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 -- 2527 2023-01-05 12:40:39 |
2 format + 2 word(s) 2529 2023-01-06 03:50:10 |

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.
Kruk, K.;  Winnicka, K. Applications of Alginates. Encyclopedia. Available online: https://encyclopedia.pub/entry/39792 (accessed on 01 July 2024).
Kruk K,  Winnicka K. Applications of Alginates. Encyclopedia. Available at: https://encyclopedia.pub/entry/39792. Accessed July 01, 2024.
Kruk, Katarzyna, Katarzyna Winnicka. "Applications of Alginates" Encyclopedia, https://encyclopedia.pub/entry/39792 (accessed July 01, 2024).
Kruk, K., & Winnicka, K. (2023, January 05). Applications of Alginates. In Encyclopedia. https://encyclopedia.pub/entry/39792
Kruk, Katarzyna and Katarzyna Winnicka. "Applications of Alginates." Encyclopedia. Web. 05 January, 2023.
Applications of Alginates
Edit

Alginates (ALG) have been used in biomedical and pharmaceutical technologies for a long time. ALG are natural polymers occurring in brown algae and feature multiple advantages, including biocompatibility, low toxicity and mucoadhesiveness. Moreover, ALG demonstrate biological activities per se, including anti-hyperlipidemic, antimicrobial, anti-reflux, immunomodulatory or anti-inflammatory activities. ALG are characterized by gelling ability, one of the most frequently utilized properties in the drug form design. ALG have numerous applications in pharmaceutical technology that include micro- and nanoparticles, tablets, mucoadhesive dosage forms, wound dressings and films. However, there are some shortcomings, which impede the development of modified-release dosage forms or formulations with adequate mechanical strength based on pure ALG. Other natural polymers combined with ALG create great potential as drug carriers, improving limitations of ALG matrices. 

alginate polymer blends biomedical applications

1. Introduction

Beneficial properties of alginates (ALG) enable their wide application in the pharmaceutical, medical, and food industries [1]. ALG were discovered at the end of the nineteenth, century and their commercial production started in the United States in 1929. In 1983, the Food and Drug Administration (FDA) approved the use of ALG as a food ingredient [2]. Since then, ALG have been investigated thoroughly and have been the subject of many studies. ALG are polymers of natural origin, and their main source of acquisition is brown algae (Phaeophyta). They occur most frequently in algae cell walls [2]; however, they can be found in the intracellular matrix in Laminaria and Fucus species [3]. The most common form of ALG, which appears in cell walls, is calcium salt [4], with sodium or potassium salts appearing less frequently. The algae species most commonly used for ALG acquisition are inter alia: Ascophyllun, Macrocystis, Laminaria, Eisenia, Ecklonia, Fucus, Alario, Necrocystis, Sargassum, Ascophyllum nodosum, Macrocystis porifera, and Laminaria [1][2][5][6]. Another valid means of obtaining ALG is technology using bacteria from Pseudomonas sp. or Azotobacter (Acetobacter) sp.

2. Pharmaceutical Applications of ALG

Among many advantages of ALG, SA monographs are included in both European Pharmacopeia and United States Pharmacopeia [7][8]; thus, their properties for pharmaceutical and biomedical use are strictly regulated. Due to the low toxicity, biocompatibility and inert character, European Food Safety Authority (EFSA) approved ALG for use in a wide range of products, even in young children with special delivery requirements. The versality of ALG leads them to be widely applied in drug delivery systems, such as micro-, nanoparticles, tablets, semi-solid drug forms (as thickener and suspending agent), in tissue engineering, bone and cartilage regeneration, and in wound treatment (ALG dressings). Drug dosage forms designed with the use of ALG are mainly micro-, nanoparticles, tablets, capsules, hydrogels, and beads [9][10].
Such a wide application of ALG is also possible due to other numerous advantages, such as gelation properties, cell affinity, and high capacity to incorporate and release various active substances, including proteins [9]. Moreover, ALG are natural and relatively inexpensive polymers. Mucoadhesiveness of ALG makes them a promising excipient for mucosal (ocular [11], nasal [12], vaginal [13], oral [14]) delivery systems.

2.1. Microparticles

ALG microparticles are being studied to protect the drug from degradation as well as modify, release, and increase drug bioavailability. Drugs encapsulated in microparticles can be both hydrophilic and hydrophobic. Due to biocompatibility, biodegradability, non-toxicity, relatively low cost, properties favorable for the spray-drying process, and ease of gelation, SA presents itself as an excellent matrix material for microparticles technology [15][16][17][18]. ALG microparticles are biocompatible and highly bioavailable dosage forms [19], they are relatively small (1–500 μm), possess a large surface area, protect the core from external agents, mask taste and odor, and enable different release modification [20][21][22][23]. Microparticles are also delivery platforms for proteins, nucleic acids, enzymes, cells, growth factors and genes in tissue engineering [24][25][26][27]. However, disadvantages of microparticles include relative high production costs, specialized equipment, low reproducibility of the process and instability of the final product [28]. External gelation is one of the most often used methods to obtain microparticles by dropping ALG solution into the crosslinking medium [29]. Particle size is a key parameter in the context of drug distribution in the gastrointestinal tract, drug release and degradation process [15]. Currently, the tendency is to receive the finest possible microparticle size by enhancing the dripping method through the involvement of external forces, such as vibration, electrostatic force, or coaxial air flow. It is also possible to improve the emulsification method by using sonification, membrane emulsification, or microfluidic methods [30][31]. Microcapsules obtained by external gelation provide a protective coating for sensible components such as living cells, cytokines, growth factors and proteins [32][33][34][35][36]. ALG microparticles are semi-permeable and provide immune protection to cells, which allows them to differentiate and proliferate [37]. Martin et al. [38] proposed a novel mucoadhesive system consisting of ALG microparticles loaded with nystatin for the treatment of oral candidiasis to increase residence time on the buccal mucosa and to improve the effectiveness of the treatment. Designed microparticles demonstrated antifungal activity for up to 48 h in an in vitro study. An interesting approach to the topic was demonstrated by Benavides et al. [39], Faidi et al. [40] and Fermandizz et al. [41], who proposed thyme, clove essential and cedar wood oil-loaded ALG microparticles to protect against evaporation and disintegration under the influence of oxygen, UV, and heat. All the researchers obtained microparticles with relatively high encapsulation efficiency, however further studies are required to refine the conditions for microparticle production. Hussein N. et al. [42] designed ALG microparticles loaded with ropinirole hydrochloride using the spray-drying method for intranasal delivery. Ropinirole is an agonist of D2 dopamine receptors in the brain, employed in the treatment of Parkinson’s disease in monotherapy or in combination with other drugs, most frequently levodopa. The drug possesses an approximately 50% hepatic first-pass effect; thus, it is characterized by low bioavailability. ALG microparticles were utilized as a drug form for intranasal use to increase ropinirole bioavailability by reducing the first-pass effect, since the nasal mucosa possesses a relatively large absorption area (approx. 150 cm2), and the drug reaches its target site of action faster: the nervous system. In addition, ALG microparticles, due to their mucoadhesive properties, prolonged the contact of the formulation with the mucosa. The study showed that designed particles were non-toxic to an isolated sheep mucosa and were stable for 2 months of storage at 5 °C ± 1 and 20 °C ± 2. There have also been attempts to load insulin into ALG microparticles [43][44][45][46]. To improve the comfort of insulin administration, novel drug delivery systems, including ALG microparticles, are being developed to protect insulin from degradation. Mild conditions for producing microparticles by the internal gelation method did not damage its secondary structure [44], and microparticles prepared from SA blended with mucin showed a hypoglycemic effect in diabetic rats comparable to subcutaneous administration of insulin [45]. Insulin loaded into microparticles by the spray drying maintained its bioactivity [46].
ALG microparticles were also utilized to improve the stability of papain during storage and enteral release, which extended the shelf life of papain during storage to about 3.6 years, compared to free papain (about 0.48 years) and a higher rate of drug release at pH >6.8 than at pH <5 [47]. A similar premise supported the design of hydrogel beads with subtilisin on a matrix of SA blended with guar gum [48] or the encapsulation of plasmid DNA (pDNA) encoding a green fluorescent protein (GFP)-reported gene using SA [49]. Szekalska M. et al. [50] designed ALG microspheres with ranitidine hydrochloride to improve their bioavailability after oral administration, and the obtained formulations were characterized by sustained drug release and beneficial mucoadhesive properties.
Microparticles can also provide protection for orally administered probiotic bacteria from harmful environmental factors, such as stomach acidic pH, and high temperature during the manufacturing and storage process. Faarez I.M. et al. [51] placed Lactobacillus plantarum lactic acid bacteria (LAB)12 cells in ALG microcapsules encapsulated with cellulose derivatives: methylcellulose (MC), sodium carboxymethylcellulose (NaCMC) or hydroxypropyl methylcellulose (HPMC) to increase heat resistance. Bacteria cells gained higher survivability during storage when HPMC or MC were blended with ALG. Mirmazloum I. et al. [52] encapsulated Lactobacillus acidophilus with Reishi medicinal mushroom (Ganoderma lingzhi) extract as a prebiotic for the bacteria in a matrix composed of SA blended with maltose, HPMC or hydroxyethyl cellulose (HEC) to increase the stability of the formulation and to mask the bitter taste of the fungus, which resulted in increased stability of the formulation during storage and slower release, especially with the addition of maltose.

2.2. Nanoparticles

Nanoparticles are characterized by the same order of size as proteins and by a relatively large surface area, which creates the possibility of placing specific ligands [53][54][55]. In addition, nanoparticles enable modifications in drug pharmacokinetics, they can reduce drug toxicity and the possibility of damaging healthy cells, and they enable precise drug delivery to the targeted site of action, especially in targeted cancer therapy [55]. Nanoparticles can also improve solubility and bioavailability of poorly water-soluble substances and can provide modified drug release while reducing toxicity [10][15]. Moreover, nanoparticle size enables passage through the smallest capillaries [15], and they are characterized by the ability to inhibit P-glycoprotein activity, thus to reduce the resistance of tumor cells to cytostatics [56]. In order to improve the bioavailability of metformin, to reduce its side effects and to obtain sustained drug release, Kumar S. et al. [57] developed metformin-loaded ALG nanoparticles. The study conducted on adult Wistar albino rats showed sustained drug release from nanoparticles (up to 30 h), and a lower dose of the drug was needed. Thomas D. et al. [58] proposed ALG nanoparticles loaded with rifampicin to improve drug bioavailability after oral administration and to obtain controlled drug release. The study showed a pH-dependent release profile with controlled release of the drug for 6 h at pH 7.4. Ahmad Z. et al. [59] developed ALG nanoparticles with isoniazid, rifampicin, pyrazinamide or ethambutol as an oral delivery system for tuberculosis treatment to improve pharmacokinetics of these drugs and to reduce the high potential of side effects. The researchers created nanoparticles with high encapsulation efficiency and higher bioavailability of encapsulated drugs compared to free drugs. Higher tissue concentrations of drugs were observed on a Laca mice model after administration in encapsulated form compared to free drugs. In addition, tissue concentrations were maintained at >MIC levels for 15 days, making it possible to use these drugs less frequently than daily. Kirtane A.R. et al. [14] developed ALG nanoparticles for a chemiotherapeutic drug, doxorubicin, in order to improve its oral bioavailability. A study on Madin–Darby Canine Kidney II cells (MDCK) showed improved doxorubicin transport across tumor cells overexpressing P-glycoprotein after nanoparticle administration. Studies on a mice model showed higher oral drug bioavailability during nanoparticle administration compared to free drugs. Bakhshi M. et al. [60] designed a vaccine against Escherichia coli 0157:H7 consisting of IgY encapsulated within ALG nanoparticles. The study showed that IgY was released from the nanoparticles in the stomach in a minimal amount (up to 10%); thus, the formulation managed to protect the drug against low pH and did not adversely affect the biological activity of immunoglobulins.

2.3. Tablet Technology

ALG are used in a variety of applications in tablet technology. SA at a concentration of 1–5% acts as a disintegrating agent in powder form and in the form of a 1% solution as a binder. ALG are also used as fillers as well as taste and odor maskers. Controlled drug release might be achieved through the application of ALG matrix tablets, which undergo superficial swelling and slow dissolution. Thus, drugs that are well soluble in water slowly dissolve in water, flowing into the tablet interior, and such a solution diffuses to the outside of the tablet. In contrast, substances that hardly dissolve in water are released from the matrix by erosion of the tablet as it passes through the gastrointestinal tract [61]. The biopharmaceutical properties of ALG tablets significantly depend on intrinsic factors, namely the properties of the ALG itself. The molecular weight directly proportionally affects the viscosity of the ALG solution, and thus, an increase in viscosity entails a slower release from the matrix tablet [62]. The concentration of ALG in the tablet affects the release rate in a similar way [63]. The M/G ratio is also not without influence, ALG with higher G-content form stiffer gel structures, which slow down drug release [29][64].

3. Biomedical Applications of ALG

3.1. Tissue Regeneration

The application of ALG in tissue engineering and bone and cartilage regeneration seems interesting and potentially promising, as treating damage in these tissues is often a difficult and lengthy process [65]. The most desirable strategy is to stimulate osteogenesis and chondrogenesis in situ [39], which can be achieved by supplying the site of damage with stem cells capable of proliferation and differentiation [39][66][67][68][69][70][71]. As ALG gels are characterized by a structure similar to the extracellular matrix in tissues, they are being studied for potential use in tissue engineering or cell transplantation [65]. The principle of gel action is to deliver cells to a specific location in the body and to provide conditions for tissue reconstruction [72][73]. The influence of cells such as osteoblasts, chondrocytes, or bone marrow mesenchymal cells (MCSs) on osteogenesis and chondrogenesis is being studied [66][74][75][76][77]. Numerous studies showed bone regeneration using injectable ALG scaffolds containing MCSs [70][71][74][75][76][77][78][79]. ALG can be inserted into tissues in a non-invasive manner, they fill irregular spaces accurately, are easily chemically modified, and possess good regenerative properties, as proven in an animal model [68][80]. In mouse embryonic stem cell studies, the ability of ALG to promote stem cell differentiation into bone cells was demonstrated [81]. ALG gels are biodegradable and do not possess sufficient mechanical properties to allow for load transfer during the initial stages of regeneration [65]. In order to improve mechanical properties, ALG were mixed with ceramides, hydroxyapatite [82][83], CH [84], or bio-glass [85]. ALG gels might also be carriers for growth factors such as bone morphogenetic proteins (BMPs) [31] or tumor growth factor β (TNF-β) [74][75][76][78].

3.2. Wound Care

One of the most common dressings used for centuries has been gauze, which is easy to use, inexpensive, and has a high water absorption capacity. However, it can stick to the wound and cause re-damage during dressing changes [39][86]. Current emphasis is on modern dressings providing a moist wound environment while managing the exudate. Among many advantages of ALG dressings are biocompatibility, optimal water vapor permeability, mild antiseptic properties combined with non-toxicity, and biodegradability [39][87]. The principle of ALG dressings is to absorb exudate from the wound, exchange calcium ions from the dressing for sodium ions from the wound, convert to gel form and provide moisture to the wound. These processes promote granulation and epithelialization of the epidermis and thus wound healing [65][88][89]. The ratio of mannuronic to guluronic residues affects the ability to absorb exudate [90][91]. A high content of mannuronic acid is positively correlated with the ability to retain water; however, the fibers of such dressings are weaker. Some studies have indicated that pathogens from the wound were trapped in the gelled structure of the dressing [92]. The manufacturing process of ALG dressings begins with crosslinking the ALG with calcium ions and impregnating the material with the resulting gel. Such a semi-finished product is freeze-dried and mechanically smoothed to obtain flexible, delicate fiber mats [93] or foam sheets [94]. To improve the properties of ALG dressings, several compounds such as silver [95], zinc ions [96][97][98][99][100][101][102][103][104], chitosan (CH), fucoidan, asiaticoside [105], gelatin (GEL) [39], polyvinyl alcohol (PVA) [106][107], or cellulose [108] can be introduced. The aim of adding these components is to improve antibacterial properties (silver, zinc ions, CH), swelling rate, tensile strength (PVA), and other mechanical properties (fucoidan, GEL, cellulose). Another excipient added to ALG dressings are antibiotics, such as clindamycin [107] vancomycin [109], aminoglycosides [110], curcumin [99], aloe vera [104], or active carbon [111] for the elimination of unpleasant wound odor.

References

  1. Szekalska, M.; Puciłowska, A.; Szymańska, E.; Ciosek, P.; Winnicka, K. Alginate: Current use and future perspectives in pharmaceutical and biomedical applications. Int. J. Polym. Sci. 2016, 2016, 7697031.
  2. Sachan, N.K.; Pushkar, S.; Jha, A.; Bhattcharya, A. Sodium alginate: The wonder polymer for controlled drug delivery. J. Pharm. Res. 2009, 2, 1191–1199.
  3. Kloareg, B.; Quatrano, R.S. Structure of the cell walls of marine algae and ecophysiological functions of the matrix polysaccharides. Oceanogr. Mar. Biol. Annu. Rev. 1988, 26, 259–315.
  4. Myklestad, S. Ion-exchange properties of brown algae I. Determination of rate mechanism for calciumhydrogen ion exchange for particles from Laminaria hyperborea and Laminaria digitata. J. Appl. Chem. 1968, 18, 30–36.
  5. Guo, X.; Wang, Y.; Qin, Y.; Shen, P.; Peng, Q. Structures, properties and application of alginic acid: A review. Int. J. Biol. Macromol. 2020, 1, 618–628.
  6. Goh, C.H.; Heng, P.W.S.; Chan, L.W. Alginates as a useful natural polymer for microencapsulation and therapeutic applications. Carbohydr. Polym. 2012, 88, 1–12.
  7. Council of Europe. The European Pharmacopoeia 10.0; Council of Europe: Strasbourg, France, 2020.
  8. The United States Pharmacopeia. “The United States Pharmacopeial Convention”, USP 44-NF 39; The United States Pharmacopeial Convention: Rockville, MD, USA, 2021.
  9. Puscaselu, R.G.; Lobiuc, A.; Dimian, M.; Covasa, M. Alginate: From food industry to biomedical applications and management of metabolic disorders. Polymers 2020, 12, 2417.
  10. Severino, P.; da Silva, C.F.; Andrade, L.N.; de Lima Oliveira, D.; Campos, J.; Souto, E.B. Alginate nanoparticles for drug delivery and targeting. Curr. Pharm. Des. 2019, 25, 1312–1334.
  11. Motwani, S.K.; Chopra, S.; Talegaonkar, S.; Kohli, K.; Ahmad, F.J.; Khar, R.K. Chitosan-sodium alginate nanoparticles as submicroscopic reservoirs for ocular delivery: Formulation, optimisation and in vitro characterization. Eur. J. Pharm. Biopharm. 2008, 68, 513–525.
  12. Dehghan, S.; Kheiri, M.T.; Abnous, K.; Eskandari, M.; Tafaghodi, M. Preparation, characterization and immunological evaluation of alginate nanoparticles loaded with whole inactivated influenza virus: Dry powder formulation for nasal immunization in rabbits. Microb. Pathog. 2018, 115, 74–85.
  13. Wong, T.W.; Dhanawat, M.; Rathbone, M.J. Vaginal drug delivery: Strategies and concerns in polymeric nanoparticle development. Expert. Opin. Drug. Deliv. 2014, 11, 1419–1434.
  14. Kirtane, A.R.; Narayan, P.; Liu, G.; Panyam, J. Polymer-surfactant nanoparticles for improving oral bioavailability of doxorubicin. J. Pharm. Investig. 2017, 47, 65–73.
  15. Jadach, B.; Świetlik, W.; Froelich, A. Sodium alginate as a pharmaceutical excipient: Novel applications of a well-known polymer. J. Pharm. Sci. 2022, 111, 1250–1261.
  16. Bowey, K.; Neufeld, R.J. Systemic and mucosal delivery of drugs within polymeric microparticles produced by spray drying. BioDrugs 2010, 24, 359–377.
  17. Lopes, M.; Abrahim, B.; Veiga, F.; Seiça, R.; Cabral, M.R.; Arnaud, P.; Andrade, J.C.; Ribeiro, A.J. Preparation methods and applications behind alginate-based particles. Expert. Opin. Drug Deliv. 2017, 14, 769–782.
  18. Ranjan, S.; Fontana, F.; Ullah, H.; Hirvonen, J.; Santos, H.A. Microparticles to enhance delivery of drugs and growth factors into wound sites. Ther. Deliv. 2016, 7, 711–732.
  19. Birnbaum, D.T.; Brannon-Peppas, L. Microparticle Drug Delivery Systems. In Drug Delivery Systems in Cancer Therapy. Cancer Drug Discovery and Development; Brown, D.M., Ed.; Humana Press: Totowa, NJ, USA, 2004; pp. 117–135.
  20. Patel, R.P.; Baria, A.H.; Pandya, N.B. Stomach-specific drug delivery of famotidine using floating alginate beads. Int. J. PharmTech Res. 2009, 1, 288–291.
  21. Patel, N.; Lalwani, D.; Gollmer, S.; Injeti, E.; Sari, Y.; Nesamony, J. Development and evaluation of a calcium alginate based oral ceftriaxone sodium formulation. Prog. Biomater. 2016, 5, 117–133.
  22. Agarwal, T.; Narayana, S.N.; Pal, K.; Pramanik, K.; Giri, S.; Banerjee, I. Calcium alginate-carboxymethyl cellulose beads for colon-targeted drug delivery. Int. J. Biol. Macromol. 2015, 75, 409–417.
  23. De Cicco, F.; Russo, P.; Reverchon, E.; García-González, C.A.; Aquino, R.P.; Del Gaudio, P. Prilling and supercritical drying: A successful duo to produce core-shell polysaccharide aerogel beads for wound healing. Carbohydr. Polym. 2016, 147, 482–489.
  24. Springer, M.L.; Hortelano, G.; Bouley, D.M.; Wong, J.; Kraft, P.E.; Blau, H.M. Induction of angiogenesis by implantation of encapsulated primary myoblasts expressing vascular endothelial growth factor. J. Gene. Med. 2000, 2, 279–288.
  25. Tan, H.; Huang, D.; Lao, L.; Gao, C. RGD modified PLGA/gelatin microspheres as microcarriers for chondrocyte delivery. J. Biomed. Mater. Res. B Appl. Biomater. 2009, 91, 228–238.
  26. Patil, S.B.; Sawant, K.K. Mucoadhesive microspheres: A promising tool in drug delivery. Curr. Drug Deliv. 2008, 5, 312–318.
  27. Basmanav, B.F.; Kose, G.T.; Hasirci, V. Sequential growth factor delivery from complexed microspheres for bone tissue engineering. Biomaterials 2008, 29, 4195–4204.
  28. Ribeiro, A.J.; Neufeld, R.J.; Arnaud, P.; Chaumeil, J.C. Microencapsulation of lipophilic drugs in chitosan-coated alginate microspheres. Int. J. Pharm. 1999, 187, 115–123.
  29. Urtuvia, V.; Maturana, N.; Acevedo, F.; Peña, C.; Díaz-Barrera, A. Bacterial alginate production: An overview of its biosynthesis and potential industrial production. World J. Microbiol. Biotechnol. 2017, 33, 198.
  30. Zhang, C.; Grossier, R.; Candoni, N.; Veesler, S. Preparation of alginate hydrogel microparticles by gelation introducing cross-linkers using droplet-based microfluidics: A review of methods. Biomater. Res. 2021, 25, 41.
  31. Wang, Y.-L.; Hu, J.J. Sub-100-micron calcium-alginate microspheres: Preparation by nitrogen flow focusing, dependence of spherical shape on gas streams and a drug carrier using acetaminophen as a model drug. Carbohydr Polym. 2021, 269, 118262.
  32. Serra, M.; Correia, C.; Malpique, R.; Brito, C.; Jensen, J.; Bjorquist, P.; Carrondo, M.J.; Alves, P.M. Microencapsulation technology: A powerful tool for integrating expansion and cryopreservation of human embryonic stem cells. PLoS ONE 2011, 6, e23212.
  33. Kong, H.J.; Smith, M.K.; Mooney, D.J. Designing alginate hydrogels to maintain viability of immobilized cells. Biomaterials 2003, 24, 4023–4029.
  34. Man, Y.; Wang, P.; Guo, Y.; Xiang, L.; Yang, Y.; Qu, Y.; Gong, P.; Deng, L. Angiogenic and osteogenic potential of platelet-rich plasma and adipose-derived stem cell laden alginate microspheres. Biomaterials 2012, 33, 8802–8811.
  35. Yu, J.; Du, K.T.; Fang, Q.; Gu, Y.; Mihardja, S.S.; Sievers, R.E.; Wu, J.C.; Lee, R.J. The use of human mesenchymal stem cells encapsulated in RGD modified alginate microspheres in the repair of myocardial infarction in the rat. Biomaterials 2010, 31, 7012–7020.
  36. Freiberg, S.; Zhu, X.X. Polymer microspheres for controlled drug release. Int. J. Pharm. 2004, 282, 1–18.
  37. Sun, J.; Tan, H. Alginate-Based Biomaterials for Regenerative Medicine Applications. Materials 2013, 6, 1285–1309.
  38. Martín, M.J.; Calpena, A.C.; Fernández, F.; Mallandrich, M.; Gálvez, P.; Clares, B. Development of alginate microspheres as nystatin carriers for oral mucosa drug delivery. Carbohydr. Polym. 2015, 117, 140–149.
  39. Benavides, S.; Cortes, P.; Parada, J.; Franco, W. Development of alginate microspheres containing thyme essential oil using ionic gelation. Food Chem. 2016, 204, 77–83.
  40. Faidi, A.; Lassoued, M.A.; Becheikh, M.E.H.; Touati, M.; Stumbé, J.F.; Farhat, F. Application of sodium alginate extracted from a Tunisian brown algae Padina pavonica for essential oil encapsulation: Microspheres preparation, characterization and in vitro release study. Int. J. Biol. Macromol. 2019, 136, 386–394.
  41. Ferrandiz, M.; Lopez, A.; Franco, E.; Garcia-Garcia, D.; Fenollar, D.; Balart, R. Development and characterization of bioactive alginate microcapsules with cedarwood essential oil. Flavour Fragran. J. 2017, 32, 184–190.
  42. Hussein, N.; Omer, H.; Ismael, A.; Albed Alhnan, M.; Elhissi, A.; Ahmed, W. Spray-dried alginate microparticles for potential intranasal delivery of ropinirole hydrochloride: Development, characterization and histopathological evaluation. Pharm. Dev. Technol. 2020, 25, 290–299.
  43. Meneguin, A.B.; Silvestre, A.L.P.; Sposito, L.; de Souza, M.P.C.; Sábio, R.M.; Araújo, V.H.S.; Cury, B.S.F.; Chorilli, M. The role of polysaccharides from natural resources to design oral insulin micro- and nanoparticles intended for the treatment of Diabetes mellitus: A review. Carbohydr. Polym. 2021, 256, 117504.
  44. Reis, C.P.; Ribeiro, A.J.; Neufeld, R.J.; Veiga, F. Alginate microparticles as novel carrier for oral insulin delivery. Biotechnol. Bioeng. 2007, 96, 977–989.
  45. Builders, P.F.; Kunle, O.O.; Okpaku, L.C.; Builders, M.I.; Attama, A.A.; Adikwu, M.U. Preparation and evaluation of mucinated sodium alginate microparticles for oral delivery of insulin. Eur. J. Pharm. Biopharm. 2008, 70, 777–783.
  46. Bowey, K.; Swift, B.E.; Flynn, L.E.; Neufeld, R.J. Characterization of biologically active insulin-loaded alginate microparticles prepared by spray drying. Drug. Dev. Ind. Pharm. 2013, 39, 457–465.
  47. Sankalia, M.G.; Mashru, R.C.; Sankalia, J.M.; Sutariya, V.B. Papain entrapment in alginate beads for stability improvement and site-specific delivery: Physicochemical characterization and factorial optimization using neural network modeling. AAPS PharmSciTech 2005, 6, E209–E222.
  48. Simi, C.K.; Emilia Abraham, T. Encapsulation of crosslinked subtilisin microcrystals in hydrogel beads for controlled release applications. Eur. J. Pharm. Sci. 2007, 32, 17–23.
  49. Nograles, N.; Abdullah, S.; Shamsudin, M.N.; Billa, N.; Rosli, R. Formation and characterization of pDNA-loaded alginate microspheres for oral administration in mice. J. Biosci. Bioeng. 2012, 113, 133–140.
  50. Szekalska, M.; Amelian, A.; Winnicka, K. Alginate microspheres obtained by the spray drying technique as mucoadhesive carriers of ranitidine. Acta Pharm. 2015, 65, 15–27.
  51. Fareez, I.M.; Lim, S.M.; Zulkefli, N.A.A.; Mishra, R.K.; Ramasamy, K. Cellulose derivatives enhanced stability of alginate-based beads loaded with Lactobacillus plantarum LAB12 against low pH, high temperature and prolonged storage. Probiotics Antimicrob. Proteins 2018, 10, 543–557.
  52. Mirmazloum, I.; Ladányi, M.; Omran, M.; Papp, V.; Ronkainen, V.-P.; Pónya, Z.; Papp, I.; Némedi, E.; Kiss, A. Co-encapsulation of probiotic Lactobacillus acidophilus and Reishi medicinal mushroom (Ganoderma lingzhi) extract in moist calcium alginate beads. Int. J. Biol. Macromol. 2021, 192, 461–470.
  53. Choukaife, H.; Doolaanea, A.A.; Alfatama, M. Alginate nanoformulation: Influence of process and selected variables. Pharmaceuticals 2020, 13, 335.
  54. He, L.; Shang, Z.; Liu, H.; Yuan, Z.-X. Alginate-based platforms for cancer-targeted drug delivery. Biomed. Res. Int. 2020, 2020, 17.
  55. Nitta, S.K.; Numata, K. Biopolymer-based nanoparticles for drug/gene delivery and tissue engineering. Int. J. Mol. Sci. 2013, 14, 1629–1654.
  56. Suri, S.S.; Fenniri, H.; Singh, B. Nanotechnology-based drug delivery systems. J. Occup. Med. Toxicol. 2007, 2, 16.
  57. Kumar, S.; Bhanjana, G.; Verma, R.K.; Dhingra, D.; Dilbaghi, N.; Kim, K.H. Metformin-loaded alginate nanoparticles as an effective antidiabetic agent for controlled drug release. J. Pharm. Pharmacol. 2017, 69, 143–150.
  58. Thomas, D.; KurienThomas, K.; Latha, M.S. Preparation and evaluation of alginate nanoparticles prepared by green method for drug delivery applications. Int. J. Biol. Macromol. 2020, 154, 888–895.
  59. Ahmad, Z.; Pandey, R.; Sharma, S.; Khuller, G.K. Pharmacokinetic and pharmacodynamic behaviour of antitubercular drugs encapsulated in alginate nanoparticles at two doses. Int. J. Antimicrob. Agents 2006, 27, 409–416.
  60. Bakhshi, M.; Ebrahimi, F.; Nazarian, S.; Zargan, J.; Behzadi, F.; Gariz, D.S. Nano-encapsulation of chicken immunoglobulin (IgY) in sodium alginate nanoparticles: In vitro characterization. Biologicals 2017, 17, 69–75.
  61. Sanchez-Ballester, N.M.; Bataille, B.; Soulairol, I. Sodium alginate and alginic acid as pharmaceutical excipients for tablet formulation: Structure-function relationship. Carbohydr. Polym. 2021, 270, 118399.
  62. Andriamanantoanina, H.; Rinaudo, M. Relationship between the molecular structure of alginates and their gelation in acidic conditions. Polym. Int. 2010, 59, 1531–1541.
  63. Rehm, B.H.; Valla, S. Bacterial alginates: Biosynthesis and applications. Appl. Microbiol. Biotechnol. 1997, 48, 281–288.
  64. Chee, S.Y.; Wong, P.K.; Wong, C.L. Extraction and characterization of alginate from brown seaweeds (Fucales, Phaeophyceae) collected from Port Dickson, Peninsular Malaysia. J. Appl. Phycol. 2011, 23, 191–196.
  65. Lee, K.Y.; Mooney, D.J. Alginate: Properties and biomedical applications. Prog. Polym. Sci. 2012, 37, 106–126.
  66. Alsberg, E.; Anderson, K.W.; Albeiruti, A.; Franceschi, R.T.; Mooney, D.J. Cell-interactive alginate hydrogels for bone tissue engineering. J. Dent. Res. 2001, 80, 2025–2029.
  67. Abbah, S.A.; Lu, W.W.; Chan, D.; Cheung, K.M.; Liu, W.G.; Zhao, F.; Li, Z.Y.; Leong, J.C.; Luk, K.D. In vitro evaluation of alginate encapsulated adipose-tissue stromal cells for use as injectable bone graft substitute. Biochem. Biophys. Res. Commun. 2006, 347, 185–191.
  68. Durrieu, M.C.; Pallu, S.; Guillemot, F.; Bareille, R.; Amedee, J.; Baquey, C.H.; Labrugère, C.; Dard, M. Grafting RGD containing peptides onto hydroxyapatite to promote osteoblastic cells adhesion. J. Mater. Sci. Mater. Med. 2004, 15, 779–786.
  69. Grellier, M.; Granja, P.L.; Fricain, J.C.; Bidarra, S.J.; Renard, M.; Bareille, R.; Bourget, C.; Amédée, J.; Barbosa, M.A. The effect of the co-immobilization of human osteoprogenitors and endothelial cells within alginate microspheres on mineralization in a bone defect. Biomaterials 2009, 30, 3271–3278.
  70. Jin, H.H.; Kim, D.H.; Kim, T.W.; Shin, K.K.; Jung, J.S.; Park, H.C.; Yoon, S.Y. In vitro evaluation of porous hydroxyapatite/chitosan-alginate composite scaffolds for bone tissue engineering. Int. J. Biol. Macromol. 2012, 51, 1079–1085.
  71. Rubert, M.; Monjo, M.; Lyngstadaas, S.P.; Ramis, J.M. Effect of alginate hydrogel containing polyproline-rich peptides on osteoblast differentiation. Biomed. Mat. 2012, 7, 055003.
  72. Langer, R.; Vacanti, J.P. Tissue engineering. Science 1993, 260, 920–926.
  73. Lee, K.Y.; Mooney, D.J. Hydrogels for tissue engineering. Chem. Rev. 2001, 101, 1869–1879.
  74. Tang, M.; Chen, W.; Weir, M.D.; Thein-Han, W.; Xu, H. Human embryonic stem cel encapsulation in alginate microbeads in macroporous calcium phosphate cement for bone tissue engineering. Acta Biomater. 2012, 8, 3436–3445.
  75. Chen, W.; Zhou, H.; Weir, M.D.; Bao, C.; Xu, H. Umbilical cord stem cells released from alginate-fibrin microbeads inside macroporous and biofunctionalized calcium phosphate cement for bone regeneration. Acta Biomater. 2012, 8, 2297–2306.
  76. Xia, Y.; Mei, F.; Duan, Y.; Gao, Y.; Xiong, Z.; Zhang, T.; Zhang, H. Bone tissue engineering using bone marrow stromal cells and an injectable sodium alginate/gelatin scaffold. J. Biomed. Mater. Res. A 2012, 100, 1044–1050.
  77. Brun, F.; Turco, G.; Accardo, A.; Paoletti, S. Automated quantitative characterization of alginate/hydroxyapatite bone tissue engineering scaffolds by means of micro-CT image analysis. J. Mater. Sci. Mater. Med. 2011, 22, 2617–2629.
  78. Florczyk, S.J.; Leung, M.; Jana, S.; Li, Z.; Bhattarai, N.; Huang, J.; Hopper, R.A.; Zhang, M. Enhanced bone tissue formation by alginate gel-assisted cell seeding in porous ceramic scaffolds and sustained release of growth factor. J. Biomed. Mater. Res. A 2012, 100, 3408–3415.
  79. Kolambkar, Y.M.; Dupont, K.M.; Boerckel, J.D.; Huebsch, N.; Mooney, D.J.; Hutmacher, D.W.; Guldberg, R.E. An alginate-based hybrid system for growth factor delivery in the functional repair of large bone defects. Biomaterials 2011, 32, 65–74.
  80. Zhao, L.; Weir, M.D.; Xu, H.H. An injectable calcium phosphate-alginate hydrogel-umbilical cord mesenchymal stem cell paste for bone tissue engineering. Biomaterials 2010, 31, 6502–6510.
  81. Hwang, Y.S.; Cho, J.; Tay, F.; Heng, J.Y.; Ho, R.; Kazarian, S.G.; Williams, D.R.; Boccaccini, A.R.; Polak, J.M.; Mantalaris, A. The use of murine embryonic stem cells, alginate encapsulation, and rotary microgravity bioreactor in bone tissue engineering. Biomaterials 2009, 30, 499–507.
  82. Fay, D.; Halsey, J. Anatomy of Bones; World Technologies: Bellingham, WA, USA, 2012.
  83. Lin, H.-R.; Yeh, Y.-J. Porous alginate/hydroxyapatite composite scaffolds for bone tissue engineering: Preparation, characterization, and in vitro studies. J. Biomed. Mater. Res. B Appl. Biomater. 2004, 71, 52–65.
  84. Florczyk, S.J.; Kim, D.J.; Wood, D.L.; Zhang, M. Influence of processing parameters on pore structure of 3D porous chitosan-alginate polyelectrolyte complex scaffolds. J. Biomed. Mater. Res. A 2011, 98, 614–620.
  85. Zamani, D.; Moztarzadeh, F.; Bizari, D. Alginate-bioactive glass containing Zn and Mg composite scaffolds for bone tissue engineering. Int. J. Biol. Macromol. 2019, 137, 1256–1267.
  86. Jones, V.J. The use of gauze: Will it ever change? Int. Wound. J. 2006, 3, 79–86.
  87. Ahmad Raus, R.; Wan Nawawi, W.M.F.; Nasaruddin, R.R. Alginate and alginate composites for biomedical applications. Asian J. Pharm. Sci. 2021, 16, 280–306.
  88. Lloyd, L.; Kennedy, J.F.; Methacanon, P.; Paterson, M.; Knill, C.J. Carbohydrate polymers as wound management aids. Carbohydr. Polym. 1998, 37, 315–322.
  89. Bishop, S.M.; Walker, M.; Rogers, A.A.; Chen, W.Y. Importance of moisture balance at the wound-dressing interface. J. Wound Care. 2003, 12, 125–128.
  90. Thomas, A.; Harding, K.G.; Moore, K. Alginates from wound dressings activate human macrophages to secrete tumour necrosis factor-α. Biomaterials 2000, 21, 1797–1802.
  91. Walker, M.; Hobot, J.A.; Newman, G.R.; Bowler, P.G. Scanning electron microscopic examination of bacterial immobilisation in a carboxymethyl cellulose (AQUACEL®) and alginate dressings. Biomaterials 2003, 24, 883–890.
  92. Miraftab, M.; Qiao, Q.; Kennedy, J.F.; Anand, S.C.; Groocock, M.R. Fibres for wound dressings based on mixed carbohydrate polymer fibres. Carbohydr. Polym. 2003, 53, 225–231.
  93. Grothe, T.; Grimmelsmann, N.; Homburg, S.V.; Ehrmann, A. Possible applications of nano-spun fabrics and materials. Mater. Today Proc. 2017, 4, S154–S159.
  94. Namviriyachote, N.; Lipipun, V.; Akkhawattanangkul, Y.; Charoonrut, P.; Ritthidej, G.C. Development of polyurethane foam dressing containing silver and asiaticoside for healing of dermal wound. Asian J. Pharm. Sci. 2019, 14, 63–77.
  95. Wiegand, C.; Heinze, T.; Hipler, U.C. Comparative in vitro study on cytotoxicity, antimicrobial activity, and binding capacity for pathophysiological factors in chronic wounds of alginate and silver-containing alginate. Wound Repair Regen. 2009, 17, 511–521.
  96. Agren, M.S. Zinc in wound repair. Arch. Dermatol. 1999, 135, 1273–1274.
  97. Wiśniewska-Wrona, M.; Kucharska, M.; Struszczyk, M.H.; Cichecka, M.; Wilbik-Hałgas, B.; Szymonowicz, M.; Paluch, D.; Guzińska, K.; Rybak, Z. Hemostatic, resorbable dressing of natural polymers-hemoguard. Autex Res. J. 2016, 16, 29–34.
  98. Xue, W.; Zhang, M.; Zhao, F.; Gao, J.; Wang, L. Long-term durability antibacterial microcapsules with plant-derived Chinese nutgall and their applications in wound dressing. E-Polymers 2019, 19, 268–276.
  99. Dai, M.; Zheng, X.; Xu, X.; Kong, X.; Li, X.; Guo, G.; Luo, F.; Zhao, X.; Wei, Y.Q.; Qian, Z. Chitosan-alginate sponge: Preparation and application in curcumin delivery for dermal wound healing in rat. J. Biomed. Biotechnol. 2009, 2009, 8.
  100. Yan, X.; Khor, E.; Lim, L.Y. PEC films prepared from chitosan-alginate coacervates. Chem. Pharm. Bull. 2000, 48, 941–946.
  101. Venkatrajah, B.; Vanitha Malathy, V.; Elayarajah, B.; Mohan; Rajendran, R.; Rammohan, R. Biopolymer and Bletilla striata herbal extract coated cotton gauze preparation for wound healing. J. Med. Sci. 2012, 12, 148–160.
  102. Holak, P.; Adamiak, Z.; Babinska, I.; Jalynski, M.; Jastrzebski, P.; Grabarczyk, L.; Brzezinski, M.; Bory, J.; Tobolska, A.; Glodek, J. The influence of haemostatic dressing prototypes for the emergency services on the histopathological parameters of porcine muscle. In Vivo 2019, 33, 723–729.
  103. Straccia, M.C.; d’Ayala, G.G.; Romano, L.; Oliva, A.; Laurienzo, P. Alginate hydrogels coated with chitosan for wound dressing. Mar. Drugs 2015, 13, 2890–2908.
  104. Gomez Chabala, L.F.; Cuartas, C.E.E.; Lopez, M.E.L. Release behavior and antibacterial activity of chitosan/alginate blends with aloe vera and silver nanoparticles. Mar. Drugs 2017, 15, 328.
  105. Murakami, K.; Aoki, H.; Nakamura, S.; Nakamura, S.; Takikawa, M.; Hanzawa, M.; Kishimoto, S.; Hattori, H.; Tanaka, Y.; Kiyosawa, T.; et al. Hydrogel blends of chitin/chitosan, fucoidan and alginate as healing-impaired wound dressings. Biomaterials 2010, 31, 83–90.
  106. Kamoun, E.A.; Kenawy, E.-R.S.; Tamer, T.M.; El-Meligy, M.A.; Mohy Eldin, M.S. Poly (vinyl alcohol)-alginate physically crosslinked hydrogel membranes for wound dressing applications: Characterization and bio-evaluation. Arab. J. Chem. 2015, 8, 38–47.
  107. Kim, J.O.; Choi, J.Y.; Park, J.K.; Kim, J.H.; Jin, S.G.; Chang, S.W.; Li, D.X.; Hwang, M.-R.; Woo, J.S.; Kim, J.-A.; et al. Development of clindamycin-loaded wound dressing with polyvinyl alcohol and sodium alginate. Biol. Pharm. Bull. 2008, 31, 2277–2282.
  108. Benselfelt, T.; Wågberg, L. Unidirectional swelling of dynamic cellulose nanofibril networks: A platform for tunable hydrogels and aerogels with 3D shapeability. Biomacromolecules 2019, 20, 2406–2412.
  109. Kurczewska, J.; Pecyna, P.; Ratajczak, M.; Gajecka, M.; Schroeder, G. Halloysite nanotubes as carriers of vancomycin in alginate-based wound dressing. Saudi Pharm. J. 2017, 25, 911–920.
  110. Kumar, L.; Brice, J.; Toberer, L.; Klein-Seetharaman, J.; Knauss, D.; Sarkar, S.K. Antimicrobial biopolymer formation from sodium alginate and algae extract using aminoglycosides. PLoS ONE 2019, 14, 1492.
  111. Osmokrović, A.; Jančić, I.; Jankovic-Castvan, I.; Petrović, P.; Milenkovic, M.; Obradovic, B. Novel composite zinc-alginate hydrogels with activated charcoal aimed for potential applications in multifunctional primary wound dressings. Hem. Ind. 2019, 73, 37–46.
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: 270
Revisions: 2 times (View History)
Update Date: 06 Jan 2023
1000/1000
Video Production Service