111 , 106396 (2021).

Li, J. et al. Self-assembly of plant protein fibrils interacting with superparamagnetic iron oxide nanoparticles. Sci. Rep. 9 , 8939 (2019).

Article   PubMed   PubMed Central   Google Scholar  

Herneke, A. et al. Protein nanofibrils for sustainable food–characterization and comparison of fibrils from a broad range of plant protein isolates. ACS Food Sci. Technol. 1 , 854–864 (2021).

Brunchi, C.-E., Bercea, M., Morariu, S. & Dascalu, M. Some properties of xanthan gum in aqueous solutions: effect of temperature and pH. J. Polym. Res. 23 , 123 (2016).

Beaucage, G. Small-angle scattering from polymeric mass fractals of arbitrary mass-fractal dimension. J. Appl. Crystallogr. 29 , 134–146 (1996).

Beaucage, G. Approximations leading to a unified exponential/power-law approach to small-angle scattering. J. Appl. Crystallogr. 28 , 717–728 (1995).

McCutchen, C. W. Mechanism of animal joints: sponge-hydrostatic and weeping bearings. Nature 184 , 1284–1285 (1959).

Klein, J. Molecular mechanisms of synovial joint lubrication. Proc. Inst. Mech. Eng. Part J J. Eng. Tribol. 220 , 691–710 (2006).

Crouzier, T. et al. Modulating mucin hydration and lubrication by deglycosylation and polyethylene glycol binding. Adv. Mater. Interfaces 2 , 1500308 (2015).

Liamas, E., Connell, S. D., Ramakrishna, S. N. & Sarkar, A. Probing the frictional properties of soft materials at the nanoscale. Nanoscale 12 , 2292–2308 (2020).

Sarkar, A., Andablo-Reyes, E., Bryant, M., Dowson, D. & Neville, A. Lubrication of soft oral surfaces. Curr. Opin. Colloid Interface Sci. 39 , 61–75 (2019).

Ma, L., Gaisinskaya-Kipnis, A., Kampf, N. & Klein, J. Origins of hydration lubrication. Nat. Commun. 6 , 6060 (2015).

Dong, Y. et al. Dehydration does not affect lipid-based hydration lubrication. Nanoscale 14 , 18241–18252 (2022).

Klein, J. & Kumacheva, E. Simple liquids confined to molecularly thin layers. I. Confinement-induced liquid-to-solid phase transitions. J. Chem. Phys. 108 , 6996–7009 (1998).

Raviv, U., Tadmor, R. & Klein, J. Shear and frictional interactions between adsorbed polymer layers in a good solvent. J. Phys. Chem. B 105 , 8125–8134 (2001).

Sapir, L. & Harries, D. Macromolecular compaction by mixed solutions: bridging versus depletion attraction. Curr. Opin. Colloid Interface Sci. 22 , 80–87 (2016).

Seror, J., Zhu, L., Goldberg, R., Day, A. J. & Klein, J. Supramolecular synergy in the boundary lubrication of synovial joints. Nat. Commun. 6 , 6497 (2015).

Jahn, S. & Klein, J. Hydration lubrication: the macromolecular domain. Macromolecules 48 , 5059–5075 (2015).

Raviv, U. et al. Lubrication by charged polymers. Nature 425 , 163–165 (2003).

Raviv, U. & Klein, J. Fluidity of bound hydration layers. Science 297 , 1540–1543 (2002).

Hodge, W. A. et al. Contact pressures in the human hip joint measured in vivo. Proc. Natl Acad. Sci. 83 , 2879–2883 (1986).

Hodge, W. A. et al. Contact pressures from an instrumented hip endoprosthesis. JBJS 71 , 1378–1386 (1989).

Pabois, O. et al. Benchmarking of a microgel-reinforced hydrogel-based aqueous lubricant against commercial saliva substitutes. Sci. Rep. 13 , 19833 (2023).

Hu, J., Andablo-Reyes, E., Soltanahmadi, S. & Sarkar, A. Synergistic microgel-reinforced hydrogels as high-performance lubricants. ACS Macro Lett. 9 , 1726–1731 (2020).

Heenan, R. K. et al. Small angle neutron scattering using Sans2d. Neutron N. 22 , 19–21 (2011).

Mildner, D. F. R. & Carpenter, J. M. Optimization of the experimental resolution for small-angle scattering. J. Appl. Crystallogr. 17 , 249–256 (1984).

Arnold, O. et al. Mantid—data analysis and visualization package for neutron scattering and μ SR experiments. Nucl. Instrum. Methods Phys. Res. Sect. A Accel. Spectrom. Detect. Assoc. Equip. 764 , 156–166 (2014).

Souza, P. C. T. et al. Martini 3: a general purpose force field for coarse-grained molecular dynamics. Nat. Methods 18 , 382–388 (2021).

Qi, Y. et al. CHARMM-GUI Martini maker for coarse-grained simulations with the martini force field. J. Chem. Theory Comput. 11 , 4486–4494 (2015).

Lee, J. et al. CHARMM-GUI supports the Amber force fields. J. Chem. Phys. 153 , 035103 (2020).

Wassenaar, T. A., Pluhackova, K., Böckmann, R. A., Marrink, S. J. & Tieleman, D. P. Going backward: a flexible geometric approach to reverse transformation from coarse grained to atomistic models. J. Chem. Theory Comput. 10 , 676–690 (2014).

Lee, J. et al. CHARMM-GUI input generator for NAMD, GROMACS, AMBER, OpenMM, and CHARMM/OpenMM simulations using the CHARMM36 additive force field. J. Chem. Theory Comput. 12 , 405–413 (2016).

Danne, R. et al. doGlycans–Tools for preparing carbohydrate structures for atomistic simulations of glycoproteins, glycolipids, and carbohydrate polymers for GROMACS. J. Chem. Inf. Model. 57 , 2401–2406 (2017).

Kirschner, K. N. et al. GLYCAM06: a generalizable biomolecular force field. Carbohydrates. J. Comput. Chem. 29 , 622–655 (2008).

Santana-Bonilla, A., López-Ríos de Castro, R., Sun, P., Ziolek, R. M. & Lorenz, C. D. Modular software for generating and modeling diverse polymer databases. J. Chem. Inf. Model. 63 , 3761–3771 (2023).

Yabe, M., Mori, K., Ueda, K. & Takeda, M. Development of PolyParGen software to facilitate the determination of molecular dynamics simulation parameters for polymers. J. Comput. Chem. Jpn. 5 , 1–5 (2019).

Ziolek, R. M. et al. Conformational heterogeneity and interchain percolation revealed in an amorphous conjugated polymer. ACS Nano 16 , 14432–14442 (2022).

Hernández-Rodríguez, M., Rosales-Hernández, C. M., Mendieta-Wejebe, E. J., Martínez-Archundia, M. & Basurto, C. J. Current tools and methods in molecular dynamics (md) simulations for drug design. Curr. Med. Chem. 23 , 3909–3924 (2016).

Article   PubMed   Google Scholar  

Nosé, S. A molecular dynamics method for simulations in the canonical ensemble. Mol. Phys. 52 , 255–268 (1984).

Mark, P. & Nilsson, L. Structure and dynamics of the TIP3P, SPC, and SPC/E water models at 298 K. J. Phys. Chem. A 105 , 9954–9960 (2001).

McKinley, G. H. Visco-elasto-capillary thinning and break-up of complex fluids. Annu. Rheol. Rev. 3 , 1–49 (2005).

Liamas, E., Connell, S. D., Zembyla, M., Ettelaie, R. & Sarkar, A. Friction between soft contacts at nanoscale on uncoated and protein-coated surfaces. Nanoscale 13 , 2350–2367 (2021).

Sarkar, A., Soltanahmadi, S., Chen, J. & Stokes, J. R. Oral tribology: providing insight into oral processing of food colloids. Food Hydrocoll. 117 , 106635 (2021).

de Vicente, J., Stokes, J. R. & Spikes, H. A. The frictional properties of Newtonian fluids in rolling–sliding soft-EHL contact. Tribol. Lett. 20 , 273–286 (2005).

Myant, C., Fowell, M., Spikes, H. A. & Stokes, J. R. An investigation of lubricant film thickness in sliding compliant contacts. Tribol. Trans. 53 , 684–694 (2010).

Download references

Acknowledgements

The authors acknowledge the European Research Council (ERC) for the provision of funding under the European Union’s Horizon 2020 research and innovation programme (grant agreements: no. 890644) and UKRI Horizon Europe Guarantee Fund (grant agreement: no. EP/X03514X/1). The authors also thank the UKRI Healthy Ageing Catalyst Award (ES/X006565/1) for supporting this research. O.P. acknowledges: the British Society of Rheology (BSR) for the award of the Undergraduate Summer Research Bursary, and the Royal Society for the award of the International Exchanges Scheme that allowed her to conduct a research visit at Weizmann Institute. O.P. thanks ISIS for the provision of beam time on SANS2D ( https://doi.org/10.5286/ISIS.E.RB2000279 ). This work benefitted from the use of the SasView application, originally developed under NSF award DMR-0520547. SasView contains code developed with funding from the European Union’s Horizon 2020 research and innovation programme under the SINE2020 project, grant agreement No. 654000.

Author information

These authors contributed equally: Olivia Pabois, Yihui Dong.

Authors and Affiliations

School of Food Science and Nutrition, University of Leeds, Leeds, LS2 9JT, UK

Olivia Pabois, Mingduo Mu, Yasmin Message, Evangelos Liamas, Arwen I. I. Tyler & Anwesha Sarkar

Department of Molecular Chemistry and Materials Science, Weizmann Institute of Science, 76100, Rehovot, Israel

Yihui Dong, Nir Kampf & Jacob Klein

Department of Engineering, King’s College London, London, WC2R 2LS, UK

Christian D. Lorenz

ISIS Neutron and Muon Source, Science and Technology Facilities Council, Rutherford Appleton Laboratory, Didcot, OX11 ODE, UK

James Doutch

Université Paris-Saclay, INRAE, AgroParisTech, UMR SayFood, 91120, Palaiseau, France

Alejandro Avila-Sierra & Marco Ramaioli

Unilever Research & Development Port Sunlight, Quarry Road East, Bebington, Merseyside, CH63 3JW, UK

Evangelos Liamas

You can also search for this author in PubMed   Google Scholar

Contributions

O.P., J.K., C.L., M.R., and A.S. conceptualised the study and designed the experimental protocol. A.S. supervised the study. O.P. designed the study protocol. O.P., M.M., Y.M., and E.L. performed the rotational rheology, macroscale tribology, and QCM-D experiments. O.P. conducted the TEM and AFM measurements with the guidance of N.K. and J.K. O.P. thanks Dr Nadav Elad for his technical support with the TEM experiments. O.P. performed the SANS measurements with the help of J.D. and A.T. C.L. designed and carried out the MD simulation study. Y.D., N.K., and J.K. designed the SFB experiments and performed the data analyses. A.A.S. and M.R. developed and conducted the extensional rheology experiments. O.P., Y.D., A.A.S., C.L., and A.S. prepared the figures, and wrote the main manuscript text. All authors reviewed the manuscript.

Corresponding author

Correspondence to Anwesha Sarkar .

Ethics declarations

Competing interests.

The authors declare no competing interests.

Peer review

Peer review information.

Communications Materials thanks Yiping Cao and the other, anonymous, reviewer(s) for their contribution to the peer review of this work. Primary Handling Editor: Jet-Sing Lee.

Additional information

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

Supplementary information

Peer review file, supplementary information, rights and permissions.

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

About this article

Cite this article.

Pabois, O., Dong, Y., Kampf, N. et al. Self-assembly of sustainable plant protein protofilaments into a hydrogel for ultra-low friction across length scales. Commun Mater 5 , 158 (2024). https://doi.org/10.1038/s43246-024-00590-5

Download citation

Received : 29 April 2024

Accepted : 25 July 2024

Published : 03 September 2024

DOI : https://doi.org/10.1038/s43246-024-00590-5

Share this article

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

Quick links

  • Explore articles by subject
  • Guide to authors
  • Editorial policies

Sign up for the Nature Briefing newsletter — what matters in science, free to your inbox daily.

journal of biomedical materials research

COMMENTS

  1. Journal of Biomedical Materials Research Part A

    Journal of Biomedical Materials Research Part A is an interdisciplinary biomaterials journal that features contributions related to the preparation, performance, and evaluation of biomaterials in the fields of medicine, biology, and engineering.. At the interface of materials science and biomedicine, the journal publishes research on all relevant biomaterial topics, including the science and ...

  2. Journal of Biomedical Materials Research Part B: Applied Biomaterials

    JOURNAL METRICS >. Journal of Biomedical Materials Research Part B: Applied Biomaterials is a highly interdisciplinary journal that focuses on the applied aspects of biomaterials, including their design, development, interactions, and behavior in medical devices and other clinical applications. We are serving the needs of biomaterials ...

  3. Journal of Biomedical Materials Research Part A

    Opportunities for biomaterials to address the challenges of COVID-19 [REVIEW ARTICLE] J. Biomed. Mater. Res. 2020, vol. 108, no. 10, p. 1974. Journal of Biomedical Materials Research Part A is a biomaterials journal at the interface of materials science & biomedicine, covering medicine, biology & engineering.

  4. Journal of Biomedical Materials Research

    A peer-reviewed journal of original contributions on biomaterials science and technology, published by John Wiley & Sons. Find its scope, impact factor, quartile, and similar journals on ScimagoJR, a platform for scientific indicators.

  5. Journal of Biomedical Materials Research

    Learn about the history, scope, and impact of this peer-reviewed scientific journal of biomedical material science. It has two parts, Part A and Part B, published by John Wiley & Sons since 2003.

  6. JBMR-A&B

    JBMR-A&B is the official journal of the Society for Biomaterials, covering biomaterials research and applications. It is the #1 journal in materials science and biomaterials, and requires subscription access.

  7. Journal of Biomedical Materials Research

    A peer-reviewed journal serving biomaterials professionals who design, develop, produce and apply biomaterials and medical devices. It covers all disciplines where medical devices are used and publishes original research, reviews, concepts, reports and editorials.

  8. Journal of Biomedical Materials Research

    The Journal of Biomedical Materials Research is an international, interdisciplinary, English-language publication of original contributions concerning studies of the preparation, performance, and ...

  9. Journal of Biomedical Materials Research Part A

    A peer-reviewed journal of original contributions on biomaterials science and technology, published by Wiley. The journal covers alloys, polymers, ceramics, tissues, and medical devices, and is the official journal of several biomaterials societies.

  10. Journal of Biomedical Materials Research

    Scimago Journal Rank (SJR) is a prestige indicator that ranks journals by their average citations per article. See the SJR, quartile, documents, and citations per document of Journal of Biomedical Materials Research, a biomaterials and biomedical engineering journal published in the US.

  11. Journal of Biomedical Materials Research Part A

    Journal of Biomedical Materials Research Part A<br/>The Journal of Biomedical Materials Research is an international, interdisciplinary, English-language publication of original contributions concerning studies of the preparation, performance, and evaluation of biomaterials; the chemical, physical, toxicological, and mechanical behavior of materials in physiological environments; and the ...

  12. Journal of Biomedical Materials Research Part A

    Issue Edited by: Chris A. Bashur, Ngan F. Huang. Volume 112, Pages: 307-483. March 2024. Volume 112, Pages: 139-306. February 2024. Journal of Biomedical Materials Research Part A is a biomaterials journal at the interface of materials science & biomedicine, covering medicine, biology & engineering.

  13. Journal of Biomedical Materials Research

    Journal of Biomedical Materials Research. [...] Membrane molecular weight (MW) cut-off is a critical factor for immunoprotection of transplanted microencapsulated cells as well as for graft ...

  14. Society for Biomaterials (SFB)

    Journal of Biomedical Materials Research Part B: Applied Biomaterials, Volume 112, Issue 9, September 2024. Login or join to access the journal. MS&T24. The Materials Science & Technology (MS&T) technical meeting and exhibition is the long-standing, recognized forum for fostering technical innovation at the intersection of materials science ...

  15. Journal of Biomedical Materials Research Part B: Applied Biomaterials

    Aims and Scope. Journal of Biomedical Materials Research - Part B: Applied Biomaterials is a highly interdisciplinary peer-reviewed journal serving the needs of biomaterials professionals who design, develop, produce and apply biomaterials and medical devices. It has the common focus of biomaterials applied to the human body and covers all ...

  16. Collagen-based biomaterials for biomedical applications

    Subsequently, some of the superior properties of collagen-based biomaterials, such as biocompatibility, biodegradability, mechanical properties, and cell activities, are pinpointed. These properties make collagen applicable in biomedicine, such as wound healing, tissue engineering, surface coating of medical devices, and skin supplementation.

  17. Journal of Biomedical Materials Research Part A

    Journal of Biomedical Materials Research Part A is a biomaterials journal at the interface of materials science & biomedicine, covering medicine, biology & engineering.

  18. Human Amniotic Membrane: A review on tissue engineering ...

    Abstract. Human amniotic membrane (hAM) has been employed as scaffolding material in a wide range of tissue engineering applications, especially as a skin dressing and as a graft for corneal treatment, due to the structure of the extracellular matrix and excellent biological properties that enhance both wound healing and tissue regeneration.

  19. Journal of Biomedical Materials Research Part A: List of Issues

    Journal of Biomedical Materials Research Part A is a biomaterials journal at the interface of materials science & biomedicine, covering medicine, biology & engineering.

  20. Biomedical Materials

    Biomedical Materials. Biomedical Materials publishes original research findings and critical reviews that contribute to our knowledge about the composition, properties, and performance of materials for all applications relevant to human healthcare. Mohd Aiman Hakimi Abdul Rahim et al 2024 Biomed. Mater. 19 045040. Chenlu Li et al 2024 Biomed.

  21. Journal of Biomaterials Applications: Sage Journals

    The Journal of Biomaterials Applications publishes original articles that emphasize the development, manufacture and clinical applications of biomaterials. Biomaterials continue to be one of the most rapidly growing areas of research today and certainly one of the biggest technical challenges, and in particular the interaction of the surrounding tissues and cells on the biomaterials is ...

  22. Journal of the Mechanical Behavior of Biomedical Materials

    About the journal. The Journal of the Mechanical Behavior of Biomedical Materials is concerned with the mechanical deformation, damage and failure under applied forces, of biological material (at the tissue, cellular and molecular levels) and of biomaterials, i.e. those materials which are designed to mimic or …. View full aims & scope.

  23. Protein Immobilization on Bacterial Cellulose for Biomedical Application

    New carriers for protein immobilization are objects of interest in various fields of biomedicine. Immobilization is a technique used to stabilize and provide physical support for biological micro- and macromolecules and whole cells. Special efforts have been made to develop new materials for protein immobilization that are non-toxic to both the body and the environment, inexpensive, readily ...

  24. 3D‐printed silicate porous bioceramics ...

    Journal of Biomedical Materials Research Part B: Applied Biomaterials. Volume 112, Issue 9 e35469. RESEARCH ARTICLE. ... Among multiple materials applied in the repair of bone defect, silicate bioceramics attracted researchers because of its excellent bioactivity. The purpose of this study was to detect the effect of specific bioactive glass ...

  25. Self-assembly of sustainable plant protein protofilaments into a

    Superlubricity is important for energy and biomedical applications but typical building blocks are limited to synthetically-sourced polymeric materials. Here, self-assembly of plant-based ...