Submitted Successfully!
To reward your contribution, here is a gift for you: A free trial for our video production service.
Thank you for your contribution! You can also upload a video entry or images related to this topic.
Version Summary Created by Modification Content Size Created at Operation
1 -- 2650 2022-07-19 03:29:34 |
2 Reference format revised. -15 word(s) 2635 2022-07-20 03:35:24 |

Video Upload Options

Do you have a full video?

Confirm

Are you sure to Delete?
Cite
If you have any further questions, please contact Encyclopedia Editorial Office.
Ho, T.;  Chang, C.;  Chan, H.;  Chung, T.;  Shu, C.;  Chuang, K.;  Duh, T.;  Yang, M.;  Tyan, Y. Applications of Hydrogels in Biomedicine. Encyclopedia. Available online: https://encyclopedia.pub/entry/25254 (accessed on 23 April 2024).
Ho T,  Chang C,  Chan H,  Chung T,  Shu C,  Chuang K, et al. Applications of Hydrogels in Biomedicine. Encyclopedia. Available at: https://encyclopedia.pub/entry/25254. Accessed April 23, 2024.
Ho, Tzu-Chuan, Chin-Chuan Chang, Hung-Pin Chan, Tze-Wen Chung, Chih-Wen Shu, Kuo-Pin Chuang, Tsai-Hui Duh, Ming-Hui Yang, Yu-Chang Tyan. "Applications of Hydrogels in Biomedicine" Encyclopedia, https://encyclopedia.pub/entry/25254 (accessed April 23, 2024).
Ho, T.,  Chang, C.,  Chan, H.,  Chung, T.,  Shu, C.,  Chuang, K.,  Duh, T.,  Yang, M., & Tyan, Y. (2022, July 19). Applications of Hydrogels in Biomedicine. In Encyclopedia. https://encyclopedia.pub/entry/25254
Ho, Tzu-Chuan, et al. "Applications of Hydrogels in Biomedicine." Encyclopedia. Web. 19 July, 2022.
Applications of Hydrogels in Biomedicine
Edit

Hydrogels are crosslinked polymer chains with three-dimensional (3D) network structures, which can absorb relatively large amounts of fluid. Because of the high water content, soft structure, and porosity of hydrogels, they closely resemble living tissues. Research in recent years shows that hydrogels have been applied in various fields, such as agriculture, biomaterials, the food industry, drug delivery, tissue engineering, and regenerative medicine. Along with the underlying technology improvements of hydrogel development, hydrogels can be expected to be applied in more fields. 

hydrogel medical application 3D cell culture drug delivery wound dressing tissue engineering

1. Introduction

Hydrogels comprise a three-dimensional (3D) network which can absorb a large amount of water and swell in the water due to their hydrophilic groups, such as -NH2, -COOH, -OH, -CONH2, -CONH, and -SO3H [1][2][3][4][5][6][7][8][9]. Its network is usually constructed by crosslinked polymer chains that sometimes can be formed through crosslinked colloidal clusters [10][11][12][13][14][15][16][17]. They can be flexible and soft, which are results of their water absorption ability [18]. Chemical or physical crosslinking of natural or synthetic polymer chains can be used to design the hydrogels [19][20][21][22][23]. Because of the high water content, soft structure, and porosity of hydrogels, they closely resemble living tissue. Wichterle and Lim first developed hydrogels for biomaterials in 1960. They produced a synthetic poly-2-hydroxyethyl methacrylate (PHEMA) hydrogel, which was then used as a filler for eye enucleation and contact lenses [24]. Since then, the expense of hydrogels in drug delivery and bioactive compound release has been elevated in several early studies from the 1970s to the 1990s [25][26][27][28][29]. In the 1990s, hydrogels were applied in tissue engineering [30][31][32][33]. The application of hydrogels was restricted to only the surface environment from the 1970s to the 1990s, for applications in the eye or open wounds, for example. The properties (e.g., swelling–deswelling rate, stiffness, degradability, mech size) of hydrogels can be adjusted by changing the hydrophilic and hydrophobic ratios, the initiator or polymer concentrations, and the reaction conditions (time, temperature, container, etc.) [34][35][36][37]. The biomedical application of hydrogels is not limited to the surface environment due to in situ gelation after infection and the stimuli responsiveness of the hydrogel [38][39].
Over the past 60 years, hydrogels have been engineered to be implantable, injectable, and sprayable for many organs and tissues [38][39]. Recently, hydrogels have gained attention in the field of environmental engineering [40], soft robotics [41], and wastewater treatment [42]. With the underlying technological improvement of hydrogel generation, hydrogels can be expected to be used in more fields.

2. Biomedical Applications of Hydrogels

2.1. 3D Cell Cultures

Three-dimensional cell cultures provide a useful platform for the cell to grow in vitro in all directions. Compared with the 2D culture system, it is easier to understand the in vivo cell behavior, since cells form a 3D structure in living tissue. The 3D cell culture is achieved by culturing the cells on a 3D scaffold. In the in vivo 3D cell structures, the cells are embedded in the extracellular matrix (ECM) and form a 3D structure. ECM is known to play an important role in regulating the cell behavior [43]. Hydrogels have a 3D structure and a hydrophilic polymer network capable of absorbing water in addition to biological fluid [1][2][3][4][5][6][7][8][9][44][45]. Thus, they can construct the soft and wet 3D structure which is like the extracellular matrix (ECM), which is available to encapsulate the cells. This results in those hydrogels which have gained increasing attention in the application of scaffolds for 3D cell cultures [46][47].
Hydrogels can comprise natural, synthetic, and semi-synthetic polymers. These hydrogels provide distinct biochemical, physical, and mechanical properties for the 3D cell culture [43]. Table 1 describes the recent application of these hydrogels for 3D cell culture. Natural hydrogels have good biocompatibility, endogenous factors, and the similar viscoelasticity and fibrils of the ECM. These hydrogels can support cell activity for 3D cell cultures.
Table 1. Natural, synthetic, and semi-synthetic hydrogels for 3D cell cultures.
Source of Hydrogels Properties Materials Cell Applications
Natural Provide comparable viscoelasticity and fibrils to the ECM; having good biocompatibility; endogenous factors can support cellular activity Collagen Rat chondrocyte [48], hMSCs [49][50], rMSC [51], HUVECs/hASCs [52] Maintain the chondrocyte phenotype [48]; facilitate chondrogenic differentiation of hBMSCs [49] and rBMSCs [51]; form stable EC networks [52]; promote cell viability; promote growth of hMSCs [50].
HA hiPSC-NPCs [53], hiPSCs [54], rMSCs [55], human breast cancer MCF-7 cells [56], HepG2 cells [57], human dental pulp stem cells [58], hNS/PC [59], and hMSCs [60] Promote the neural differentiation of hiPSC-NPCs [53]; cardiac differentiation of hiPSCs [54]; osteogenic differentiation of human dental pulp stem cells [58]; the adhesion and proliferation of HepG2 cells [57]; cell spreading, fiber remodeling, and focal adhesion of hMSCs [60]; maintain the stemness of rMSCs and induce the direct cartilage differentiation [61]; enhance the tumorigenic capability of MCF-7 cells [56]; increase the oligodendrocytes and neural differentiation of hNS/PC and support long-term cell viability [59].
Fibrin HUVECs/hMSCs [62], porcine cumulus–oocyte complexes (COCs) [63], primary human chondrocytes [64], mHPSCs [65], and hiPSCs/HUVECs/human dermal fibroblast [66] Prevascular formation of HUVECs, improve cell viability and proliferation of hMSCs and enhance their osteogenic differentiation and bone mineral deposition [62]; maintain the functional relationship between oocytes and follicular cells [63]; induce the production of glycosaminoglycans and collagen type II of primary human chondrocytes [64]; enhance the murine hematopoietic stem/progenitor cells (mHPSCs) expansion and differentiation [65]; no effect viability and prevascular formation of encapsulated cells [66].
Alginate hESCs/hiPSCs [67], hiPSCs-derived neurons [68] Enhance the generation of retinal pigmented epithelium and neural retina of hESCs/hiPSCs [67]; form complex neural networks [68].
Synthetic Have the good mechanical strength to provide structural support for various cell types in 3D cell culture PVA mHSCs [69], mSCCs [70], human glioma cell lines LN299, U87MG and Gli36 [71], human breast cancer Hs578T cells, and human pancreatic cancer cell lines Sui67 and Sui72 [72] Enhance the expansion of murine hematopoietic stem cells (mHSCs) [69]; promote the meiotic and post-meiotic differentiation rate of mSCCs [70]; form tumor spheroids [71][72].
PEG hiPSCs [73], mMSCs [74], chondrocyte [75], and hMSCs [76] Enhance the hematopoietic differentiation of hiPSCs [73]; evaluate the behavior of mMSCs [74] and hMSCs at the specific condition [76]; prolong the oxygen release of chondrocytes [75].
Semi-synthetic Have a feature of ECM microenvironment and faster stress relaxation HA–PEG hiPS-HEPs [77] and HUVECs [78] Enhance viability and functionality of hiPS-HEPs [77]; promote the capillary-like sprouts formation of HUVECs spheroids [78].
RGD–alginate–PEG Fibroblasts and mMSCs [79] Increase the spread and proliferation of fibroblasts and the osteogenic differentiation of mMSCs [79].

2.2. Drug Delivery

Polymers are one of the most promising substances for the preparation of drug delivery systems. Polymers can be prepared for various nanostructures, including polymeric micelles, polymeric vesicles, and hydrogels. Those nanostructures are great for drug delivery [80]. Increased interest in hydrogels is focused on smart hydrogels due to the stimuli-responsive properties of polymeric moieties. Stimuli-responsive properties can enable the formulation of novel targeted drugs and control drug release through non-intravenous administration. It can also delay the effect of opsonization by low blood contact.
The basic advantage is the ability of a smart hydrogel to change its properties (such as mechanical properties, swelling capacity, hydrophilicity, or permeability of bioactive molecules) under the effect of surroundings, including temperature, pH, electromagnetic radiation, magnetic field, and biological factors. Smart hydrogels can be prepared by the natural or synthetic polymers. There is a problem with natural hydrogels, which is that its mechanical properties make it difficult to maintain consistency.
Although this problem with natural hydrogel can be overcome by extensive chemical modification for natural polymers, it is very difficult to process [81]. In contrast, synthetic polymers are easy to alter their chemical or physical properties. The biodegradable and hydrophilic synthetic polymers are the most competitive substances for the synthesis of smart hydrogels for drug delivery. Those synthetic polymers endow smart hydrogels with low toxicity, low side effects, and low blood material adhesion. Of these, the low blood material adhesion can slow the effect of opsonization and reduce the phagocyte elimination. Table 2 shows the application of those synthetic smart hydrogels for drug delivery.
Table 2. Smart hydrogels for drug delivery.
Hydrogels Drug Materials Sustained-Release Time Proposed Applications Ref.
Thermoresponsive hydrogel Dexamethasone HPMA More than 30 days Osteoarthritis and rheumatoid arthritis [81]
Topotecan Poloxamer 407 and poloxamer 188 28 days Colorectal cancer [82]
Lamivudine and zidovudine Pluronic® F-127 168 h AIDS [83]
Antibody PEGMA 13 days Enhance the efficacy of antibody treatment [84]
pH-responsive hydrogel Bortezomib mPEG-LUT 50 h Colorectal cancer [85]
Amifostine (S-2(3-aminopropylamino) ethyl phosphorothioate MAC-g-PCL 6 h Acute radiation syndrome [86]
Photoresponsive hydrogel Doxycycline SPCOOH modified-silicone-hydrogel (poly(HEMA-co-PEGMEA)) 42 h Inflammation disease [87]
Insulin BP, pNIPAM, PEG, and ETPTA Not detected Diabetic disease [88]
Daul-responsive hydrogel          
pH/thermo Doxorubicin chemosensitizer curcumin poly (NIPAAm-co-DMAEMA) 168 h Colon cancer [89]
  Methotrexate   50 h Breast cancer [90]
pH/redox Magnesium ions poly (NIPAAm-co-DMAEMA)
PLP-CDE
6 h Ionic therapeutics [91]

2.3. Wound Dressings

The skin is the largest human organ and consists of epidermis, endothelium, and subcutaneous tissue from outside to inside. Skin is attacked by physical, chemical, or thermal damage, which results in wounding. Wounds lead to the destruction of skin structure and function [92]. The creation of a wound will trigger a series of physiological responses that promote wound repair, known as wound healing [93]. Wounds can be categorized by the nature of the repair as acute and chronic wounds. Acute wounds are mainly caused by mechanical injuries, such as abrasions, cuts, burns, scalds, or surgical incisions, and can heal completely in about 8–12 weeks [94]. Chronic wounds are wounds with delayed healing, 12 weeks after the initial injury [89]. These wounds are mainly caused by repeated tissue damage, underlying physiological factors (such as diabetes, impaired angiogenesis, innervation, or cell migration), or acquired physiological factors (such as malignancy or infection) [95][96]. Once a chronic wound forms, it can eventually lead to amputations or even mortality [97]. The wound healing process is dynamic and complicated. It involves four phases: hemostasis, inflammation, proliferation, and remodeling [98]. The hemostasis phase occurs within minutes of the injury. During this period, platelets stick to the wound site, engage with collagen, and release thrombin, which activates fibrin to form a network that stops blood loss. The inflammation phase occurs when immune cells (especially neutrophils and macrophages) are recruited to the wound site by platelets. Immune cells engulf damaged cells, dead cells, bacteria, and other pathogens at the wound site. At the same time, various peptide growth factors are released by platelets and inflammatory cells, which promote the migration of fibroblasts to the wound site.
During the proliferation phase, fibroblasts proliferate at the wound site and rebuild the dermal tissue, employing granulation tissue formation and extracellular matrix protein deposition. Within the granulation tissue, blood vessel networks will be formed, providing sufficient oxygen and nutrients to improve cell survival. Epithelial cells then migrate from the wound edge to the center to cover the defect: this process is termed re-epithelialization. During the remodeling phase, excess collagen fibers are degraded in the dermis, and the wound shrinks and heals rapidly. Therefore, the use of wound dressings to quickly stop bleeding, prevent infection, and promote repair can speed up wound healing and reduce unnecessary mortality.
Characteristics of an ideal wound dressing should (1) provide and maintain a moist environment, (2) permit the easy transmission of gases, (3) remove exudates and absorb blood from the wound, (4) have low adherence to skin, (5) reduce wound necrosis, (6) prevent infection, (7) allow heat insulation, (8) enhance epidermal migration, (9) promote angiogenesis, (10) have low toxicity and be biocompatible and biodegradable [98]. Several studies have shown that hydrogels can form a physical barrier and remove excess exudate. They also provide a moist environment to promote the process of wound healing. In addition, hydrogels can be applied as a sprayable or injectable wound dressing, which may fill irregularly shaped wounds [99][100][101]. They also present with similar properties as the natural extracellular matrix (ECM), biocompatibility, biodegradability, and tunable properties (such as shape, gel state, and mechanical strength). These advantages of hydrogels can simulate the development of hydrogels for different dressings for different types of wounds. Recently, functional hydrogels have received a lot of attention in wound dressing research. These hydrogels can exhibit high-performance biological activities, such as antibacterial properties, promoting blood coagulation, or promoting blood regeneration, etc. [102].
Wound dressings with hemostasis, angiogenesis, antibacterial infection, and anti-inflammation characteristic have a good impact on wound healing. Natural polymers, such as cellulose, chitosan, collagen, and HA, contain endogenous bioactivation factors. These natural hydrogels are a good wound dressing for wound healing. For example, an in-situ-formed collagen–HA hydrogel was adapted to promote spontaneous wound healing. In addition, this hydrogel inhibited the growth of planktonic Escherichia coli (E. coli) and Staphylococcus aureus (S. aureus) [103]. PEG-modified collagen–chitosan hydrogels further reduce the zone diameters of E. coli and S. aureus biofilms. This hydrogel also exhibits hemostatic ability, which can enhance wound healing [104]. Zhu L. and Chen L. developed CF-encapsulated graphene–silk fibroin macromolecular hydrogel dressings, which have functions of antibacterial (both planktonic and biofilm S. aureus and Pseudomonas aeruginosa (P. aeruginosa)) and enhanced fibroblasts growth. These have a great healing ability for burn wounds [105]. Khaliq T et al. used the chitosan HCl, κ-carrageenan, and PVA-based, physically crosslinked hydrogel to load cefotaxime sodium (CTX), which displays a high oxygen permeability and antibacterial capacity for inhibiting the biofilm size of S. aureus, P. aeruginosa, and E. coli. This hydrogel provided higher re-epithelialization and good granulation tissue formation for healing burn wounds in a diabetic rat model [106]. In addition, the silver-nanoparticle-loaded pH hydrogel also showed the effective elimination of P. aeruginosa and Staphylococcus epidermidis (S. epidermidis) in in vitro antibacterial biofilm studies. This hydrogel provides a promising strategy to enhance the healing of drug-resistant-bacteria-infected wounds [107]. The in vivo effect of this hydrogel needs further investigation. Collagen–PEG injectable hydrogels containing umbilical cord stem cell factor (SCF) can induce neovascularization and skew toward M2 macrophages in diabetic wounds. They can promote diabetic wound repair based on their angiogenesis and anti-inflammation abilities [108]. It is unknown whether this hydrogel has functions in the inhibition of bacterial growth due to the lack of antibacterial activity assay studies. A 3-carboxy-phenylboronic-acid-grafted gelatin–PVA hydrogel exhibits excellent hemostasis properties enhancing cell adhesion. This hydrogel further encapsulates the vancomycin-conjugated silver nanoclusters (VAN-AgNCs) and nimesulide (NIM), endowing an anti-inflammatory effect. It also has the capacity to inhibit the planktonic S. aureus and P. aeruginosa growth in a VAN-AgNCs dose-dependent manner. In an in vivo experiment, this VAN-AgNCs- and NIM-loaded 3-carboxy-phenylboronic-acid-grafted gelatin–PVA hydrogel can induce the sequential healing processes to promote the healing of chronically infected diabetic wounds [109]. Plasma-exosomes-loaded, pH-responsive carboxymethylcellulose (P-Exos-loaded CMC) hydrogel stimulates the activation of the vascular endothelial growth factor (VEGF) signaling pathway. This pathway further enhances angiogenesis and re-epithelialization to promote the wound healing process in diabetic type 1 mice [110]. Another study uses the umbilical-cord-derived mesenchymal stem cell exosomes combined with Pluronic F127 hydrogel to demonstrate that this hydrogel can induce the expression of transforming growth factor beta-1 (TGFβ-1) and cell proliferation in addition to VEGF production. Based on the above ability, it can enhance the regeneration of granular tissue and angiogenesis in chronic diabetic wound healing [111]. However, the antibacterial activity of exosomes-loaded hydrogels is unclear in these two studies [110][111].

2.4. Tissue Engineering

Tissue engineering is a promising and challenging strategy to treat patients who suffer functional failure and irreparable tissue destruction [112]. The aim of tissue engineering is to develop a scaffold mimicking an in vivo extracellular matrix to support tissue regeneration. Hydrogels have gained great interest in tissue engineering due to their mechanical strength, biocompatibility, biodegradability, and the resemblance to in vivo extracellular matrix [113].
A hydrogel scaffold can be useful in tissue regeneration of nerves, cardiac tissue, cartilage, and bone. For example, the 3D printing of collagen–chitosan is beneficial in decreasing scar and cavity formation and can improve the regeneration of nerve fibers, as well as functional recovery, when tested in an animal model [114]. Another example is HA combined with alginate and fibrin. This was applied as an ink ingredient of 3D printing in peripheral nerve tissue regeneration [115]. In addition, the HA–cellulose hydrogels can repair the central nerves [116]. Li J et al. used horseradish peroxidase (HRP) and choline oxidase (ChOx) crosslinked gelation hydrogel to encapsulate the mMSCs. This hydrogel displays a high capacity to promote cellular viability, neural differentiation, and neurotrophic secretion of loaded mMSCs. Based on that capacity, it can enhance the survival and proliferation of endogenous neural cells and neurological function recovery of traumatic-brain-injured mice [117].

References

  1. Velmurugan, N.; Kumar, G.G.; Han, S.S.; Nahm, K.S.; Lee, Y.S. Synthesis and characterization of potential fungicidal silver nano-sized particles and chitosan membrane containing silver particles. Iran. Polym. J. 2009, 18, 383–392.
  2. Jiao, T.F.; Zhou, J.A.; Zhou, J.X.; Gao, L.H.; Xing, Y.Y.; Li, X.H. Synthesis and characterization of chitosan-based schiff base compounds with aromatic substituent groups. Iran. Polym. J. 2011, 20, 123–136.
  3. Gbenebor, O.P.; Adeosun, S.O.; Lawal, G.I.; Jun, S.; Olaleye, S.A. Acetylation, crystalline and morphological properties of structural polysaccharide from shrimp exoskeleton. Eng. Sci. Technol. 2017, 20, 1155–1165.
  4. Peng, Z.; Li, Z.; Shen, Y. Influence of Chemical Cross-Linking on Properties of Gelatin/Chitosan Micropheres. Polym. Plast Technol. 2012, 51, 381–385.
  5. Zamani, A.; Taherzaden, M.J. Effects of partial dehydration and freezing temperature on the morphology and water binding capacity of carboxymehty chitosan-based superabsorbents. Ind. Eng. Chem. Res. 2010, 49, 8094–8099.
  6. Adair, A.; Kaesaman, A.; Klinpituksa, P. Superabsorbent materials derived fromhydroxyethyl cellulose and bentonite: Preparation, Characterization and swelling capacities. Polym. Test. 2017, 64, 321–329.
  7. Olad, A.; Pourkhiyabi, M.; Gharekhani, H.; Doustdar, F. Semi-IPN superabsorbent nanocomposite based on sodium alginate and montmorillonite: Reaction parameters and swelling characteristics. Carbohydr. Polym. 2018, 190, 95–306.
  8. Rop, K.; Mbui, D.; Njomo, N.; Karuku, G.N.; Michira, I.; Ajayi, R.F. Biodegradable water hyacinth cellulose-graft-poly(ammunium acrylate-co-acrylic acid)polymer hydrogel for potential agricultural application. Heliyon 2019, 5, e01416.
  9. Zain, G.; Nada, A.A.; El-Sheikh, M.A.; Attaby, F.A.; Waly, A.I. Superabsorbent hydrogel based on sulfonated-starch for improving water and saline absorbency. Int. J. Biol. Macromol. 2018, 115, 61–68.
  10. Peppas, N.A.; Merrill, E.W. Development of semicrystalline poly(vinyl alcohol) hydrogels for biomedical applications. J. Biomed. Mater. Res. 1977, 11, 423–434.
  11. Dror, M.; Elsabee, M.Z.; Berry, G.C. Interpenetrating polymer networks for biological applications. Biomater. Med. Devices Artif. Organs 1979, 7, 31–39.
  12. Peppas, N.A.; Moynihan, H.J.; Lucht, L.M. The structure of highly crosslinked poly(2-hydroxyethyl methacrylate) hydrogels. J. Biomed. Mater. Res. 1985, 19, 397–411.
  13. Gander, B.; Gurny, R.; Doelker, E.; Peppas, N.A. Effect of polymeric network structure on drug release from cross-linked poly(vinyl alcohol) micromatrices. Pharm. Res. 1989, 6, 578–584.
  14. Douglas, A.M.; Fragkopoulos, A.A.; Gaines, M.K.; Lyon, L.A.; Fernandez-Nieves, A.; Barker, T.H. Dynamic assembly of ultrasoft colloidal networks enables cell invasion within restrictive fibrillar polymers. Proc. Natl. Acad. Sci. USA 2017, 114, 885–890.
  15. Nair, S.K.; Basu, S.; Sen, B.; Lin, M.H.; Kumar, A.N.; Yuan, Y.; Cullen, P.J.; Sarkar, D. Colloidal gels with tunable mechanomorphology regulate endothelial morphogenesis. Sci. Rep. 2019, 9, 1072.
  16. Lu, P.J.; Zaccarelli, E.; Ciulla, F.; Schofield, A.B.; Sciortino, F.; Weitz, D.A. Gelation of particles with short-range attration. Nature 2008, 453, 449–503.
  17. Tsurusawa, H.; Leocmach, M.; Russo, J.; Tanaka, H. Direct link between mechanical stability in gels and percolation of isostatic particles. Sci. Adv. 2019, 5, 1–8.
  18. Yuqi, Z.; Yao, L.; Jiaqi, L.; Pu, G.; Liping, H. Super water absorbecy OMMT/PAA hydrogel matrials with excellent mechanical properties. RSC Adv. 2017, 7, 14504.
  19. Hahn, S.K.; Park, J.K.; Tomimatsu, T.; Shimoboji, T. Synthesis and degradation test of hyaluronic acid hydrogels. Int. J. Biol. Macromol. 2007, 40, 374–380.
  20. Ranjha, N.; Mudassir, J.; Akhtar, N. Methyl methacrylate-co-itaconic acid (MMA-co-IA) hydrogels for controlled drug delivery. J. Sol-Gel Sci. Technol 2008, 47, 23–30.
  21. Eagland, D.; Crowther, N.J.; Butler, C.J. Complexation between polyoxyethyl and polymethacrylic acid- the importance of the molar-mass of polyoxyethylene. Eur. Polym. J. 1994, 30, 767–773.
  22. Yokoyama, F.; Masada, I.; Shimamura, K.; Ikawa, T.; Monobe, K. Morphology and structure of highy elastic polyvinyl-aclohol hydrogel prepared by repeated freezing-and-melting. Colloid Polym. Sci. 1986, 264, 595–601.
  23. Cerchiara, T.; Luppi, B.; Bigucci, F.; Orienti, I.; Zecchi, V. Physically cross-linked chitosan hydrogels as topical vehicles for hydrophilic drugs. J. Pharm. Pharmacol. 2002, 54, 1453–1459.
  24. Wichterle, O.; Lím, D. Hydrophilic Gels for Biological Use. Nature 1960, 185, 117–118.
  25. Haldon, R.; Lee, B. Structure and permeability of porous films of poly (hydroxy ethyl methacrylate). Br. Polym. J. 1972, 4, 491–501.
  26. Korsmeyer, R.W.; Peppas, N.A. Effect of the morphology of hydrophilic polymeric matrices on the diffusion and release of water soluble drugs. J. Membr. Sci. 1981, 9, 211–227.
  27. Song, S.; Cardinalx, J.; Kim, S.; Kim, S. Progestin permeation through polymer membranes V: Progesterone release from monolithic hydrogel devices. J. Pharm. Sci. 1981, 70, 216–219.
  28. Yean, L.; Bunel, C.; Vairon, J.P. Reversible immobilization of drugs on a hydrogel matrix, 2. Diffusion of free chloramphenicol from poly (2-hydroxyethyl methacrylate) hydrogels. Makromol. Chem. 1990, 191, 1119–1129.
  29. Roorda, W.; De Vries, M.; De Leede, L.; De Boer, A.; Breimer, D.; Junginger, H. Zero-order release of oxprenolol-HCl, a new approach. J. Control. Release 1988, 7, 45–52.
  30. Rowley, J.A.; Madlambayan, G.; Mooney, D.J. Alginate hydrogels as synthetic extracellular matrix materials. Biomaterials 1999, 20, 45–53.
  31. Soon-Shiong, P.; Heintz, R.E.; Merideth, N.; Yao, Q.X.; Yao, Z.; Zheng, T.; Murphy, M.; Moloney, M.K.; Schmehl, M.; Harris, M. Insulin independence in a type 1 diabetic patient after encapsulated islet transplantation. Lancet 1994, 343, 950.
  32. Bellamkonda, R.; Ranieri, J.P.; Bouche, N.; Aebischer, P. Hydrogel-based three-dimensional matrix for neural cells. J. Biomed. Mater. Res. 1995, 29, 663–671.
  33. Tabata, Y.; Yamada, K.; Miyamoto, S.; Nagata, I.; Kikuchi, H.; Aoyama, I.; Tamura, M.; Ikada, Y. Bone regeneration by basic fibroblast growth factor complexed with biodegradble hydrogels. Biomaterials 1998, 19, 807–815.
  34. Miao, L.; Hu, J.; Lu, M.; Tu, Y.; Chen, X.; Li, Y.; Lin, S.; Li, F.; Hu, S. Alkynyl-functionalization of hydroxypropyl cellulose and thermoresponsive hydrogel thereof prepared with P(NIPAAm-co-HEMAPCL). Carbohydr. Polym. 2016, 137, 433–440.
  35. Lin, C.Y.; Battistoni, C.M.; Liu, J.C. Redox-Responsive Hydrogels with Decoupled Initial Stiffness and Degradation. Biomacromolecules 2021, 22, 5270–5280.
  36. Liu, T.; Zhang, Y.; Sun, M.; Jin, M.; Xia, W.; Yang, H.; Wang, T. Effect of Freezing Process on the Microstructure of Gelatin Methacryloyl Hydrogels. Front. Bioeng. Biotechnol. 2021, 9, 810155.
  37. Antić, K.; Onjia, A.; Vasiljević-Radović, D.; Veličković, Z.; Tomić, S.L. Removal of Nickel Ions from Aqueous Solutions by 2-Hydroxyethyl Acrylate/Itaconic Acid Hydrogels Optimized with Response Surface Methodology. Gels 2021, 7, 225.
  38. Kopecek, J. Hydrogels: From soft contact lenses and implants to self-assembled nanomaterials. J. Polym. Sci. Part A Polym. Chem. 2009, 47, 5929–5946.
  39. Kouchak, M. In situ gelling systems for drug delivery. Jundishapur J. Nat. Pharm. Prod. 2014, 9, e20126.
  40. Yu, A.C.; Lopez Hernandez, H.; Kim, A.H.; Stapleton, L.M.; Brand, R.J.; Mellor, E.T.; Bauer, C.P.; McCurdy, G.D.; Wolff, A.J., 3rd. Wildfire prevention through prophylactic treatment of high-risk landscapes using viscoelastic retardant fluids. Proc. Natl. Acad. Sci. USA 2019, 116, 20820–20827.
  41. Migliorini, L.; Santaniello, T.; Yan, Y.; Lenardi, C.; Milani, P. Low-voltage electrically driven homeostatic hydrogel-based actuators for underwater soft robotics. Sens. Actuators B Chem. 2016, 228, 758–766.
  42. Mittal, H.; Ray, S.S.; Okamoto, M. Recent Progress on the Design and Applications of Polysaccharide-Based Graft Copolymer Hydrogels as Adsorbents for Wastewater Purification. Macromol. Mater. Eng. 2016, 301, 496–522.
  43. Park, Y.; Huh, K.M.; Kang, S.W. Applications of Biomaterials in 3D Cell Culture and Contributions of 3D Cell Culture to Drug Development and Basic Biomedical Research. Int. J. Mol. Sci. 2021, 22, 2491.
  44. Qing, X.; He, Z.; Liu, Y.; Yin, Y.; Cai, W.; Fan, L.; Fardim, P. Preparation and propeties of polyvinyl alcohol/N–succinyl chitosan/lincomycin composite antibacterial hydrogels for wound dressing. Carbohydr. Polym. 2021, 261, 117875.
  45. Nurzynska, A.; Klimek, K.; Palka, K.; Szajnecki, Ł.; Ginalska, G. Curdlan-Based Hydrogels for Potential Application as Dressings for Promotion of Skin Wound Healing-Preliminary In Vitro Studies. Materials 2021, 14, 2344.
  46. Jose, G.; Shalumon, K.T.; Chen, J.P. Natural Polymers Based Hydrogels for Cell Culture Applications. Curr. Med. Chem. 2020, 27, 2734–2776.
  47. Habanjar, O.; Diab-Assaf, M.; Caldefie-Chezet, F.; Delort, L. 3D Cell Culture Systems: Tumor Application, Advantages, and Disadvantages. Int. J. Mol. Sci. 2021, 22, 12200.
  48. Jin, G.Z.; Kim, H.W. Effects of Type I Collagen Concentration in Hydrogel on the Growth and Phenotypic Expression of Rat Chondrocytes. Tissue Eng. Regen. Med. 2017, 14, 383–391.
  49. Tamaddon, M.; Burrows, M.; Ferreira, S.A.; Dazzi, F.; Apperley, J.F.; Bradshaw, A.; Brand, D.D.; Czernuszka, J.; Gentleman, E. Monomeric, porous type II collagen scaffolds promote chondrogenic differentiation of human bone marrow mesenchymal stem cells in vitro. Sci. Rep. 2017, 7, 43519.
  50. Buitrago, J.O.; Patel, K.D.; El-Fiqi, A.; Lee, J.H.; Kundu, B.; Lee, H.H.; Kim, H.W. Silk fibroin/collagen protein hybrid cell-encapsulating hydrogels with tunable gelation and improved physical and biological properties. Acta Biomater. 2018, 69, 218–233.
  51. Kilmer, C.E.; Battistoni, C.M.; Cox, A.; Breur, G.J.; Panitch, A.; Liu, J.C. Collagen Type I and II Blend Hydrogel with Autologous Mesenchymal Stem Cells as a Scaffold for Articular Cartilage Defect Repair. ACS Biomater. Sci. Eng. 2020, 6, 3464–3476.
  52. Andrée, B.; Ichanti, H.; Kalies, S.; Heisterkamp, A.; Strauß, S.; Vogt, P.M.; Haverich, A.; Hilfiker, A. Formation of three-dimensional tubular endothelial cell networks under defined serum-free cell culture conditions in human collagen hydrogels. Sci. Rep. 2019, 9, 5437.
  53. Wu, S.; Xu, R.; Duan, B.; Jiang, P. Three-Dimensional Hyaluronic Acid Hydrogel-Based Models for In Vitro Human iPSC-Derived NPC Culture and Differentiation. J. Mater. Chem. B 2017, 5, 3870–3878.
  54. Xu, J.; Shamul, J.G.; Staten, N.A.; White, A.M.; Jiang, B.; He, X. Bioinspired 3D Culture in Nanoliter Hyaluronic Acid-Rich Core-Shell Hydrogel Microcapsules Isolates Highly Pluripotent Human iPSCs. Small 2021, 17, e2102219.
  55. Ren, Y.; Zhang, H.; Wang, Y.; Du, B.; Yang, J.; Liu, L.; Zhang, Q. Hyaluronic Acid Hydrogel with Adjustable Stiffness for Mesenchymal Stem Cell 3D Culture via Related Molecular Mechanisms to Maintain Stemness and Induce Cartilage Differentiation. ACS Appl. Bio Mater. 2021, 4, 2601–2613.
  56. Suo, A.; Xu, W.; Wang, Y.; Sun, T.; Ji, L.; Qian, J. Dual-degradable and injectable hyaluronic acid hydrogel mimicking extracellular matrix for 3D culture of breast cancer MCF-7 cells. Carbohydr. Polym. 2019, 211, 336–348.
  57. Bucatariu, S.M.; Constantin, M.; Varganici, C.D.; Rusu, D.; Nicolescu, A.; Prisacaru, I.; Carnuta, M.; Anghelache, M.; Calin, M. A new sponge-type hydrogel based on hyaluronic acid and poly(methylvinylether-alt-maleic acid) as a 3D platform for tumor cell growth. Int. J. Biol Macromol. 2020, 165, 2528–2540.
  58. La Gatta, A.; Tirino, V.; Cammarota, M.; La Noce, M.; Stellavato, A.; Pirozzi, A.; Portaccio, M.; Diano, N.; Laino, L. Gelatin-biofermentative unsulfated glycosaminoglycans semi-interpenetrating hydrogels via microbial-transglutaminase crosslinking enhance osteogenic potential of dental pulp stem cells. Regen. Biomater. 2021, 8, rbaa052.
  59. Seidlits, S.K.; Liang, J.; Bierman, R.D.; Sohrabi, A.; Karam, J.; Holley, S.M.; Cepeda, C.; Walthers, C.M. Peptide-modified, hyaluronic acid-based hydrogels as a 3D culture platform for neural stem/progenitor cell engineering. J. Biomed. Mater. Res. A 2019, 107, 704–718.
  60. Lou, J.; Stowers, R.; Nam, S.; Xia, Y.; Chaudhuri, O. Stress relaxing hyaluronic acid-collagen hydrogels promote cell spreading, fiber remodeling, and focal adhesion formation in 3D cell culture. Biomaterials 2018, 154, 213–222.
  61. Geuss, L.R.; Allen, A.C.; Ramamoorthy, D.; Suggs, L.J. Maintenance of HL-1 cardiomyocyte functional activity in PEGylated fibrin gels. Biotechnol. Bioeng. 2015, 112, 1446–1456.
  62. Heo, D.N.; Hospodiuk, M.; Ozbolat, I.T. Synergistic interplay between human MSCs and HUVECs in 3D spheroids laden in collagen/fibrin hydrogels for bone tissue engineering. Acta Biomater. 2019, 95, 348–356.
  63. Gorczyca, G.; Wartalski, K.; Tabarowski, Z.; Duda, M. Proteolytically Degraded Alginate Hydrogels and Hydrophobic Microbioreactors for Porcine Oocyte Encapsulation. J. Vis. Exp. 2020, 161, 61325.
  64. Bachmann, B.; Spitz, S.; Schädl, B.; Teuschl, A.H.; Redl, H.; Nürnberger, S.; Ertl, P. Stiffness Matters: Fine-Tuned Hydrogel Elasticity Alters Chondrogenic Redifferentiation. Front. Bioeng. Biotechnol. 2020, 8, 373.
  65. Garcia-Abrego, C.; Zaunz, S.; Toprakhisar, B.; Subramani, R.; Deschaume, O.; Jooken, S.; Bajaj, M.; Ramon, H.; Verfaillie, C. Towards Mimicking the Fetal Liver Niche: The Influence of Elasticity and Oxygen Tension on Hematopoietic Stem/Progenitor Cells Cultured in 3D Fibrin Hydrogels. Int. J. Mol. Sci. 2020, 21, 6367.
  66. Jarrell, D.K.; Vanderslice, E.J.; Lennon, M.L.; Lyons, A.C.; VeDepo, M.C.; Jacot, J.G. Increasing salinity of fibrinogen solvent generates stable fibrin hydrogels for cell delivery or tissue engineering. PLoS ONE 2021, 16, e0239242.
  67. Hunt, N.C.; Hallam, D.; Karimi, A.; Mellough, C.B.; Chen, J.; Steel, D.; Lako, M. 3D culture of human pluripotent stem cells in RGD-alginate hydrogel improves retinal tissue development. Acta Biomater. 2017, 49, 329–343.
  68. Moxon, S.R.; Corbett, N.J.; Fisher, K.; Potjewyd, G.; Domingos, M.; Hooper, N.M. Blended alginate/collagen hydrogels promote neurogenesis and neuronal maturation. Mater. Sci. Eng. C Mater. Biol. Appl. 2019, 104, 109904.
  69. Wilkinson, A.C.; Ishida, R.; Kikuchi, M.; Sudo, K.; Morita, M.; Crisostomo, R.V.; Yamamoto, R.; Loh, K.M.; Nakamura, Y. Long-term ex vivo haematopoietic-stem-cell expansion allows nonconditioned transplantation. Nature 2019, 571, 117–121.
  70. Ziloochi Kashani, M.; Bagher, Z.; Asgari, H.R.; Najafi, M.; Koruji, M.; Mehraein, F. Differentiation of neonate mouse spermatgonial stem cells on three-dimensional agar/polyvinyl alcohol nanofiber scaffold. Syst. Biol. Reprod. Med. 2020, 66, 202–215.
  71. Molyneaux, K.; Wnek, M.D.; Craig, S.; Vincent, J.; Rucker, I.; Wnek, G.E.; Brady-Kalnay, S.M. Physically-cross-linked poly(vinyl alcohol) cell culture plate coatings facilitate preservation of cell-cell interactions, spheroid formation, and stemness. J. Biomed. Mater. Res. B Appl. Biomater. 2021, 109, 1744–1753.
  72. Okita, Y.; Zheng, L.; Kawanishi, K.; Miyoshi, H.; Yanagihara, K.; Kato, M. Polyvinyl alcohol scaffolds and supplementation support 3D and sphere culturing of human cancer cell lines by reducing apoptosis and promoting cellular proliferation. Genes Cells 2021, 26, 336–343.
  73. Sidhu, I.; Barwe, S.P.; Kiick, K.L.; Kolb, E.A.; Gopalakrishnapillai, A. A 3-D hydrogel based system for hematopoietic differentiation and its use in modeling down syndrome associated transient myeloproliferative disorder. Biomater. Sci. 2021, 9, 6266–6281.
  74. Sylvester, C.B.; Pugazenthi, A.; Grande-Allen, K.J.; Ghanta, R.K. Cell-Laden Bioactive Poly(ethylene glycol) Hydrogels for Studying Mesenchymal Stem Cell Behavior in Myocardial Infarct-Stiffness Microenvironments. Cardiovasc. Eng. Technol. 2021, 12, 183–199.
  75. Chen, S.C.; Yang, M.H.; Chung, T.W.; Jhuang, T.S.; Yang, J.D.; Chen, K.C.; Chen, W.J.; Huang, Y.F.; Jong, S.B. Preparation and Characterization of Hyaluronic Acid-Polycaprolactone Copolymer Micelles for the Drug Delivery of Radioactive Iodine-131 Labeled Lipiodol. Biomed. Res. Int. 2017, 2017, 4051763.
  76. Jansen, L.E.; Kim, H.; Hall, C.L.; McCarthy, T.P.; Lee, M.J.; Peyton, S.R. A poly(ethylene glycol) three-dimensional bone marrow hydrogel. Biomaterials 2022, 280, 121270.
  77. Christoffersson, J.; Aronsson, C.; Jury, M.; Selegård, R.; Aili, D.; Mandenius, C.F. Fabrication of modular hyaluronan-PEG hydrogels to support 3D cultures of hepatocytes in a perfused liver-on-a-chip device. Biofabrication 2018, 11, 015013.
  78. Zapp, C.; Mundinger, P.; Boehm, H. Natural Presentation of Glycosaminoglycans in Synthetic Matrices for 3D Angiogenesis Models. Front. Cell Dev. Biol. 2021, 9, 729670.
  79. Nam, S.; Stowers, R.; Lou, J.; Xia, Y.; Chaudhuri, O. Varying PEG density to control stress relaxation in alginate-PEG hydrogels for 3D cell culture studies. Biomaterials 2019, 200, 15–24.
  80. Chung, T.W.; Tyan, Y.C.; Lin, S.W.; Yang, M.H.; Liu, Y.H.; Wang, R.P. Developing photothermal-responsive and anti-oxidative silk/dopamine nanoparticles decorated with drugs which were incorporated into silk films as a depot-based drug delivery. Int. J. Biol. Macromol. 2021, 185, 122–133.
  81. Sgambato, A.; Cipolla, L.; Russo, L. Bioresponsive Hydrogels: Chemical Strategies and Perspectives in Tissue Engineering. Gels 2016, 2, 28.
  82. Xing, R.; Mustapha, O.; Ali, T.; Rehman, M.; Zaidi, S.S.; Baseer, A.; Batool, S.; Mukhtiar, M.; Shafique, S. Development, Characterization, and Evaluation of SLN-Loaded Thermoresponsive Hydrogel System of Topotecan as Biological Macromolecule for Colorectal Delivery. Biomed. Res. Int. 2021, 2021, 9968602.
  83. Witika, B.A.; Stander, J.C.; Smith, V.J.; Walker, R.B. Nano Co-Crystal Embedded Stimuli-Responsive Hydrogels: A Potential Approach to Treat HIV/AIDS. Pharmaceutics 2021, 13, 127.
  84. Huynh, V.; Ifraimov, N.; Wylie, R.G. Modulating the Thermoresponse of Polymer-Protein Conjugates with Hydrogels for Controlled Release. Polymers 2021, 13, 2772.
  85. Qing, W.; Xing, X.; Feng, D.; Chen, R.; Liu, Z. Indocyanine green loaded pH-responsive bortezomib supramolecular hydrogel for synergistic chemo-photothermal/photodynamic colorectal cancer therapy. Photodiagnosis Photodyn. Ther. 2021, 36, 102521.
  86. Lin, X.; Miao, L.; Wang, X.; Tian, H. Design and evaluation of pH-responsive hydrogel for oral delivery of amifostine and study on its radioprotective effects. Colloids Surf. B Biointerfaces 2020, 195, 111200.
  87. Ghani, M.; Heiskanen, A.; Thomsen, P.; Alm, M.; Emnéus, J. Molecular-Gated Drug Delivery Systems Using Light-Triggered Hydrophobic-to-Hydrophilic Switches. ACS Appl. Bio Mater. 2021, 4, 1624–1631.
  88. Fan, L.; Zhang, X.; Liu, X.; Sun, B.; Li, L.; Zhao, Y. Responsive Hydrogel Microcarrier-Integrated Microneedles for Versatile and Controllable Drug Delivery. Adv. Healthc. Mater. 2021, 10, e2002249.
  89. Abedi, F.; Davaran, S.; Hekmati, M.; Akbarzadeh, A.; Baradaran, B.; Moghaddam, S.V. An improved method in fabrication of smart dual-responsive nanogels for controlled release of doxorubicin and curcumin in HT-29 colon cancer cells. J. Nanobiotechnology 2021, 19, 18.
  90. Najafipour, A.; Gharieh, A.; Fassihi, A.; Sadeghi-Aliabadi, H.; Mahdavian, A.R. MTX-Loaded Dual Thermoresponsive and pH-Responsive Magnetic Hydrogel Nanocomposite Particles for Combined Controlled Drug Delivery and Hyperthermia Therapy of Cancer. Mol. Pharm. 2021, 18, 275–284.
  91. Huang, Y.; Wang, Z.; Zhang, G.; Ren, J.; Yu, L.; Liu, X.; Yang, Y.; Ravindran, A.; Wong, C. A pH/redox-dual responsive, nanoemulsion-embedded hydrogel for efficient oral delivery and controlled intestinal release of magnesium ions. J. Mater. Chem. B 2021, 9, 1888–1895.
  92. Lazarus, G.S.; Cooper, D.M.; Knighton, D.R.; Percoraro, R.E.; Rodeheaver, G.; Robson, M.C. Definitions and guidelines for assessment of wounds and evaluation of healing. Wound Repair Regen. 1994, 2, 165–170.
  93. Gonzalez, A.C.; Costa, T.F.; Andrade, Z.A.; Medrado, A.R. Wound healing—A literature review. An. Bras. Dermatol. 2016, 91, 614–620.
  94. Percival, N.J. Classification of Wounds and their Management. Surgery 2002, 20, 114–117.
  95. Golinko, M.S.; Clark, S.; Rennert, R.; Flattau, A.; Boulton, A.J.; Brem, H. Wound emergencies: The importance of assessment, documentation, and early treatment using a wound electronic medical record. Ostomy Wound Manag. 2009, 55, 54–61.
  96. Moore, K.; McCallion, R.; Searle, R.J.; Stacey, M.C.; Harding, K.G. Prediction and monitoring the therapeutic response of chronic dermal wounds. Int. Wound J. 2006, 3, 89–96.
  97. Herndon, D.N.; Barrow, R.E.; Rutan, R.L.; Rutan, T.C.; Desai, M.H.; Abston, S. A comparison of conservative versus early excision. Therapies in severely burned patients. Ann. Surg. 1989, 209, 547–553.
  98. Dhivya, S.; Padma, V.V.; Santhini, E. Wound dressings—A review. Biomedicine 2015, 5, 22.
  99. Pan, Z.; Ye, H.; Wu, D. Recent advances on polymeric hydrogels as wound dressings. APL Bioeng. 2021, 5, 011504.
  100. Tavakoli, S.; Klar, A.S. Advanced Hydrogels as WoundDressings. Biomolecules 2020, 10, 1169.
  101. Fan, F.; Saha, S.; Hanjaya-Putra, D. Biomimetic Hydrogels to Promote Wound Healing. Front. Bioeng. Biotechnol. 2021, 9, 718377.
  102. Liang, Y.; He, J.; Guo, B. Functional Hydrogels as Wound Dressing to Enhance Wound Healing. ACS Nano 2021, 15, 12687–12722.
  103. Ying, H.; Zhou, J.; Wang, M.; Su, D.; Ma, Q.; Lv, G.; Chen, J. In situ formed collagen-hyaluronic acid hydrogel as biomimetic dressing for promoting spontaneous wound healing. Mater. Sci. Eng. C Mater. Biol. Appl. 2019, 101, 487–498.
  104. Ding, C.; Tian, M.; Feng, R.; Dang, Y.; Zhang, M. Novel Self-Healing Hydrogel with Injectable, pH-Responsive, Strain-Sensitive, Promoting Wound-Healing, and Hemostatic Properties Based on Collagen and Chitosan. ACS Biomater. Sci. Eng. 2020, 6, 3855–3867.
  105. Zhu, L.; Chen, L. Facile design and development of nano-clustery graphene-based macromolecular protein hydrogel loaded with ciprofloxacin to antibacterial improvement for the treatment of burn wound injury. Polym Bull. 2021, 1–16, online ahead of print.
  106. Khaliq, T.; Sohail, M.; Minhas, M.U.; Ahmed Shah, S.; Jabeen, N.; Khan, S.; Hussain, Z.; Mahmood, A.; Kousar, M. Self-crosslinked chitosan/κ-carrageenan-based biomimetic membranes to combat diabetic burn wound infections. Int. J. Biol. Macromol. 2022, 197, 157–168.
  107. Haidari, H.; Kopecki, Z.; Sutton, A.T.; Garg, S.; Cowin, A.J.; Vasilev, K. pH-Responsive “Smart” Hydrogel for Controlled Delivery of Silver Nanoparticles to Infected Wounds. Antibiotics 2021, 10, 49.
  108. Zhang, L.; Zhou, Y.; Su, D.; Wu, S.; Zhou, J.; Chen, J. Injectable, self-healing and pH responsive stem cell factor loaded collagen hydrogel as a dynamic bioadhesive dressing for diabetic wound repair. J. Mater. Chem. B 2021, 9, 5887–5897.
  109. Wang, Y.; Wu, Y.; Long, L.; Yang, L.; Fu, D.; Hu, C.; Kong, Q.; Wang, Y. Inflammation-Responsive Drug-Loaded Hydrogels with Sequential Hemostasis, Antibacterial, and Anti-Inflammatory Behavior for Chronically Infected Diabetic Wound Treatment. ACS Appl. Mater. Interfaces 2021, 13, 33584–33599.
  110. Huang, L.; Shi, Y.; Li, M.; Wang, T.; Zhao, L. Plasma Exosomes Loaded pH-Responsive Carboxymethylcellulose Hydrogel Promotes Wound Repair by Activating the Vascular Endothelial Growth Factor Signaling Pathway in Type 1 Diabetic Mice. J. Biomed. Nanotechnol. 2021, 17, 2021–2033.
  111. Yang, J.; Chen, Z.; Pan, D.; Li, H.; Shen, J. Umbilical Cord-Derived Mesenchymal Stem Cell-Derived Exosomes Combined Pluronic F127 Hydrogel Promote Chronic Diabetic Wound Healing and Complete Skin Regeneration. Int. J. Nanomed. 2020, 15, 5911–5926.
  112. Zhao, Y.; Song, S.; Ren, X.; Zhang, J.; Lin, Q.; Zhao, Y. Supramolecular Adhesive Hydrogels for Tissue Engineering Applications. Chem. Rev. 2022, 122, 5604–5640.
  113. Cascone, S.; Lamberti, G. Hydrogel-based commercial products for biomedical applications: A review. Int. J. Pharm. 2020, 573, 118803.
  114. Sun, Y.; Yang, C.; Zhu, X.; Wang, J.J.; Liu, X.Y.; Yang, X.P.; An, X.W.; Liang, J.; Dong, H.J. 3D printing collagen/chitosan scaffold ameliorated axon regeneration and neurological recovery after spinal cord injury. J. Biomed. Mater. Res. A 2019, 107, 1898–1908.
  115. Ning, L.; Ning, Z.; Mohabatpour, F.; Sarker, M.D.; Schreyerd, D.J.; Chen, X. Bioprinting Schwann cell-laden scaffolds from low-viscosity hydrogel compositions. J. Mater. Chem. B 2019, 7, 4538–4551.
  116. Loh, E.Y.X.; Mohamad, N.; Fauzi, M.B.; Ng, M.H.; Ng, S.F.; Mohd Amin, M.C.I. Development of a bacterial cellulose-based hydrogel cell carrier containing keratinocytes and fibroblasts for full-thickness wound healing. Sci. Rep. 2018, 8, 2875.
  117. Li, J.; Zhang, D.; Guo, S.; Zhao, C.; Wang, L.; Ma, S.; Guan, F.; Yao, M. Dual-enzymatically cross-linked gelatin hydrogel promotes neural differentiation and neurotrophin secretion of bone marrow-derived mesenchymal stem cells for treatment of moderate traumatic brain injury. Int. J. Biol. Macromol. 2021, 187, 200–213.
More
Information
Contributors MDPI registered users' name will be linked to their SciProfiles pages. To register with us, please refer to https://encyclopedia.pub/register : , , , , , , , ,
View Times: 323
Revisions: 2 times (View History)
Update Date: 20 Jul 2022
1000/1000