<<

Advanced Drug Delivery Reviews 146 (2019) 97–125

Contents lists available at ScienceDirect

Advanced Drug Delivery Reviews

journal homepage: www.elsevier.com/locate/addr

Therapeutic strategies for enhancing in wound

Austin P. Veith a,1, Kayla Henderson a,1, Adrianne Spencer a, Andrew D. Sligar a, Aaron B. Baker a,b,c,d,⁎ a Department of Biomedical Engineering, University of Texas at Austin, Austin, TX, USA b Institute for Cellular and Molecular Biology, University of Texas at Austin, Austin, TX, USA c The Institute for Computational Engineering and Sciences, University of Texas at Austin, Austin, TX, USA d Institute for Biomaterials, Drug Delivery and Regenerative Medicine, University of Texas at Austin, Austin, TX, USA article info abstract

Article history: The enhancement of has been a goal of medical practitioners for thousands of years. The develop- Received 17 March 2018 ment of chronic, non-healing wounds is a persistent medical problem that drives patient morbidity and increases Received in revised form 15 September 2018 healthcare costs. A key aspect of many non-healing wounds is the reduced presence of vessel growth through the Accepted 24 September 2018 process of angiogenesis. This review surveys the creation of new treatments for healing cutaneous wounds Available online 26 September 2018 through therapeutic angiogenesis. In particular, we discuss the challenges and advancement that have been made in delivering biologic, pharmaceutical and cell-based therapies as enhancers of wound vascularity and Keywords: Wound healing healing. Angiogenesis © 2018 Elsevier B.V. All rights reserved. therapy Cell-based therapies therapy Neovascularization Diabetic ulcers Chronic wounds Non-healing wounds

Contents

1. Introduction...... 98 2. Proteintherapeuticsforwoundangiogenesis...... 99 2.1. Growthfactors...... 99 2.1.1. Plateletderivedgrowthfactor...... 99 2.1.2. Epidermalgrowthfactor...... 100 2.1.3. Fibroblastgrowthfactors...... 100 2.1.4. Vascularendothelialgrowthfactor...... 101 2.2. Non-growthfactorproteindelivery...... 101 2.3. Peptides...... 103 2.4. Bloodderivedfactors...... 103 3. Geneandnucleicacid-basedtherapiesforwoundangiogenesis...... 104 3.1. MicroRNAsinwoundangiogenesis...... 104 3.1.1. MicroRNAsthattargetangiogenesisinwoundhealing...... 104 3.2. Deliverymethods...... 107 3.2.1. Freeviralvectordelivery...... 107 3.2.2. Scaffoldsandhydrogelsemployedingenedelivery...... 107 3.2.3. Electroporationandsonoporationenhanceddelivery...... 108 4. Druganddrug-likecompoundsforenhancingwoundangiogenesis...... 108 4.1. Strategiesusingdeliveryofsmallmoleculestopromoteangiogenesisandrescueimpairedwoundhealinginanimalmodels...... 108 4.2. Adenosinetriphosphate...... 108 4.3. Statins...... 109

⁎ Corresponding author at: Department of Biomedical Engineering, University of Texas at Austin, University Station BME 5.202D, Austin, TX 78712, USA. E-mail address: [email protected] (A.B. Baker). 1 Authors contributed equally to this work.

https://doi.org/10.1016/j.addr.2018.09.010 0169-409X/© 2018 Elsevier B.V. All rights reserved. 98 A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125

4.4. Deferoxamine...... 110 4.5. Naturalcompounds...... 111 4.6. Hyaluronanoligosaccharides...... 113 5. Stemcellbasedtherapiesforwoundangiogenesis...... 114 5.1. Challengesandstrategiesfortherapeuticcelldelivery...... 114 5.2. Adiposederivedstemcells...... 114 5.3. Bonemarrowderivedmesenchymalstemcells...... 116 5.4. Inducedpluripotentstemcells...... 118 5.5. Otherstemcellpopulations...... 118 6. Conclusion...... 118 Acknowledgements...... 119 References...... 119

1. Introduction factors that regulate new blood vessel growth [21–23]. The M2 pheno- type is thought to be more strongly associated with angiogenic activity In modern times, non-healing, cutaneous wounds have become [24,25]. The presence of therapeutics and/or biomaterials for delivery of major medical and social burden in the United States and worldwide. compounds can also induce a response from . The foreign Diabetes is the leading cause of chronic, non-healing ulcers that leave body response to some biomaterials involves the formation of foreign patients at increased risk for limb amputation [1,2]. It is estimated that body giant cells and fibrotic encapsulation of the material [26]. The for- chronic wounds cost the U.S. over $25 billion annually and significantly eign body reaction to implanted biomaterials can be separated into contribute to the rising cost of healthcare [3–6]. Astonishingly, diabetic stages including protein adsorption on blood or biofluid contact, initial ulcers are responsible for 25-50% of the total cost of diabetes treatment cells response/recruitment from cells of the innate in- [7] and are the most common cause for limb amputations in the United cluding macrophages, production of cytokines and soluble factors that States [8]. Debridement of necrotic tissues, offloading, infection control, can control inflammation and the recruitment of cells involved in fi- surgical revascularization, and limb elevation/compression are often re- brous tissue formation, and finally formation of fibrous capsule around quired concurrently and lead to the immense cost and complexity the foreign body [27]. The polarization of macrophages can be altered [9,10]. Additionally, a host of techniques and dressings have been ap- by interactions with biomaterials and this is a potential strategy for al- plied with the goal of improving healing, including the tering the course of wound healing and modulating the result of the for- delivery of growth factors, gene therapy, stem cell therapies and eign body reaction [28,29]. mechanical/pressure-based stimulation [11–15]. Currently, only The injured tissue at this stage undergoes rapid proliferation in a platelet-derived growth factor-BB (PDGF-BB; /Regranex), loose provisional (ECM). In a well healing wound, is approved in the U.S. for clinical treatment of chronic wounds. How- this granulation tissue is characterized by intense angiogenesis giving ever, the mixed efficacy of this treatment has prevented it from achiev- it its characteristic pink swollen appearance [30]. Angiogenesis in the ing widespread clinical adoption [16,17]. A major aspect of non-healing provisional matrix is guided by several factors including the develop- wounds in diabetes and peripheral vascular disease is a reduced ability ment of gradients of hypoxia caused by local microvascular disruption to regrow microvasculature through the process of angiogenesis [18]. and cytokines produced by macrophages. Among other processes, hyp- This disruption of the normal healing process contributes to the dys- oxia activates the transcription factor hypoxia-inducible factor-1-alpha functional healing response and is major therapeutic target for creating (HIF-1α), which promotes angiogenesis through multiple mechanisms new treatments for non-healing and ischemic wounds. In this review, including the regulation of a large number of angiogenesis related we will examine recent advances in therapies targeting the enhance- genes [31] and the production of angiogenic growth factors including ment of angiogenesis in wound healing. for Ang-2, stromal cell-derived factor-1 (SDF-1) and vascular endothe- The healing of wounds is a complex process that greatly depends on lial growth factor (VEGF-A) [32]. Heparan sulfate proteoglycans also the extent and mechanism of injury. Broadly speaking, the wound play a key role in regulating the angiogenic activity of many of these healing process can be separated into a several stages that include im- growth factors including VEGF-A and fibroblast growth factor-2 (FGF- mediate hemostasis, acute inflammation, proliferation and maturation 2) [33,34]. [19]. A key accompanying activity in the proliferation stage is the forma- In dermal wounds with robust healing, the angiogenic activity dur- tion of new blood vessels, a process known as angiogenesis. In intact tis- ing the proliferative phase initially creates a disorganized vascular net- sues, the microvasculature remains in a state of homeostasis where work with vessels of high tortuosity and many dead-end flow adequate nutrients and oxygen are delivered to the tissues in balance pathways, often reaching higher vessel numbers than the skin prior to with removal of waste products and carbon dioxide. Upon injury, the injury [35]. Following this peak in vessel density, there is increased ex- microvasculature is disrupted leading to fluid accumulation, inflamma- pression of anti-angiogenic growth factors, such as Sprouty2 and pig- tion and the development of hypoxia. Hypoxia and inflammatory cyto- ment epithelium-derived factor (PEDF), leading to subsequent kines activate endothelial cells, leading to recruitment of immune cells. regression of the vascular network [36,37]. Maturation of the vascular The injured tissue recruits an early influx of followed later network to include larger vessels and prevent regression requires the by macrophages. Macrophages have been recently recognized to polar- stabilization of the capillaries by pericytes and vascular smooth muscle ize into a continuum of phenotypes and the nature of this polarization cells (vSMCs) [38]. A key growth factor in promoting stabilization of the has a critical effect on wound healing, formation and angiogenesis vascular network is PDGF-BB, which aids in recruitment and differenti- [20]. Classically, the polarization of macrophages ranges in a continuum ation of pericytes [39]. In addition, -1 (Ang-1) and from a pro-inflammatory M1 phenotype to alternatively activated M2a- angiopoietin-2 (Ang-2), in combination with the receptor Tie2, play a d phenotypes that have been associated with anti-inflammatory and complex role in regulating angiogenesis and maturity of vascular net- wound healing activity [20]. Macrophages are essential to the coordi- works [40–42]. Specifically, Tie2 is expressed on pericytes and signaling nate the angiogenic response in wound healing and produce many through Ang-2/Tie2 in pericytes can regulate angiogenesis [43,44]. A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125 99

Reduced angiogenesis plays a prominent role in the non-healing na- 2. Protein therapeutics for wound angiogenesis ture of diabetic foot ulcers and other chronic wounds [45]. Multiple mechanisms are supported for the reduced angiogenesis in these 2.1. Growth factors chronic wounds. Diabetic patients with foot ulcers have higher circula- tory levels of PEDF in their plasma in comparison to non-diabetic pa- There have been many years of intensive study into the use of tients and diabetic patients without foot ulcers [46]. In diabetic skin, growth factors for both inducing angiogenesis and enhancing wound there is also a marked reduction in syndecan-4 and glypican-1, two closure in chronic wounds. Endogenous growth factors play a critical key cell surface heparan sulfate proteoglycans that enable effective role in orchestrating the wound healing process and are essential for ef- binding of FGF-2 and other growth factors to their cognate receptors fective wound healing [53]. While the delivery of exogenous growth [47,48]. Similarly, wound exudates isolated from venous leg ulcers factors has shown benefits to wound closure and angiogenesis in animal have anti-angiogenic properties, even in spite of increased expression studies, almost all growth factor-based agents for wound healing have of angiogenic factors [49]. These analyses are consistent with a proteo- not shown clear benefits for enhancing wound healing in patients in mic analysis of venous leg ulcer exudates that found increases in anti- clinical trials. One hypothesis regarding the failure of these trials is angiogenic proteins in non-healing ulcers [50]. In chronic venous ulcers, that human patients develop resistance to angiogenic growth factors there was also increased of proteolysis of VEGF-A [51]aswellasin- during the development of vascular disease and the aging process creased levels of soluble VEGFR-1, which may serve to neutralize [54]. It is also uncertain whether a single growth factor therapy could VEGF-A activity [52]. be effective for wound healing [55]. The mixed results from clinical tri- In this review, we summarize the recent progress in the delivery of als, combined with the prohibitive cost of the therapies, have slowed compounds to wounds with the goal of inducing therapeutic angiogen- the clinical adoption of these treatments, however research is still un- esis. In particular, we will focus on new systems and approaches for de- derway to develop more effective protein therapies for wound healing. livering compounds and cells for enhancing angiogenesis in cutaneous wounds. As non-healing wounds are major clinical problem, we will 2.1.1. Platelet derived growth factor emphasize studies with results relevant to chronic wounds. The review Platelet-derived growth factors (PDGFs) are important is arranged by the overall therapeutic strategy of using proteins, gene/ chemoattractants in the wound healing process. Currently, becaplermin nucleic acid delivery, small molecules/drug-like compounds or cell ther- (100 μg/g PDGF-BB) is the only FDA approved growth factor treatments apies (Fig. 1). for wound healing [16,56]. In some clinical trials, becaplermin has been shown expedite wound healing in venous leg ulcers [57,58]. In a clinical trial on becaplermin for treating diabetic foot ulcers, 48% of patients (29/ 61) had healed ulcers versus 25% (14/57) in the placebo group with a dosage of 2.2 μg/cm2 of ulcer area [59]. Clinical trials in 1998 and 1999 showed statistically significant improvements in this type of chronic wound with becaplermin treatment [60,61]. In the groups treated once a day with becaplermin, diabetic foot ulcers exhibited decreased wound areas and faster healing rates over the placebo controls. These positive results can be attributed to the ability of PDGF-BB to enhance recruitment of cells to the wound site and the stimulation of an angio- genic response. However, other clinical trials have not found improve- ments in wound healing with becaplermin treatment [62]. Therapeutic resistance may also play a significant role in preventing the effectiveness of PDGF-BB therapies, thus necessitating higher con- centrations and increasing the risk of side effects of the growth factor therapy. In addition, the FDA gave becaplermin a black box warning due to a five-fold increase in death rate among patients that received three tubes of becaplermin and who had pre-existing malignancy [63]. While the risk of getting a new cancer was not increased in patients without cancer, this finding highlights the risk of using pro-growth and angiogenic therapies in patients with an existing cancer. Novel de- livery strategies are needed to increase the efficacy of PDGF-BB without increasing the dose and to increase its cost effectiveness [64]. The simultaneous application of multiple growth factors and pro- teins may provide the extra benefit needed to see therapies applied in a clinical environment. The simultaneous application of PDGF-BB and (EGF) has been shown to direct pericyte re- cruitment to help stabilize blood vessels [65,66]. Other multiple growth factor studies have shown that the simultaneous application of PDGF-BB and FGF-2 has improved vascular stability in hind limb ischemia models in rats and rabbits [67]. The dual delivery of PDGF-BB and VEGF-A from electrospun nanofiber scaffolds accelerated wound healing in rats and promoted angiogenesis at the wound site [68]. Similar results were achieved in the simultaneous delivery of VEGF-A and PDGF-BB bound by laminin heparin binding domains from fibrin matrices [69]. In an- other study, researchers created an electrospun nanofiber scaffold of Fig. 1. Summary diagram of the cellular and molecular components of wound healing and and (HA) that released FGF-2, EGF, VEGF-A wound angiogenesis (created using BioRender and Adobe Illustrator). and PDGF-BB with varying release kinetics. This construct accelerated 100 A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125

Table 1 Multi-growth factor approaches for enhancing wound healing and angiogenesis.

Growth factors Results Exp. model Ref.

PDGF-BB and VEGF-A Promotes angiogenesis at the wound site, with significantly increased numbers of capillaries in the wound bed Rats [68] PDGF-BB and FGF-2 Increased the levels of neovascularization significantly Rats [67] PDGF-BB, FGF-2, VEGF-A and EGF Accelerates wound closure and enhances the maturation of blood vessels in the wound bed in diabetic rats Diabetic Rats [70] PlGF domains and VEGF-A Chimeric protein promoted angiogenesis with reduced vascular permeability Rats [135] PDGF-BB and IGF-1 Increased numbers of blood vessels at the wound site, Increased VEGF expression Rats [324]

wound closure and the maturation of blood vessels in the wound beds comparison to the non-EGF dressings and control wounds. Finally, to of rats with streptozotocin-induced diabetes [70]. Finally, combining address another major issue in protein therapeutics, dextrin conjuga- PDGF-BB with a syndecan-4 proteoliposomes significantly increased tion has been used to reduce the proteolytic degradation of EGF in circu- its activity in wound closure, angiogenesis and increased the proportion lation, thereby increasing its wound healing capacity [90]. By day 11 in a of M2 macrophages in a diabetic mouse model of wound healing [71]. study of full thickness excisional wounds in diabetic mice, neo-dermal Table 1 summarizes the recent studies using multiple growth factors tissue generation was greatest in the group that received both 1 μg/mL to enhance angiogenesis in wound healing. and 10 μg/mL of dextrin-EGF over the control that received 10 μg/mL of EGF without dextrin conjugation. The dextrin-EGF construct also ac- 2.1.2. Epidermal growth factor celerated wound closure and the formation of granulated tissue at the Epidermal growth factor (EGF) is another promising growth factor 10 μg/mL dose over the control. Further research into novel materials for wound healing therapies and a known mediator of angiogenesis. and protective strategies will help develop more effective EGF therapies. Early clinical trials showed reduced healing times and increase re- epithelialization in wounds treated with recombinant EGF [72,73], how- 2.1.3. growth factors ever, concerns over the role of EGF in cancer have reduced enthusiasm Fibroblast growth factors (FGFs) are a family of growth factors that for these therapies in the U.S. [74]. Despite this fact, there has been in- show great promise for applications in dermal wound healing as they creased interest in EGF treatment to enhance wound healing in chronic affect a large number of physical processes [91]. There are several differ- wounds in other parts of the world. There are several commercially ent FGF family members expressed during the healing process including available forms of recombinant human EGF outside of the U.S. including FGF-1, FGF-2, FGF-7, FGF-10 and FGF-22 [92,93]. In general, FGF ligands Easyef (Nepidermin), Regen-D 150, and Heberprot-P. Easyef is available bind to their receptors in the presence of heparin or heparan sulfate and as a spray or ointment and has shown showed benefits for diabetic foot these signals can promote angiogenesis and proliferation/migration of ulcers in a prospective, open-label trial, crossover study with 89 patients endothelial cells [94]. The most studied member of the FGF family [75]. Regen-D 150 is a gel formulation containing EGF that has shown used in wound healing applications is FGF-2 [95], partly due to its strong improvements in wound healing time and reduces the risk of amputa- association with wound angiogenesis. The other FGFs are more often as- tion in patients with chronic diabetic foot ulcers [76]. Heberprot-P is a sociated with the enhancement of re-epithelization [93,96,97], the most lyophilized formulation of EGF that has been tested in over 100,000 pa- notable being FGF-7. FGF-7 has been shown to be integral in the migra- tients in Cuba and has also been shown to increase healing of chronic ul- tion of keratinocytes to the wound area [98] and a lack of FGF-7 delays cers and reduce diabetes-related amputations [77]. Finally, clinical trials cutaneous wound healing in mice [99]. Further, the dual delivery of in have supported that recombinant human EGF has positive FGF-7 and FGF-2 has been shown to enhance skin wound healing, healing effects on moderate to severe diabetic foot ulcers [78,79]. showing more developed vascular networks than the scaffold control Mechanistically, EGF has been shown to enhance angiogenesis [100]. through phosphatidylinositide 3-kinase/protein kinase B (PI3K/Akt) Multiple studies have reported the benefits of delivery of FGF-2 for and extracellular signaling-regulated kinase-1/2 (ERK1/2) through wound healing. In freeze injured skin grafts, the delivery of FGF-2 stim- VEGF-mediated pathways [80]. EGF can induce production of VEGF-A ulated angiogenesis and accelerated wound healing [101]. More recent to enhance tumor angiogenesis in cancer cells [81] and is a powerful studies have corroborated this evidence in a wide range of wound chemoattractant for keratinocyte migration to encourage wound re- models from skin flaps to incisional hernias [102,103]. In one study, epithelialization [82] with heparin-bound EGF being the one of the FGF-2 released from heparin-fibrin composite matrices increased fibro- most potent forms [83]. EGF also regulates nuclear factor-κB (NF-κB) blast proliferation and collagen density, decreased the necrotic area in controlled activation of CCCTC-binding factor (CTCF), a gene regulator the skin flap and the vascular density in comparison to control wounds that plays a significant role in epithelial cell migration [84]. In mice [102]. Research on incisional hernias has shown that tissues treated with a knockout of the NF-κB p50 subunit, the wound closure percent- with FGF-2 after the incision had increased breaking strength and age was significantly reduced in comparison to wild type (WT) control lower hernia development over untreated incisions [103]. In clinical mice upon EGF treatment [85]. studies, FGF-2 increased wound healing in patients with burns and Recently, new delivery methods have been explored to help localize chronic ulcers [104,105]. A second clinical study also demonstrated in- EGF and increase its circulation time [86]. Silk biomaterials have been creased wound closure in patients with burns and enhanced strength developed to deliver EGF to promote angiogenesis and wound closure in wounds treated with FGF-2 [106]. in cutaneous excisional wounds in mice [87,88]. In a study by Gil et al. Novel FGF-2 delivery methods have been explored to enhance its three different material structures and two different drug incorporation wound healing activity. Delivery of FGF-2 from gelatin sheets methods were examined [87]. The silk biomaterial was formulated as a crosslinked with glutaraldehyde accelerated re-epithelialization, forma- solid film, a porous film, or electrospun nanofibers. Silk biomaterials tion of granulation tissue and angiogenesis in comparison to control show wound closure percentages significantly greater than control wounds in a mouse model [107]. For injectable applications, chitosan- wounds in a full thickness dermal wound model. The delivery of EGF heparin hydrogels have been developed to deliver FGF-2 via injection from silk biomaterials decreased scar tissue formation and increased ep- to subcutaneous wound sites [108]. When the gel was injected into ithelialization over a Tegaderm control. Another group developed a the ischemic hind limb of a diabetic mice, there was increased blood hyaluronic acid (HA)/collagen sponge that could be used to deliver vessel formation after four days. Furthermore, there was a significant in- EGF as a wound dressing [89]. In a wound model in rats, the EGF- crease in blood flow and oxygen saturation in the calf and ankle until at containing dressing increased wound closure and angiogenesis in least day 28. Diabetic patients have a reduction in cell surface A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125 101 proteoglycans that are essential for robust FGF-2 signaling, including infiltration at a greater rate than the control treatments. VEGF-A encap- syndecan-4 and glypican-1 [48,54]. Several studies support that co- sulated PLGA microspheres and nanoparticles released from chitosan/ delivering syndecan-4 or glypican-1 in a proteoliposome with FGF-2 hyaluronic acid hydrogels and fibrin scaffolds respectively have also can markedly enhance its activity for angiogenesis and wound healing been used to promote angiogenesis in non-healing infected wounds in diabetic animal models [47,48,54,109,110]. and in diabetic mice [126,127]. The release of VEGF-A from calcium algi- The delivery of growth factors and specifically FGF-2 from nano- nate microspheres has been shown to increase capillary network forma- particles and nanocrystals has also been investigated [111–113]. tion and neovascularization as well in subcutaneous implantation Mesoporous silica nanoparticles (MSNs) are a means to create con- models in rats [128,129]. When histological analysis was performed, trolled drug release profiles by fine tuning the surface area to volume all sites displayed a low level of fibrotic tissue and the microspheres ratio, pore sizes, and surface properties. Zhang et al. delivered FGF-2 loaded with VEGF-A showed considerable capillary growth and newly from MSNs with a formulation that allowed release over 20 days formed blood vessels. [111]. In addition, Li et al. combined tunability of gelatin micro- Placenta growth factor (PlGF) has also been shown to enhance an- spheres and the structural stability of a collagen/cellulose nanocrystal giogenesis and can have angiogenic synergy with VEGF-A [130–132]. scaffold shown by Li et al. provided a suitable environment for en- Mice overexpressing PlGF exhibit increased vascularization in both hancing angiogenic growth [112]. Microspheres, impregnated with fetal and adult life [133]. These mice display hypervascularized skin FGF-2, were loaded into a scaffold composed of collagen and cellulose and enlarged blood vessels. Gene transfer of PlGF increases wound an- nanocrystals (CNCs), which increased the mechanical stability of the giogenesis in health and diabetic mice [131]. PlGF also promotes chemo- collagen network. In a subcutaneous implantation model, the FGF-2 taxis of mesenchymal stem cells (MSCs), which may express angiogenic releasing gelatin microspheres significantly increased the number of factors or induce increased angiogenesis/angiogenic factor secretion in new blood vessels compared to FGF-2 from the scaffold alone or con- other cell types. When MSCs were cultured with fibroblast-like trol group. synoviocytes, the synoviocytes were found to secrete higher levels of PlGF and induce greater angiogenesis compared to fibroblast-like 2.1.4. Vascular endothelial growth factor synoviocyte injection controls [134]. The ECM binding domains from The vascular endothelial growth factor (VEGF) family is com- PlGF have also been used to engineer variants of other growth factors prised of five separate growth factors, VEGF-A, -B, -C, -D and pla- (VEGF-A, PDGF-BB and BMP-2) with super-affinity to the ECM, leading centa growth factor (PlGF) [114]. These growth factors are to increased repair in chronic wounds in rodents, over their wild-type primarily responsible for regulating angiogenesis and counterparts [135]. The overall concept is that the localization of the lymphangiogenesis in both embryonic and adult development growth factor may be better controlled with super-affinity to the ECM, [114]. The VEGF family also plays a significant role in the perme- reducing the side effects of angiogenic therapy. The super-affinity ability of blood vessels, vasodilation, and have been shown to stabi- VEGF-A variant promoted angiogenesis with reduced induction of per- lize new blood vessel growth during wound healing. However, the meability of the vasculature for equivalent wild type VEGF-A treatment application of VEGF therapies in clinical settings has been limited, [135]. due in part to the vasopermeability induced by the treatment [115]. Several clinical trials have been performed using recombi- 2.2. Non-growth factor protein delivery nant VEGF-A to enhance angiogenesis in critical limb or myocardial ischemia [116–119], but the clinical benefits of VEGF therapy were Many other proteins have displayed therapeutic effects on wound minor and further studies were not pursued. angiogenesis in preclinical wound healing models. is a key ex- During the course of wound healing, the levels of VEGF-A and FGF-2 ample of a protein hormone shown to be a regulator of angiogenesis in the serum have a negative correlation with one another, first high and wound healing [136,137]. The delivery of insulin from PLGA micro- levels of FGF-2 involved in the initial recruitment of endothelial cells spheres encapsulated in an alginate sponge dressing enhanced healing followed by high levels of VEGF-A used to stabilize the new vessel and reduced scarring from burn injuries in rats by stimulating angiogen- growth and provide a prolonged angiogenic signaling period [120]. esis [137]. In addition, injection of insulin into the skin of mice induced Losi et al. showed that concurrent delivery of FGF-2 and VEGF-A in- angiogenesis suggesting it is a suitable candidate for ischemic wounds creased the number of endothelial cells around the immediate vicinity [138]. Furthermore, in a double-blind placebo-controlled clinical trial, of their scaffold in a hind limb ischemia model, as well as increased per- the topical delivery of insulin to diabetic skin ulcers resulted in im- fusion of blood in the hind limbs of rats after ischemic injury [121]. Fur- proved wound healing [139]. This result presents insulin as a low-cost ther, the delivery of granulocyte- colony stimulating factor alternative to growth factor therapies. (GM-CSF) to diabetic patients with leg ulcers increased the level of The hormone (EPO) is another non-growth factor VEGF-A produced by macrophages and monocytes [95,122]. Application protein that has been shown to increase angiogenesis [140–143]. In of VEGF-A to wounds in db/db mice increased wound closure and angio- second-degree burn injury model where a substantial portion of genesis [123]. In wounds treated with VEGF-A there was a greater num- the original vasculature is damaged or lost completely, mice and ber of endothelial progenitor cells that lead to new blood vessel rats treated with topical application or infusion of EPO displayed formation, as well as increased amounts of pro-angiogenic soluble rates of angiogenesis that were significantly increased over the con- factors. trol that did not receive EPO [140]. Further, wound closure rates in To further enhance the effect of VEGF-A at the wound site, many the rats treated with EPO were accelerated over the control, with therapies also involve the delivery of the growth factor from biomate- the wounds on average 98.8% closed by day 7. In addition, EPO has rials. During degradation, poly(lactic-co-glycolic acid) (PLGA) scaffolds been used to treat ischemic injuries, where significant revasculariza- hydrolyze into lactic and glycolic acids [124]. This lactate helps promote tion is required to restore adequate blood flow [143,144], and it is angiogenesis and helps recruit endothelial progenitor cells at wound suggested that EPO works synergistically with VEGF to promote an- sites and, when combined with VEGF-A, promotes healing [125]. giogenesis at the wound site [145]. On the other hand, EPO has VEGF-A released from PLGA scaffolds lead to significantly greater been implicated in tumor angiogenesis in hepatic tumors and Lewis wound closure in full thickness skin wounds in diabetic mice (N80% of lung carcinoma tumors [146,147], so similar precautions should be the initial wound area) 10 days post injury compared to controls that re- taken as in many growth factor therapies. ceived only VEGF-A or the scaffold alone. The PLGA scaffolds promoted Stromal-cell derived factor-1 (SDF-1) is another protein that angiogenesis at the wound site and the tissue showed greater collagen has been shown to regulate angiogenesis by recruiting endothe- deposition, dermal remodeling, granulation, and blood capillary lial progenitor cells (EPC) to the wound site [148,149]and 102 A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125 promoting sustained cell proliferation [150]. SDF-1 is also a scaffold provided a proliferative environment for the cells and upon strong chemotactic agent for monocytes, such as macrophages introduction to a skin wound model in mice, the 1F9 scaffold helped [151], has been shown to promote the local recruitment of to promote wound healing and angiogenesis. Moreover, 1F9 treated monocytes with anti-inflammatory activity enhances the expan- wounds experienced marked increases in wound closure over the sion of local vasculature [152]. SDF-1 has also been shown to controls. Thymosin Beta 4 (Tβ4) is another protein has been shown promote bone mesenchymal stem cells (bMSC) and EPC chemo- to increase wound healing and angiogenesis in diabetic rats [157] taxis to wound sites [153,154]. The migration of bMSCs and and improve tissue repair. Tβ4 accelerated wound closure, with signif- EPCs promotes vascularization and in turn helps accelerate the icantly greater wound closing rates over control groups. Further, Tβ4 wound healing process. SDF-1 has also been shown to promote was shown to promote angiogenesis at the wound site as seen by angiogenesis in a diabetic mouse model, with a three-fold in- marked increased in endothelial cells and greater neovascularization. crease in blood vessel formation over a control group that did Relative capillary density was also increased in rats that were treated not receive SDF-1 [149]. The delivery of SDF-1 significantly accel- with Tβ4. erated wound closure as well as vascularization in full thickness The development of therapeutic resistance is one hypothesis for the skin wounds. Finally, in a novel approach, Vågesjö et al. trans- reason for the failure of clinical trials for growth factor therapies formed lactobacilli with a plasmid encoding SDF-1 [155]. These [48,110,158]. In diabetic patient skin, there is a marked reduction in bacteria were then applied topically to wound sites in mouse syndecan-4 and glypican-1 [47,48] proteins that are key in stabilizing models of hyperglycemia and peripheral ischemia, leading to ac- the interactions of growth factors with their receptors [109]. celerated wound closure and efficacy of bioavailable SDF-1. Syndecan-4 has been shown to modulate the migration of fibroblasts Other non-growth factor-based protein therapeutics have also and keratinocytes in dermal wounds [159]. Delivery of syndecan-4 in been explored for enhancing angiogenesis during wound healing. a liposomal formulation led to enhanced revascularization in a hind Spidroin, the protein found in spider’s silk, has been explored for its limb ischemia model in mice, diabetic ob/ob mice, and rats advanced scaffolding capacity [156]. Recombinant spidroin (1F9) was [54,109,110]. This treatment also enhanced wound closure and angio- used as a cell scaffold to promote the proliferation of cells while un- genesis in response to treatment with FGF-2 or PDGF-BB (Fig. 2A-E) dergoing bioresorption. 1F9 could potentially be used in a wide vari- [47,71]. Formulating recombinant syndecan-4 in a proteoliposome en- ety of applications. When cultured with 3T3 fibroblasts, the 1F9 hanced the activation of FGF-2 mediated signaling, increased nuclear

Fig. 2. Wound angiogenesis enhancement by syndecan-4 proteoliposomes. (A) von-Willebrand Factor (vWF) staining in the wound bed. Bar = 250 um. (B) Quantification of the small and (C) large vessels in the wound bed. (D) PECAM-1 fluorescence immunostaining for (red), αSMA (green) and nuclei (blue). Bar = 50 um. (E) Quantification of PECAM-1+/α-SMA+ vessels ⁎ † in the wound beds. p b 0.05 versus control group. p b 0.05 versus control and S4PL groups (p b 0.05). Used with permission from [71]. A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125 103 trafficking of FGF-2 and causes increased recycling of syndecan-4/FGF-2 cells, levels of VEGF-A and proliferating keratinocytes. Other natu- during uptake [47,110]. Glypican-1 proteoliposomes also enhanced re- ral AMPs from non-human sources have been shown to increase vascularization in a hind limb ischemia model in WT and ob/ob mice angiogenesis and wound healing. The topical delivery of 200 and increased FGF-2 activity on endothelial cells [48]. Monteforte et al. μg/mL of CW49, a peptide derived from frog’sskin,hasshownto isolated exosomes by immunoselection from glioblastoma cells (one increase angiogenesis in vitro and in vivo, through HUVEC tube for- of the most angiogenic tumors), applied them in a mouse hind limb is- mation assays and a diabetic mouse model respectively [182]. Sim- chemia model and found marked improvement in the recovery of per- ilarly, tylotoin, a peptide present in salamander skin, showed fusion and formation of blood vessels [160]. Other examples of growth comparable wound healing capability with EGF in a full thickness factor co-receptors include the neuropilins (NRP1, NRP2), which help dermal wound mouse model [183]. The topical application of 20 modulate the angiogenic potential of VEGF-A [161–163]. Thus, novel ef- μLof20μg/mL tylotoin to the wound site twice a day resulted in fective protein therapies may involve the delivery of co-receptors to en- almost complete closure of the wounds by day 10, similar to a hance growth factor signaling. treatment of 20 μL of 100 μg/mL EGF. Combined with tylotoin’sin- duction of endothelial tube formation in vitro,suggeststhatitmay 2.3. Peptides play an angiogenic role in wound healing. Artificial AMPs engineered in lab have also shown promise for The delivery of short, anti-microbial peptides (AMPs) for therapeu- wound healing applications. SR-0379 is a key example of the new tic angiogenesis and wound healing is a promising development in the wave of artificial AMPs [184]. In full thickness skin ulcers in rats, field [164,165]. These peptides can be cost effective [166], and may treatment with 200 μg/mL of SR-0379 significantly reduced the offer an efficacious alternative to traditional growth factor therapies. wound area over a 60 μg/mL dose of FGF2 after just two days and Cathelicidins are a notable class of natural AMPs, LL-37 being the decreased the wound healing time overall. These results are similar only human cathelicidin [167–169]. Chereddy et al. showed improved to another artificial AMP, WRL3 [185]. In a MRSA infected burn wound closure in a full thickness excisional wound model in mice wound model in mice, the delivery of 4 μg/mL of WRL3, increased with up to a 7.5 mg/wound dose of LL-37 delivered from a tunable, angiogenesis, as shown by an increased number of blood vessels nanoparticle carrier. The wounds that were treated with LL-37 nano- upon histological analysis, at the burn site as well as helped miti- particles experienced an average wound closure percentage of 90% gate the infection. by day 10, which was significantly higher than the negative control groups, who received either only LL-37 or PLGA scaffold with nano- 2.4. Blood derived factors particles. Further, LL-37 delivery upregulated IL-6 and VEGF-A produc- tion and showed higher re-epithelialization than the control tissues. Blood derived factors present a promising new area of research in Other groups have also seen the upregulation of VEGF-A by LL-37 de- the field. Platelets contain growth factors and cytokines that can en- livery [170]. Concerns remain over LL-37 treatment, as elevated levels hance normal wound healing process and the delivery of a high num- of the peptide in circulation can result in the hemolytic activity ber of these factors to a wound site could provide a relatively [171,172], and, in specific cases, the delivery of LL-37 may promote straightforward way to enhance wound healing and angiogenesis. the growth of other pathogens [173] or inhibit the innate immune re- Platelet-rich plasma (PRP) is derived from gradient centrifugation sponse to bacteria [174]. However, engineered hybrid versions of the of a patient’s whole blood [186], and some components of PRP in- LL-37 peptide have been shown to mitigate this danger [175]. AP-57 clude PDGF-BB, TGF-β1, FGF-2, VEGF-A, EGF, and platelet factor-4 is another natural AMP that is commonly found in the mucosa of to name a few [187]. The potential strengths of PRP therapy include the stomach and colon [176]. A dose of 40 μg per treatment of AP- the fact that autologous PRP is easily obtainable and that it may 57 from nanoparticles delivered in an amphiphilic polylactic acid mimic the healing process with a combination of growth factors, fi- (PLA)-Pluronic F127-PLA block copolymer hydrogel was sufficient to brin and other components, as to a single growth factor therapy. accelerate wound closure and promote angiogenesis [177]. The deliv- Some studies have supported that PRP can improve neovasculariza- ery of AP-57 to the wound site increased the number of endothelial tion and wound repair [188], however, the clinical evidence for cells in the area and increased the vascular density significantly after using PRP in chronic wounds comes from trials with relatively 14 days in vivo. small patient populations, making the current support for use in LL-37 derivatives have also shown promise in enhancing angiogen- humans inconclusive [189]. esis in wounds [178]. LLKKK18 is an 18-mer peptide derivative of LL- In animal studies, PRP therapy can induce neovascularization and 37 that displays the most powerful lipopolysaccharide neutralizing ac- has also been shown to promote graft survival in chondrocutaneous tivity of all the peptide derivatives of LL-37 [179]. For burn wound composite grafts in rabbit ears [190]. Treatment with 1 mL of PRP signif- treatments, the delivery of LLKKK18 in Carbopol hydrogels reduced icantly increased the number of endothelial cells and the number of the wound size faster than the control, especially in the first 6 days. microvessels per field of view upon histological analysis. The expression The level of re-epithelialization was increased in the groups that of VEGF-A also increased significantly in the PRP group in comparison were treated with LLKKK18. Further, LLKKK18 helps to regulate the the control group that did not receive PRP injections. Qui et al. also levels of reactive oxygen species (ROS), which are present in high showed that the controlled release of PRP from a hydrogel composed quantities in the initial stages of healing following burn injury. Finally, of poly(d,l-lactide)-poly(ethylene glycol)-poly(d,l-lactide) (PDLLA- burns treated with LLKKK18 displayed a three-fold increase in the PEG-PDLLA: PLEL) triblock copolymer increased cellular migration and number of new blood vessels at the wound site as determined by α- tube formation significantly over the control [191]. Further, the con- SMA staining. trolled release of 0.8 mL of PRP significantly increased the wound clo- Vasoactive intestinal peptide (VIP) is a 28-amino acid long pep- sure in rabbits over PRP alone and the control. Additionally, PRP tide that has angiogenic and wound healing properties [180,181]. delivered from a PLEL scaffold significantly increased the number of In one study, VIP was loaded into PCL nanospun fibers and deliv- blood vessels in comparison to PRP without the scaffold or the control ered to skin wounds on the backs of mice. The wounds that were groups. Therapy with PRP may also promote angiogenesis by modulat- treated with VIP experienced higher closure percentages than the ing the tissue to secrete inflammatory proteins into the surrounding controls, achieving 96% wound closure after 7 days. As with the de- space [192]. In tendon cells from both healthy and tendons exposed to livery of the AMPs mentioned previously, the wounds experienced IL-1β (tendinopathic conditions), the addition of PRP to the culture re- significant increases in cell proliferation and angiogenesis. The duced the secretion of IL-6, IL-8 and monocyte chemoattractant protein treatment used in this study increased numbers of endothelial 1 (MCP-1). In addition, treatment with PRP increased the production of 104 A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125

VEGF-A and in normal and tendinopathic studies using proteins/peptides other than growth factors for enhancing cells. angiogenesis in wound healing. Portions of PRP called platelet lysates have been shown to promote angiogenesis and help direct cell migration during the wound healing 3. Gene and nucleic acid-based therapies for wound angiogenesis process [193]. A major function of PRP therapy is derived from its effect on mesenchymal stem cell (MSC) homing and migration. Platelet ly- 3.1. MicroRNAs in wound angiogenesis sates have been shown to promote significantly faster adhesion and mi- gration of MSCs in culture over the controls. When exposed to platelet Gene regulation plays a critical role in angiogenesis during wound lysate, MSCs increase their secretion of VEGF-A and PlGF, both impor- healing. MicroRNAs (miRNAs) are roughly 21-25 nucleotide long non- tant chemotactic growth factors for endothelial cells. While platelet ly- coding RNAs that play a powerful role in controlling gene expression sate also promotes endothelial migration, the addition of the secreted by binding to target mRNA and either suppressing mRNA synthesis or proteins from the MSCs in the local environment may even further ac- causing mRNA degradation [196–198]. Following inflammation, gene celerate the endothelial migration. Finally, platelet lysates have been expression is regulated to increase angiogenesis and remodeling of the shown to increase the levels of neovascularization, as shown by the in- woundedtissue[199]. MiRNAs have been shown to regulate many as- creased expression of PECAM-1 in platelet lysate coated scaffolds over pects of wound healing including cell proliferation and migration, colla- the uncoated scaffolds [193]. gen biosynthesis and network maturation, inhibition of Hemoglobin is another protein that researchers have used to im- neovascularization, and improved blood vessel growth [200–206]. prove ischemic wound healing [194,195]. Hemoglobin carries oxygen throughout the body’s circulatory system and oxygen plays a significant role in wound healing, by facilitating aerobic respiration, as well as 3.1.1. MicroRNAs that target angiogenesis in wound healing evolving into ROS. In a rabbit ear ischemic model, the delivery of IKOR With the power to increase or suppress gene expression, miRNAs 2048 (a chemically modified bovine hemoglobin) was shown to en- have become a target of interest in the field of wound healing. A handful hance wound repair and increase cellular survival. The local expression of miRNAs have been shown to specifically regulate angiogenesis within of Ki-67 (a marker of cellular proliferation), PECAM-1, VEGF and colla- wound healing, summarized in Table 3. These miRNAs are involved in gen were also increased upon administration of IKOR 2048 when com- increasing vessel formation and maturity, modulating migration, reduc- pared to the saline control. The delivery of 400 mg/kg of IKOR 2048 ing scar formation, and regulating expression of pro-angiogenic pro- significantly increased the oxygen tension in the tissue over untreated teins VEGF, GATA2, Ang-2 and PDGF-B, as well as the anti-angiogenic ischemic tissue, however, the treated tissue could not return it to nor- protein TSP-1 [200,203,205,207–209]. Overall, it is becoming increas- mal levels. Liposome encapsulated hemoglobin (LEH) delivery to gas- ingly clear that miRNAs play a powerful role in regulating angiogenesis trointestinal wounds in rats yielded an increase in burst pressure two in the context of wound healing. days after delivery, over the control groups, demonstrating increased High expression of various miRNAs has been correlated with de- healing. creased angiogenesis. By targeting these inhibitors of angiogenesis, sev- Blood derived factors present a promising area of research for scien- eral groups have been able increase angiogenesis in wounds. One tists and clinicians. While longer-term studies are needed to assess the method of inhibiting miRNAs is using anti-miRNA or “antagomirs”. safety of these treatments, the autologous nature of the treatments, Antagomirs are inhibitors of miRNAs that function by binding to specific the wide range of therapies available, and the auspicious pre-clinical miRNAs to prevent them from binding to their mRNA targets [210,211]. and clinical results indicate PRP treatments could provide a better alter- Patients with chronic wounds have been found to have a more than native to single growth factor therapies alone. Table 2 summarizes the two-fold greater expression of miR-92a compared to patients with healing skin wounds [212]. Likewise, overexpression of miR-92a has

Table 2 Proteins, peptides, and blood derived factors for enhancing wound healing and angiogenesis.

Treatment Function Exp. model Ref.

Insulin Accelerates wound healing, promotes keratinocyte migration, increases angiogenesis Mice/Rats [136–138] Erythropoietin Delivery increases the rates of neovascularization and increases levels of PECAM-1+ cells at the wound site Mice/Rats [140–143] hGH Increases re-epithelialization, increases angiogenesis, and increases infiltration of T-lymphocytes Diabetic [325,326] Rats Syndecan-4 Enhances the revascularization, wound closure, and angiogenesis in response to treatment with FGF-2 and PDGF-BB ob/ob [159] Mice/Rats Glypican-1 Increases revascularization in hind limb ischemia model and increases FGF-2 activity in endothelial cells ob/ob Mice [48] SDF-1 Promotes recruitment of endothelial progenitor cells and anti-inflammatory monocytes. Helps to accelerate the wound healing Mice [148,153] process and promotes bMSC migration to the wound site. Spidroin Increases proliferation of cells at the wound site to promote wound closure and angiogenesis Mice [156] Tβ4 Promotes faster wound closure rates, greater neovascularization at the wound site, and greater capillary density. Diabetic [157] Rats LL-37 Upregulates VEGF-A production, increases wound closure percentage, increases rates of re-epithelialization Mice [91–92,94] (peptide) CW49 Increases angiogenesis in vivo, HUVEC tube formation in vitro Diabetic [171] (peptide) Mice Tylotoin Accelerated wound closure (comparable to 100 μg/mL EGF dose), induces tube formation in vitro Mice [172] SR-0379 Decreased wound healing time, accelerated wound closure Rats [173] WRL3 Increased angiogenesis in burn model as shown by increase number of blood vessels per field of view Mice [174] Ap-57 Increase wound closure rates and promotes angiogenesis Mice [93,95] (peptide) VIP Increases wound closure percentages, cell proliferation and angiogenesis at the wound site Mice [180,181] PRP Increases the number of PECAM-1+ endothelial cells and the number of microvessels in the wound site. Increases the wound Rabbits [190,191,193] closure rate and MSC migration to the wound site Hemoglobin Increases cell survival, oxygen tension in the tissue, greater macrophage infiltration and tissue granulation. Rats [194,195] A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125 105

Table 3 Gene therapies for enhancing wound healing and angiogenesis.

Gene/miRNA Function Exp. model Ref. microRNA

miR-21 Supports wound contraction and collagen network maturation Wound Model in Mice [206] miR-23a Reduction of miR expression, increased MET expression and increased angiogenic potential In vitro [204] miR-26a Inhibiting miR-26a induced angiogenesis, and increased wound closure (by 53%) Wound Model in db/db mice [214] miR-27b Improves bone marrow derived angiogenic cells by suppressing TSP-1. Overexpression reduces Wound Model in db/db mice/Skin burn [200,201] directional migration of mMSCs. model in mice miR-29b Reduces collagen biosynthesis during wound healing via inhibition of HSP47 expression, also Wound and Burn Model in Mice [327] suppressing angiogenesis, leading to reduced scar formation miR-92a Overexpression blocks angiogenesis and antagomirs enhance blood vessel growth. Inhibition improves Hind limb Ischemia Mice and Wound [213,328] wound healing and angiogenesis. Model in db/db Mice miR-184 Reduces neovascularization, reduces FOG2, PDGF-B, and PPAP2B Matrigel plug assay in Mice [203] miR-191 Modulates cell migration and angiogenesis via ZO-1 regulation miR profiling in diabetic patients/in vitro [205] tubule formation assay miR-199a-5p Downregulation of miR increases Ets-1- MMP1 pathway expression and angiogenesis Wound model in WT and Ets-1 KO mice [329] miR-200b Inhibition of miR-200b increased GATA2 and VEGFR2 leading to improved wound angiogenesis and Excisional wound model in WT and db/db [208,215] closure mice

been shown to result in decreased angiogenesis, however anti-miR-92a samples from the wound bed, margin, and normal skin were collected antagomir treatment (8 kg/mg injection) enhanced blood vessel growth and total RNA was isolated for miRNA microarray analysis. MiR-31, in an in vivo hind limb ischemia model in mice by over 50% [213]. Inhi- miR-21, miR-203 were upregulated by more than two-fold and miR- bition of miR-92a also improved wound healing and angiogenesis in a 429 was down regulated by more than two-fold, confirmed by qPCR. full thickness excisional wound model, specifically through decreasing MiR-31 has been identified as a regulator of HIF-1α, miR-203 targets activity of subunit α5 (Itga5) and sirtuin-1 (Sirt1) [212,213]. p63 which is a regulates stem cell maintenance. Meanwhile, miR-203 Some antagomir treatments have been modified to be light inducible. and miR-429 regulate genes associated with tumors. Treatment with The addition of photolabile protecting groups prevents binding with 16 μg of miR-21 antagomirs dissolved in 100 μL of PBS injected locally target miRNA until activated by a light source, this light-inducible in the surrounding dermis reduced wound closure and collagen deposi- group has been labeled a “caged” antaogmir [212]. This light inducible tion. In addition, epithelial cell migration into the wound was reduced delivery method enables spatial control of the delivery of the gene ther- with miR-21 targeted antagomirs treatment [206]. In both excisional apy [212]. Treatment of diabetic (db/db) mice with an intradermal in- and burn wound models, expression of miR-29b was decreased by jection of 1 μg of light-inducible anti-miR-92a in 50 μL of PBS resulted over 50% compared to intact skin at day 7 post wounding [202]. How- in double the population of platelet endothelial cell adhesion molecule ever, treatment with oligonucleotide inhibitors of miR-29b increased (PECAM-1) positive cells in the granulation tissue compared to the HSP47 and COL1A1 gene expression in fibroblasts in vitro by more anti-miR-Control at 11 days post injury (Fig. 3A). Wound area closure than two-fold. Furthermore, miR-29b lentivirus treatment to increase for caged antimiR-92a with light was on par with non-caged antimiR- expression of miR-29b (2 x 107 TU/wound) applied either topically in 92a. Meanwhile, caged antimiR-92a was not effective at improving a diluted Matrigel or injected intradermally in an excisional wound wound healing, with closure rates similar to the cage antimiR-control model resulted a decrease in HSP47 and COL1A1 expression by over (Fig. 3B-G). Identifying miRNAs that regulate wound healing enables a 50% compared to the control virus treatment. Treatment with miR- clear path forward with potential antagomir treatment, as seen with 29b also reduced scar tissue formation and increased breaking strength miR92-a. In diabetic mouse wounds, levels of miR-26a have been of newly formed tissue [202]. found to be elevated compared to wild type (3.5-fold increase) [214]. MicroRNA profiling of patients with and without diabetes re- To target miR-26a, Locked Nucleic Acid (LNA) anti-miR-26a inhibitors vealed significantly lower levels of circulating miR-191 and miR- were used. Locked Nucleic Acid (LNA) is a high-affinity RNA analog, 200b in patients with type 2 diabetes. Interestingly, diabetic patients which “locks” to the target RNA. Intra dermal injection of 0.63 mg/kg with chronic wounds and peripheral arterial disease (PAD) had in- of LNA-anti-miR-26a inhibitors induced angiogenesis, accelerated creased levels miR-191 and miR-200b, like that of healthy patients wound closure by 53% and resulted in an approximately 50% increase [205]. Slightly higher concentrations of miR-191 were found at the in endothelial cells (PECAM-1+ cells) in full thickness excisional wound site compared to plasma levels in diabetic patients with wounds. PAD and chronic wounds. To understand the effects of miR-191, tu- It has been reported that temporary downregulation of miR-200b bule formation assays were performed. Tubule formation assays are during wound healing enables angiogenesis [208,215]. MicroR-200b is an in vitro assay to assess angiogenic potential, where more tubule thought to inhibit epithelial to mesenchymal transition and target an- formation is indicative of greater angiogenic potential. Treatment giogenic cytokines and transcription factors. Treatment of full thickness with miR-191 reduced in vitro tubule formation by over 50% mean- excisional wounds in mice with lentiviral particles expressing miR-200b while, treatment with anti-miR-191 increased tubule formation (at a titer of 2 x 107 cfu/mL, injected intradermally 1 mm from the compared to the scramble control. wound) induced slower wound healing and reduced blood flow, com- Another study found significantly reduced levels of miR-27b in pared to the control lentivirus treatment. Additionally, reduced num- mononuclear blood cells in patients with type 2 diabetes compared to bers of PECAM-1+ cells were found in the wound in miR-200b healthy patients [200]. To confirm the role of miR-27b in angiogenesis treatment groups compared to the control [208]. Treatment with anti- during wound healing, bone marrow-derived angiogenic cells (endo- miR-200b resulted in significantly increased microvessel density (mea- thelial progenitor cells selected via LDL uptake and lectin binding via sured via PECAM1+ immunostaining) compared to treatment with the Ulex europeus agglutinin-FITC binding assay) were co-delivered with vehicle control in a diabetic (db/db) mouse full thickness excisional miR-27 mimics (0.25 nmol in a 1% F-127 pluronic gel applied locally wound model [216]. to the surface of the wound). The miR-27 mimic delivery improved MiRNA profiling of wound healing in C57BL mice identified 54 wound perfusion, as measured by laser Doppler, in a full thickness exci- miRNAs that were increased or decreased more than two-fold com- sional wound model in db/db mice. Perfusion was improved by approx- pared to miRNA expression levels in normal skin [206]. Wound tissue imately 60% compared to scramble control treatment, to levels similar 106 A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125

Fig. 3. (A) Expression of miR-92a in healthy and diabetic mice, and in acute and chronic human wounds. (B) Percent wound closure over time in full thickness excisional wound model. (C) Photographs of wound area over time for each treatment group. (D) Distance between epithelial tips as a measure of re-epithelialization. (E) Amount of granulated tissue, measured as an area. (F) Distance between the ends of the panniculus carnosus. (G). Representative images of hematoxylin and eosin staining, indicating granulation tissue (gt), epithelium (e), and ends of the panniculus carnosus (arrow heads). Used with permission from [328]. A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125 107 to that of heterozygous non-diabetic littermate control mice after 8-10 fibrin scaffold to enable release of eNOS plasmids from the fibrin gel days [200]. followed by a secondary wave of release of Rab18 plasmids from the fi- Many new miRNA therapies are emerging with the potential to im- brin microspheres within the gel. The combination of Rab18-eNOS DNA prove angiogenesis in wounds to enhance healing. These therapies are delivery increased wound closure to approximately 90% after 14 days promising, in large part due to the targetable nature of miRNAs using whereas eNOS delivery alone had only reached approximately 70% antagomir treatments. However, effective delivery and stability in the wound closure [225]. This model is useful in delivering multiple treat- wound environment remain challenges for using miRNA-based thera- ments in sequence to maximize therapeutic effects. peutics in this area. Elastin-like-polypeptide (ELP) systems have also been used for the sequential delivery of genes such as eNOS and IL-10 [226]. IL-10 is an 3.2. Delivery methods anti-inflammatory cytokine that is present in ischemic tissue [227]. Like the fibrin microspheres in a fibrin gel, this system was designed 3.2.1. Free viral vector delivery to deliver IL-10 plasmids quickly followed by delivery of eNOS from Adeno-associated viruses (AAV) are non-enveloped viruses contain- ELP micro-spheres within the delivery platform. Co-delivery of IL-10 ing linear single stranded DNA. These viruses are characterized by low (10 μg) and eNOS (20 μg) and delivery of eNOS alone (20 μg) encoding immunogenicity and the need for a helper virus to replicate [217]. plasmids resulted in increased blood vessel density after two weeks in a AAVs are an attractive method for gene therapy particularly due to subcutaneous implant model compared to IL-10 plasmids. In a hind their safety, as they show little evidence of being pathogenic limb ischemia model in mice, perfusion almost doubled in eNOS and [217,218]. Recombinant adeno-associated viruses (rAAV) encoding IL-10/eNOS treatment groups in comparison to saline controls, ELP scaf- human vascular endothelial growth factor-165 (VEGF165) was deliv- fold alone, or IL-10 treatment [226]. ered in a db/db diabetic mouse wound model in which two full thick- Electrospun poly-L-lactide (PLA) and polycaprolactone (PCL) have ness longitudinal wounds were made on the dorsum of the mice. been used to create scaffolds for gene delivery of keratinocyte growth Recombinant AAV-VEGF165 treatment, delivered by intradermal injec- factor (KGF) [228]. It was found that PLA/PCL composite scaffolds pro- tion of ~1011 particles, resulted in a significant increase in angiogenesis moted the highest levels of cell adhesion to the scaffold, indicating after 7 and 14 days compared to the control delivery of rAAV-LacZ. Re- that the scaffold could support cell growth to enable angiogenesis and combinant AAV-VEGF165 increased angiogenesis, epidermal regenera- wound healing [228]. In a full thickness excisional mouse wound tion, and granulation tissue thickness at the wound site by over two- model, treatment with KGF plasmids (approximately 3.4 μg DNA/mg fold at both 7 and 14-day time points [219]. scaffold) in a PLA/PCL scaffold place on the wound resulted in a two- The are another family of regulators of angiogenesis fold increase in re-epithelization of excisional wounds significantly at and . Specifically, Ang-1 important for vessel stabiliza- day 5 after wounding compared to untreated wounds. Greater prolifer- tion, maturation, and remodeling [220]. Delivery of rAAV encoding ation, and improved epidermal thickness were seen after 7 days com- Ang-1 more than doubled granulation tissue thickness after both 7 pared to scaffolds with control plasmids [228]. and 14 days in a diabetic mouse incisional wound model, and nearly tri- Bilayer dermal equivalents have also been used for gene delivery in pled capillaries present in the wound samples in rAAV-Ang-1 treated full thickness wounds. Dermal equivalents are useful in providing a mice compared to the delivery of rAVV-LacZ after 14 days [209]. In template for the dermis to use as it heals. Bilayer dermal equivalents these studies, it was demonstrated that VEGFR-2 was more than dou- have been used to deliver VEGF-165 to full thickness burns in a porcine bled as a result of rAAV-Ang-1 treatment compared to rAAV-LacZ treat- wound model [229]. The porous collagen-chitosan/silicone bilayer ment, but VEGF mRNA expression did not differ between treatment membrane dermal equivalents (BDE) loaded with plasmid VEGF-165 groups [209]. Overall, it appears that AAV delivery of VEGF and (50 μg per mg scaffold) demonstrated sustained release for up to 4 angiopoietin are both effective treatments for improving angiogenesis weeks in vitro. In vivo results showed significantly more vessel forma- and wound healing in diabetic mouse models of wound healing. tion in the VEGF-BDE compared to blank BDEs or plasmid DNA alone, with the greatest improvements at 21 days being an 80% increase in ves- 3.2.2. Scaffolds and hydrogels employed in gene delivery sel density compared to the control [230]. As well as significantly more A broad range of scaffold and hydrogel delivery systems have been mature vessels at 7, 14, and 21 days post wounding. This demonstrates employed for delivering gene therapies targeting angiogenesis in the benefits of combinatorial therapies providing both a tissue scaffold wound healing, each delivery system with unique benefits for gene- and gene delivery in the wound. based therapies. The delivery systems include simple fibrin scaffolds, Modification of the HIF1-α gene by truncating the oxygen depen- gel-in-gel systems that enable sequential release, electrospun gels, mul- dent domain from the gene (HIF1-α-ΔODD) has provided a novel tilayer gels, and hydrogels functionalized with peptides to encourage gene to deliver for increasing angiogenesis [231,232]. Under normoxic nuclear translocation. condition, oxygen binds to HIF1-α and eventually results in degradation Nitric oxide (NO) is a regulator of endothelial cell growth, vasomotor of the HIF1-α protein. Removal of the ODD from HIF1-α prevents oxy- tone and angiogenesis [221]. Previously, it has been demonstrated that gen binding and subsequent degradation, thereby stabilizing the pro- there is a decrease in the production of nitric oxide (NO) in wounds of tein [233]. Once stabilized, HIF1-α can induce expression of VEGF patients with diabetes [222]. Endothelial nitric oxide synthase (eNOS) longer, with the goal of increasing angiogenesis. Delivery of HIF1-α can be delivered to increase NO production and enhance wound DNA with a deleted oxygen-sensitive degradation domain (HIF1-α- healing. Delivery of adenoviral eNOS (3 x 107 pfu) in a fibrin scaffold ΔODD) improved vessel maturity as measured by the number of vascu- (60 mg/mL) led to a sustained production of eNOS for two weeks com- lar structures displaying PECAM-1 and smooth muscle actin (SMA) pared to delivery of the free virus in a diabetic rabbit ear excisional [232]. The HIF1-α-ΔODD treatment resulted in the greatest number of wound model [223,224]. Additionally, the fibrin scaffold yielded a re- vessels displaying both PECAM-1 and α-SMA compared to VEGF-A pro- duction in the volume of inflammatory cells by approximately half tein treatment or the fibrin matrix alone in an excisional wound mouse from the 7-day time point to the 14-day time point [223]. model at the 7-day time point. In another study, the HIF1-α-ΔODD gene Rab18, a gene involved in trafficking and regulating secretion, had a was delivered with poly (L-lysine)-graft-poly (ethylene glycol) (PLL-g- 40-fold decrease in expression in wounded keratinocytes under high PEG) functionalized with one of two peptides, either a TG-peptide de- glucose conditions [225]. A hyperglycemic rabbit model was used to signed to sustain release of the DNA or a poly-arginine peptide designed test the co-delivery of Rab18 and eNOS plasmid DNA. Fibrin micro- to aid in nuclear translocation [231]. The poly-arginine modified PLL-g- spheres were encapsulated in a fibrin gel (60 mg/to deliver the plasmids PEG delivery of HIF1-α-ΔODD resulted in the greatest increase in VEGF- (10 μg of Rab18 plasmid DNA). This delivery system built on the use of a A, PECAM-1, and smooth muscle alpha-2 actin (ACTA2) in wounds of 108 A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125 normal mice and diabetic rats seven days after wounding, compared to inflammatory macrophages [252–254]. Macrophages play a significant all other treatment groups. role in the initiation and progression of wound healing, but an uncon- trolled inflammatory response in the long-term can be counterproduc- 3.2.3. Electroporation and sonoporation enhanced delivery tive for tissue repair. Further studies need to be performed to evaluate Electroporation uses electrical pulses to induce temporary pores in long-term dosage response of NO-releasing nanoparticles to determine the cell membrane, allowing the passage of DNA into the cell. long-term therapeutic efficacy. Sonoporation uses ultrasound to create cavitation bubbles, which in turn create pores in the cell membrane [234]. The major benefitofelec- 4.2. Adenosine triphosphate troporation and sonoporation is that they do not require virus particles to deliver DNA to cells [235,236]. Use of electroporation during delivery Most of the energy-requiring processes, including mitotic and bio- of plasmid TGF-β1 significantly increased wound closure three and five synthetic activities of cells during wound healing, either directly or in- days post wounding in a full thickness excisional wound model in dia- directly occur via hydrolysis of adenosine triphosphate (ATP). ATP is betic (db/db) mice [237]. Approximately five-fold more endothelial used throughout the wound healing process during various metabolic cells were observed in the granulation tissue at the center of the processes involving protein and lipid biosynthesis, growth factor pro- wound bed compared to TGF-β1 plasmid treatment without electropo- duction, signal transduction and mitosis. There are changes in enzy- ration, indicative of angiogenesis. matic and metabolic activity throughout the wound healing process. Minicircle DNA is a type of plasmid DNA in which the bacterial back- Reduced blood flow and decreased oxygen supply at wound site bone has been removed to reduce the potential for immunotoxic re- often results in ischemic and potentially hypoxic conditions which sponses [238,239]. Minicircle-VEGF165 DNA (20 μg) delivery via can disrupt the balance of energy production and utilization in cells. sonoporation improved wound healing compared to the null plasmid This imbalance can result in the depletion of high-energy phosphate control diabetic mouse treatment group in a full thickness excisional ATP that in turn can decrease the cellular energy supply, potentially wound model [240]. Wound closure was significantly improved at day hindering the healing process [255]. Although the delivery of ATP is six after wounding, almost to the same level as non-diabetic mice con- a potential strategy to enhance wound healing. Extracellular ATP re- trols. Blood perfusion and endothelial cell staining indicated that leased during cell damage and apoptosis has been shown to activate VEGF165 gene delivery via sonoporation increased capillary density damage-associated molecular patterns (DAMP) signaling eliciting an and blood perfusion in the wound site compared to null plasmid control inflammatory immune response [256]. Therefore it is important to uti- treated diabetic mice. Similar studies came to the same conclusion, ul- lize a vehicle to deliver soluble ATP to mediate an undesired immune trasound delivery of plasmid VEGF165 gene in a diabetic mouse exci- response. sional wound model yielded strong increases in blood perfusion in the Encapsulation of ATP in small unilamellar lipid vesicles (ATP-vesi- wound compared to EGF gene delivery and control diabetic mice (un- cles) accelerated the healing of full-thickness skin wound in nondia- treated) [241]. These studies demonstrate the potential for electropora- betic and diabetic rabbit models [257,258]. Early in vivo studies tion and sonoporation to boost gene delivery once targetable genes evaluating the treatment of full-thickness skin wounds with highly have been identified for angiogenesis in wound healing. fusogenic lipid vesicles containing magnesium–ATP showed that intra- cellular delivery of ATP significantly enhanced wound healing [257]. 4. Drug and drug-like compounds for enhancing wound Wang et al. created wounds on the ventral side of each ear using a angiogenesis stainless-steel punch. Ischemic ears were generated via artery ligation at the base of ear. Immunostaining for PECAM-1 in the non-ischemic 4.1. Strategies using delivery of small molecules to promote angiogenesis ear wound after six days showed that wounds treated with ATP- and rescue impaired wound healing in animal models vesicles contained significantly higher number of small capillaries compared to saline-treated group 14.3 ± 2.7 capillaries/field versus A dynamic angiogenic response is critical for wound healing. Nu- 5.5 ± 1.2 capillaries/field observed respectively. The delivery of ATP- merous investigators have focused on the development of innovative vesicles (10 mM) to ischemic ear wounds in rabbits initiated early an- approaches for therapeutic angiogenesis to aid in healing of chronic giogenesis in the wound bed to promote faster wound closure and wounds [242–245]. Recent advances in our understanding of chronic generation of better developed granular tissue and re- wound biology have led to the development of small-molecular thera- epithelialization compared to nontreated rabbits [258]. In addition to pies that promote angiogenic activity and can modulate the repair pro- the presence of capillary loops in the deeper layers of the wound cess with exciting potential for clinical application. Previously, small bed in the ATP-vesicles–treated group histological analysis confirmed molecules have been used to remedy diabetic impaired wound healing an increased presence of neutrophils, macrophages and lymphocytes in animal models [246–249]. Nitric oxide (NO) has been identified as in the ATP-vesicle treated group in comparison to saline-treated one molecule that plays a vital role in wound healing. NO levels increase group six days after injury. significantly following skin injury and then gradually decrease as A follow up study by Wang J. et al. performed in diabetic rabbits wound healing progresses [250,251]. Han et al. synthesized nitric observed similar results to the study published a year earlier [258]. oxide releasing nanoparticles (NO-np) by encapsulating in a hydrogel- Full thickness and ischemic ear wounds were created in alloxan- glass composite a mixture of tetramethyl orthosilicate, polyethylene induced diabetic mice. There was a significant difference in the glycol, chitosan, glucose and sodium nitrite in sodium phosphate buffer wound closure times between ATP-vesicle treated groups in both is- (0.5 mol/L) [250]. The nitrite within the matrix reduced to NO to allow chemic and non-ischemic ears. Like their first paper, treatment with for stable release of NO from nanoparticles. Han et al. found that topical ATP-vesicles increased closure rate of wound significantly exhibiting application of NO-releasing nanoparticles increased angiogenesis in wound closure as early as day 9 for ATP-vesicle treated mice com- mice excisional skin wound model utilizing implanted silicone ring. pared to day 12 for mice treated with saline only. When ischemic Generation of dense vascularization of wounded area was observed, wounds of diabetic mice were treated with ATP-vesicles, the average while non-treated wounds experienced minimal new blood vessel for- closure time was 15.3 days in comparison to 19.3 days for control mation by measuring expression of CD34. NO-releasing nanoparticles wounds [259]. Histological analysis after two days showed early an- also increased levels of TGF-β in wounds known to play an important giogenic activity the presence of a few small capillaries and numerous role in promoting angiogenesis during wound repair. In addition to an- inflammatory cells including neutrophils, lymphocytes, and macro- giogenesis, NO also mediates inflammatory response in wound healing phages were present in granulation tissue of ATP-vesicle treated regulating recruitment and infiltration of inflammatory cells like wounds. Unfortunately, Wang et al. did not measure the formation A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125 109 of mature vascular networks. After 17 days, saline-treated wounds studies suggest that statins angiogenic activity is mediated via the Akt/ showed incomplete re-epithelialization and less granulation tissue for- PI3K pathway to regulate proliferation of endothelial cells and capillary mation, while the ATP-vesicles treated group showed full granular tis- morphogenesis, and to modulate secretion of critical factors such as sue coverage and complete re-epithelialization of the wound VEGF-A in multiple cell types [261]. In an early study evaluating the indicating ATP-vesicles have the potential to initiate formation of a therapeutic potential of simvastatin to enhance wound healing, Bitto new vasculature network in injured tissue to restore blood circulation et al. used simvastatin to treat full-thickness wounds in female db/db in wound area critical to facilitate ECM production and epithelization, mice. Intraperitoneal injection of diabetic mice with simvastatin but this was not confirmed. [259]. These findings indicate that the in- (5 mg/kg) increased VEGF-A protein and gene expression in diabetic tracellular delivery of ATP enhanced wound healing in normal and di- mice with excisional wounds, following 3 and 6 days of treatment abetic wounds and offers a potential option for treating chronic non- [262]. Improved wound healing and increased angiogenesis were ob- healing wounds. These studies evaluated the angiogenic potential of served in simvastatin treated diabetic wounds by day 12. Immunostain- ATP-vesicles several days post injury. Future studies that investigate ing used to detect PECAM-1 in injured tissue sections revealed an the therapeutic potential of ATP-vesicles in wound healing should fur- increase in PECAM-1 positive staining and nearly double the microvas- ther evaluate the long-term effect of ATP-vesicle treatment on angio- culature density in simvastatin-treated group compared to the un- genesis in wound healing, and how presence or lack of a stable treated control group [262]. Additionally, daily treatment with vasculature alters progress of wound healing. Although the mecha- simvastatin also improved the tensile strength of damaged tissue and nisms by which intracellular ATP delivery enhances wound healing enhanced production of nitric oxide in wound, nitric oxide as previously are not fully understood, it is apparent that the availability of ATP is discussed is a molecule shown to play a significant role in wound an important aspect of the healing process. healing process [250,251]. A similar study conducted a few years later by Asai et al. observed 4.3. Statins that the wounds of db/db mice treated topically with simvastatin (50 μg/wound) exhibited more than 90% re-epithelization of injured tissue Statins are 3-hydroxy-3-methyl-glutaryl-coenzyme A (HMG-CoA) and accelerated wound closure with significantly smaller wound areas reductase inhibitors primarily used to treat hypercholesterolemia and observed by day 7 compared to control. Immunostaining for PECAM-1 atherosclerosis. Over the past decade, researchers have shown that, in showed a marked increase in neovascularization in treated wounds on addition to their lipid-lowering effect, statins possess pro-angiogenic day 14 (Fig. 4A, B). In addition to enhanced wound vascularity, immu- properties that can potentially protect against ischemic injury and stim- nostaining for lymphatic endothelial cell specific marker lymphatic en- ulate angiogenesis [260]. Although still under investigation, early dothelial hyaluronan (HA) receptor-1 (LYVE-1) displayed the presence

Fig. 4. Effects of simvastatin on vascularity and lymphangiogenesis in full-thickness excision skin wound of db/db mice. (A) Neovascularization 14 days post-wounding at the wound margin of simvastatin-treated or vehicle-treated db/db mice. (B) Quantification of percent of vascularity. (C) Lymphangiogenesis 14 days post-wounding at the wound margin of simvastatin-treated or vehicle-treated db/db mice. (D) Quantification of percent of lymphatic vascularity. Green corresponds to LYVE-1-positive, newly formed lymphatic vessels. Blue fluorescence indicates DAPI-labeled nuclei. Scale bar = 100 μm. Used with permission from [246]. 110 A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125 of new lymphatic vessel at the margins of the wound (Fig. 4C, D) [246]. controlling hypoxia-mediated events in wound healing [247,269]. Dur- Presence of new lymphatic vessels suggests recovery of the ing wound repair, the active subunit HIF-1α regulates various processes lymphangiogenic function, which is important for maintaining tissue including cell metabolism, proliferation, survival, and angiogenesis homeostasis and macrophage infiltration during an immune response targeting important pro-angiogenic genes such as VEGF-A and stromal [263]. Treatment with simvastatin resulted in an increased population cell-derived factor 1-α (SDF-1α)[247]. SDF-1α also known as C-X-C of macrophages M2 (alternatively activate) macrophages expressing motif chemokine 12 (CXCL12), has been shown to be beneficial in pro- IL-13 with a significantly smaller population of TNF-α expressing M1 moting wound healing in in vivo models.[155] The chemokine CXL12 macrophages indicating that simivastatin treatment did not elicit an un- binds receptors on surface of immune cells to regulate inflammatory re- desired inflammatory response that could potentially hinder wound sponse during healing process. Lactobacillus reuteri bacteria (L. reuteri) healing. Simvastatin treated group also exhibited increased protein ex- transformed with a plasmid encoding for the chemokine CXCL12 to pro- pression of VEGF-C in simvastatin treated wounds. Gene expression for duce prolong overexpression of CXCL12 by sustaining a pH regulated PDGF-B, eNOS and FGF-2 were also significantly upregulated by simva- hypoxic wound microenvironment were shown to be beneficial to statin treatment [246]. wound healing when topically administered to full-thickness wounds More recently, improved drug delivery systems have been utilized to in mice with peripheral ischemic following hind-limb ischemia proce- enhance the angiogenic and wound healing potential of statins for dure and hyperglycemia.[155] Improvements in wound healing and treating chronic wounds. Lyophilized wafers composed of sodium blood perfusion were observed in both hyperglycemic and ischemic carboxymethyl cellulose and methyl cellulose have been developed as mice. Therefore, downstream effectors of HIF-1 particularly CXCL12 or potential wound dressings for administering simvastatin and other SDF-1α are promising targets for wound healing therapeutics. small molecules to chronic wound sites [263]. Moisture retentive dress- Deferoxamine (DFO) is an iron chelator, which has been used to ings are desired for treating chronic wounds to absorb wound exudate mimic oxygen deprivation in cells and induce the accumulation of and manage the moisture balance [264]. Lyophilized wafers are a viable HIF-1α. Hou et al. explored the effect of DFO treatment on neovascular- candidate for wound dressings that require a moist environment such ization in diabetic wounds, and the expression of HIF-1α and down- as mucosal wounds and venous ulcers [265]. When applied to wound stream genes. Full-thickness excision cutaneous wounds were sites, they absorb wound exudate and form a gel to maintain a moist en- generated on dorsal skin of rats with streptozotocin-induced diabetes vironment [263,265]. Hydrogels have also been developed that possess and treated with the vehicle control group (PBS), DFO (10 mg/mL) or desired mechanical properties that maintain optimum moisture in the VEGF-A (0.5 nM) [247]. The DFO-treated group exhibited accelerated wound area and aid in wound healing process. An ideal wound dressing wound closure at 7, 10 and 14 days compared to the vehicle and would exhibit slower drug release from the medicated dressing to help VEGF-A treated groups [247]. Histological analysis showed that DFO prolong the action and release of the active drug at wound site. A recent treatment stimulated formation of microvessels around the newly study by Yasasvini et al. evaluated the in vivo wound healing efficacy of formed granulation tissue seven days after wounding. Vessel density polyvinyl alcohol (PVA) hydrogels loaded with simvastatin encapsu- was assessed via vWF immunostaining at 7 and 14 days. The number lated chitosan microparticles. Simvastatin loaded chitosan microparti- of vessels per field was approximately four times greater in DFO treated cles were generated via modified ionic gelation method and loaded wounds compared to the vehicle only group, and slightly higher than hydrogels were prepared by chemical cross linking method.[266]Cuta- the VEGF-A treated groups with 20 and 15 vessels/field respectively. neous wounds were created in rats followed by treatment with varying Western blot analysis of tissue lysates showed that HIF-1α was associ- concentration of simvastatin-chitosan microparticle loaded PVA ated with DFO treatment enhancing neovascularization. Protein expres- hydrogels (2.5, 5 and 10 mg). Treatment with low dose simvastatin sion of HIF-α was significantly elevated in DFO treated wounds and hydrogels (2.5 mg/patch) every 7 days for 21 days displayed increased VEFG protein expression was also found to be higher in the DFO- rate of wound closure compared to control non-treated wounds and treated group compared to vehicle control group. Protein expression medium and high dose treated wounds (5 mg and 10 mg respectively) levels of SDF-1α were elevated at day 7 compared to both VEGF-A and [266]. In rat wounds treated with low dose hydrogels, histological anal- vehicle treated groups. These results suggest the potential of DFO to en- ysis revealed granulation tissue with complete epithelialization 21 days hance neovascularization in wound healing through increasing HIF-1α after wounding. Recent studies have shown that utilizing drug delivery activity. systems to administer statins have increased drug efficacy as a thera- A later study by Ram et al. found that treating wounds of diabetic in- peutic in wound healing. Unfortunately, there is still very little research duced rats with DFO ointment (0.1%) also enhanced wound healing and that explores how these systems enhance statins angiogenic potential angiogenesis. Wounds treated with DFO experienced greater wound in vivo. Previous studies exploring the angiogenic potential of free closure from day 7 to 19 compared to control group. Real time PCR statins have shown that topical administration of statins promotes vas- and western blot analysis showed that gene and protein expression of cular formation in wound injury. The field would further benefitfrom HIF-1α, VEGF-A, SDF-1α,TGF-β1 and IL-10 were significantly upregu- exploring how target drug delivery systems enhance this effect. Addi- lated in wounds treated with DFO from days 3 to 14 [270]. In contrast, tionally, although novel drug delivery systems offer a strategy for sus- the relative mRNA expressions of pro-inflammatory cytokines TNF-α, tainable controlled drug release to facilitate a prolonged therapeutic MMP-9 and IL-1β were significantly downregulated at days 7 and 14 effect of drugs and aid in long-term wound recovery, further character- post-wounding. Additionally, DFO treatment was observed to decrease ization of the in vitro and in vivo toxicity and efficacy of simvastatin- protein levels of tumor necrosis factor alpha (TNF-α) post-wounding loaded delivery systems is needed. from day 3 onward and increased protein levels of IL-10 compared to controls at day 3 and onward. Indicating that treatment with DFO 4.4. Deferoxamine does not induce an early inflammatory response in vivo. DFO also in- duced angiogenesis in newly formed granulation tissue [270]. Although In chronic wounds the healing process is often impaired and many density of microvessels was not quantified, an overall larger number of growth factors like VEGF, which are essential to angiogenesis and the small blood vessels were observed in regenerated tissue of DFO-treated wound repair process are downregulated [267,268]. As part of the group throughout the 19-day study compared to control group. wound repair mechanism, the wound microenvironment can become Researchers have tried to enhance the therapeutic potential of DFO hypoxic due to injury of the microvasculature. This hypoxia can induce using transdermal delivery systems that can obtain deeper penetration cytokine production and cell recruitment to accelerate wound closure into the wound bed. Duscher et al. developed a transdermal drug deliv- [247,269]. Hypoxia-inducible factor-1 (HIF-1) is involved in regulating ery system that used reverse micelle encapsulation of DFO by nonionic the response of cells to altered oxygen levels and plays a key role in surfactants (polysorbate 80 and sorbitan monolaurete) within a A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125 111

Fig. 5. Deferoxamine-treated diabetic ulcers exhibit increased neovascularization and improved dermal thickness. (A) Immunohistochemical staining for newly formed capillaries via endothelial cell marker PECAM-1 (red). Scale bar = 10 μm. (B) Quantification of PECAM-1+ pixels per high-power field (HPF). (C) Picrosirius red staining used to assess dermal thickness of completely healed diabetic mice ulcers. Scale bar = 10 μm. (D) Quantification of picrosirius red-positive pixels per HPF. Used with permission [271]. polyvinyprrolidone (PVP) biodegradable matrix and disperse in ethyl (Fig. 6C). Consistent with the reduction in scarring, AS-IV treatment cellulose for slow release [271]. Pressure ulcers were induced in db/db stimulated angiogenesis at wound site. A few blood vessels could be ob- mice by placing two ceramic magnets for three separate ischemia/re- served in healed tissue of AS-IV treated group 30 days after wounding perfusion cycles. Each cycle consisted of placement of magnets for 3 or (Fig. 6A-B) [272]. 6 hours followed by release or reperfusion for 3 or 6 hours. Treatment To reduce the dosing frequency and enhance therapeutic efficacy, of pressure ulcers with transdermal DFO (1%) improved healing and pharmaceutical carriers for the topical application of AS-IV have been stimulated complete wound closure by day 27 compared to day 39 for investigated. One study examined the use of solid lipid nanoparticle- control group. Histological analysis showed that localized treatment enriched hydrogel (SLN-gel) prepared by a solvent evaporation method with DFO transdermal delivery improved the dermal thickness of to encapsulate AS-IV (0.5% AS-IV/gel) and locally delivered it to full- healed diabetic ulcer and promoted neovascularization within the thickness skin excision wounds in rats [273]. They found that treating wound bed (Fig. 5A-D). Immunostaining for PECAM-1 showed that wounds with solid lipid nanoparticle-enriched hydrogel loaded with treatment with transdermal DFO had more than twice the number of AS-IV (AS-based SLN-gel) dramatically enhanced wound healing capa- PECAM-1 positive capillaries. Sustained local delivery of DFO treatment bilities compared to a blank SLN-gel, and free AS-IV solution treated was also found to increase early VEGF protein expression in diabetic ul- groups at days 4 and 12 after wounding. AS-IV solution and AS-based cers at 24 and 48 hours. Pretreatment for 48 h with transdermal deliv- SLN-gel both stimulated angiogenesis at the wound site. Histological ered DFO prior to ulcer induction visibly reduced tissue necrosis and analysis of the wound beds showed a higher density of newly formed prevented ulcer formation later. Dysregulation of the HIF-1α/VEGF sig- blood vessels in both AS-IV solution and AS-based SLN-gel, while no naling can be caused by excessive ROS production a central problem in blood vessels were observed in control and only a few microvessels in diabetic wound healing Levels of ROS were negatively affected by the SLN-gel carrier only treated group [273]. pretreating with DFO. Dihydroethidium (DHE) is a superoxide indicator In addition, AS-IV has been used to functionalize wound dressings to used to measure in situ generation of ROS. DHE immunofluorescent promote healing and prevent scar formation in burn wounds. Shan et al. staining showed that ROS levels decreased dramatically in DFO developed silk fibroin/gelatin electrospun nanofibrous dressings loaded pretreated db/db mice. Thus, DFO is a promising therapy for rescuing with AS-IV (0.5 mg). A homogeneous solution of a suitable blending impaired wound healing in diabetic patients and novel delivery systems ratio of SF/gelatin and AS-IV was flowed through a needle at a fixed offer a potential strategy for improving treatment efficacy. rate of 0.01 mL/min. under a high applied voltage, 15kV, and electrospun to produce nanofibrous dressing. Treatment with AS-IV 4.5. Natural compounds loaded dressing significantly accelerated wound healing compared to blank control (normal saline) groups at 7 and 15 days post-wounding In addition to tissue healing and regeneration, another goal of in a partial-thickness burn wound model in rats [274]. Minimal scarring wound therapy is to promote adequate healing while minimizing scar was observed in repaired tissue of both AS-IV loaded nanofibrous dress- complications such as imbalanced collagen synthesis [272]. ing and AS-IV solution treated wounds. In addition, immunostaining for Astragaloside IV (AS-IV) one of the main active ingredients in Astragali PECAM-1 showed increased blood vessels in both treatment groups. Radix derived from the root of Astragalus membranaceus has long been Newly formed vessels were observed in the newly formed granulation exploited in Traditional Chinese Medicine for its healing and anti- tissue 31 days post-wounding in both AS-IV treatment groups. Treat- scarring effects [272,273]. Chen et al. observed that AS-IV (0.5%) locally ment with AS-IV enhanced the number and dimensions of blood vessels delivered at the wound site of full-thickness skin excision wounds in compared to no vasculature formed in blank control and very few micro rats stimulated quicker recovery of wounds with minimal scarring at vessels in blank nanofibrous dressing [274]. AS-IV is a promising candi- day 30 compared to blank control wounds. A lower ratio of type I/III date for treating impaired wound healing. Previous studies have shown was observed in the AS-IV group compared to that of the blank control its ability to promote wound re-epithelization, angiogenesis and regu- group demonstrating treatment was able to regulate collagen synthesis late extracellular matrix remodeling in chronic and burn wounds. Al- and reduce scar formation during the wound remodeling process though a poorly water-soluble drug, it has been successfully 112 A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125

Fig. 6. Anti-scar effects of astragaloside IV on rat full-thickness skin excision wounds 30 days after wounding. (A) Masson’s trichrome staining of granulation tissue to assess collagen synthesis and blood vessels formation. Collagen and blood vessels identified by red and green arrows, respectively. (B) Effect on angiogenesis assessed by presence of blood vessels in granulation tissue indicated by black arrows. (C) Sirius red staining to evaluate collagen composition in healed tissue. Type I collagen indicated by white arrow, and type III collagen indicated by pink arrows. Blank control refers to the no treatment group. Used with permission from [272]. incorporated into novel biocompatible carriers that allow for sustained macrophages present in wound at 1, 5 and 7 days post treatment com- drug release to induce increased tissue regeneration and inhibit scar for- pared to control. Examination of cytokine levels in the exudates from mation during wound healing. asiaticoside-treated burn wounds showed increased levels of IL-1β at Herbal remedies have been helpful in the treatment of dermatolog- multiple time points compared to control vehicle treated group [278]. ical disorders such as skin lesions, eczema, burns and hypertrophic Topical application of asiaticoside also increased level of monocyte [275–277]. One example is the plant, Centella asiatica that contains var- chemoattractant protein-1 (MCP-1) in exudates of burn wound- ious active compounds such as centelloids, asiaticoside, madecasosside, treated mice compared to control group [278]. Histological analysis of asiatic acid and madecassic acid that have been found to be beneficial in regenerating skin of burn wound-treated mice showed increased num- wound healing [275]. Ointments containing varying doses (0.1 to 1%, w/ bers of proliferating cells, as assessed by Ki-67 staining, and increased in w) of these compounds or a combination have previously been shown number of VEGF expressing cells in the healing wound at day 9 [278]. to enhance wound repair [275,278–280]. Asiaticoside derived from Several other studies demonstrated that low doses of asiaticoside stim- Centella asiatica possesses diverse pharmacological effects including an- ulate angiogenesis by enhancing keratinocyte production of MCP-1 that giogenic promoting activity and wound healing enhancing capabilities. in turn increases VEGF-A levels in chronic wounds [278,279]. Kimura et al. observed that topical treatment of burns in mice with The poor solubility of asiaticoside in aqueous medium is a major asiaticoside at lower doses of 10−8 to 10−12 (w/w) enhanced burn limiting factor in the drug’s therapeutic efficacy [279–281]. To im- wound healing. Treating with ointment for 20 days containing various prove the pharmacological effect of free asiaticoside, carriers have doses of asiaticoside significantly reduced burn wound area on days 6 been developed to increase wound healing potential. Phaechamud to 18 compared to vehicle control (petroleum jelly only) with no signif- et al. performed the chick chorioallanotoic membrane assay (CAM) icant difference observed between three concentrations (10−8,10−10 to evaluate the angiogenic effect of chitosan-aluminum monostearate and 10−12 w/w) [278]. Kimura’s group also treated the burn surface composite sponge dressings loaded with varying concentrations of with polyethylene filter pellets containing three different concentra- asiaticoside. Homogenous mixture of chitosan solution and tions of asiaticoside 10 pg, 1 ng, and 100 ng. Asiaticoside (10 pg and 1 asiaticoside was fabricated via lyophilization technique, frozen, freeze ng/filter pellet) decreased polymorphonuclear leukocytes infiltration dried, and treated with dehydrothermal treatment to stabilize struc- at wound site at 1 and 3 days, while treatment with all three doses ture of lyophilized sponge dressing. Phaechamud et al. showed that (10 pg, 1 ng, and 100 ng/filter pellet) increased the number of asiaticoside exhibits dose-dependent angiogenic activity in vitro A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125 113

[279]. The average angiogenesis score increased with increasing con- HA-Mediated Motility (RHAMM), hyaluronic acid receptor for endocy- centration of asiaticoside. The average angiogenesis score was calcu- tosis (HARE), lymphatic vessel endothelial hyaluronic acid receptor 1 lated as the ratio of total number of embryos treated to evaluable (LYVE1), layilin, toll-like receptor 2 (TLR2) and toll-like receptor 4 embryos left after a 12-day study. Embryos treated with varying con- (TLR4) to influence receptor activation and their downstream signaling. centration of asiaticoside 40 μg, 80 μg, 160 μg, and 320 μg generated CD44 is a cell-surface receptor for hyaluronan implicated in the regula- the highest overall angiogenesis scores 1.25, 1.25, 1.83, and 2.12 re- tion of endothelial cell functions, and tumor angiogenesis [295,296]. Al- spectively. In comparison to VEGF positive control which at the fol- though the role of CD44 in angiogenesis is still being investigated, CD44 lowing concentrations 100ng, 200ng, and 300ng scored 0.43, 0.17, has been shown to possess anti-angiogenic properties impairing tube and 0.2 respectively. formation in vitro and in vivo [296,297]. CD44 mediates HA-dependent In another study, full-thickness skin wounds in rats were treated cell signaling involved in the regulation of endothelial cell proliferation, with asiaticoside-loaded porous polyvinyl alcohol (PVA) microspheres migration, adhesion, and tube formation required for angiogenesis. The [280]. The asiaticoside-microspheres enhanced wound healing com- interaction between HA and CD44 receptor is, in part, dictated by the pared to untreated wounds or those treated with free asiaticoside solu- size of the HA molecule that can influence CD44 clustering and activa- tion. Asiaticoside microspheres enhanced wound healing accelerating tion. Oligomeric HA can competitively bind to CD44 receptor in place closer rates compared to free drug significantly at day 7 and day 14 of anti-angiogenic high molecular weight HA to act as an antagonist to and minimized scar formation. In the rats treated with asiaticoside mi- CD44, triggering intracellular signaling to promote endothelial cell pro- crospheres, immunohistochemical staining for PECAM-1 showed a liferation, and migration to stimulate neovascularization in vitro higher density of microvessels in comparison to the other treatment [287,291]. groups. Overall, treatment with asiaticoside reduced scar complications, Previous studies have shown that topical delivery of exogenous stimulated re-epithelialization and collagen deposition, and promoted HA oligomers to injured tissue promotes wound repair in vivo wound angiogenesis to improve wound healing. Although free [249,292]. Gao et al. topically treated murine excisional dermal asiaticoside has been previously shown to possess properties that en- wounds with HA oligomers (50 μg/per wound) encapsulated in a hance healing and reduce scar formation, its lack of solubility is a slow-releasing gel of ethylene-vinyl acetate copolymer. Treatment major limitation and new carriers for its delivery would be beneficial with HA oligomers accelerated wound healing, with the treated for improving its bioavailability. wounds exhibiting similar closure rates and recovery as those Materials derived from natural processes have also been shown to treated with VEGF-A (10 μg per wound). The HA oligomer treated have significant pro-angiogenic activity in wound healing. Dextran wounds had increased numbers of blood vessels. Compared to the hydrogel-based dressings were highly proangiogenic in burn injury PBS and HMW-HA treated mice, there was a significantly larger in mice and pigs [282,283]. In comparison to control wounds, there number of blood vessels in the wound bed. The group treated with was increased infiltration in wounds treated with dextran HA oligomer encapsulated gel by day 6 had approximately 4 times hydrogels [282]. In addition, there was increased infiltration of endo- the number of blood vessels per field area of those wounds treated thelial cells with treatment with the dextran hydrogel dressings. Bio- with PBS and HMW-HA. By day 8 post wounding HA oligomer gels active glass microfibers doped with cupric oxide also increased contained more PECAM-1 positive blood vessels although this num- angiogenesis and wound healing [284]. These microfibers degrade ber was less significant than at day 6. Additionally, treatment with to hydroxyapatite under physiological conditions in seven days. In HA oligomers induced the formation of new lymphatic endothelium. a rat wound model, these fibers increased wound closure and vessel Immunostaining for LYVE-1, a marker for lymphatic endothelial area during healing. cells, was upregulated and dispersed throughout vessels in the wound bed of mice treated with HA oligomers, VEGF and HMW- 4.6. Hyaluronan oligosaccharides HA [292]. Overall density of lymphatic vessels in HA oligomer- treated mice was significantly greater than PBS-treated mice on Hyaluronic acid (HA), also known as hyaluronan, is a glycosamino- day 4 and 8. Although the lymphatic vessel density was less com- glycan (GAG) that is a long unbranched polymer comprised of repeating pared to high molecular weight HA and VEGF-A, suggesting that D-glucuronic acid and N-acetyl-D-glucosamine disaccharide units, HA oligomers are superior in stimulating angiogenesis over bound through alternating β-1, 4 and β-1, 3 glycosidic bonds. HA is a lymphangiogenesis. Administration of HA oligomers was found to functional component of ECM that plays a critical affect the expression of numerous factors involved in wound repair role in cell signaling, tissue development and maintenance of tissue including endothelial cell proliferation, cell migration and collagen health and homeostasis [285,286]. Unlike many other GAGs, HA is not synthesis. Treatment with HA oligomers upregulated gene expres- sulfated or covalently attached to core protein as a proteoglycan [287]. sion of endothelial eNOS, an important atheroprotective molecule, Previous research has implicated hyaluronan as a potential therapeutic and cell adhesion molecules, E-selectin and integrin β3atearly capable of both stimulating and inhibiting angiogenesis depending on time points compared to control [292]. In addition to inducing an- its form and concentration [287,288]. Physiological responses to giogenesis during wound repair of dermal injury, oligomeric HA hyaluronic acid is highly influenced by HA molecular weight, which is treatment effects the mRNA expression of key endothelial mediators classified into low (≤5 kDa), intermediate (60-800 kDa) and high and regulators of collagen deposition. (N800 kDa) molecular weight groups [289]. Physiological concentra- A recent study also supports the use of HA oligomers for treating di- tions of native long-chain, high molecular weight HA have been abetic wounds. Ointment prepared by emulsification containing HA shown to induce anti-inflammatory and anti-angiogenic activity oligomers (0.15%), petroleum jelly (2.4 g), stearic acid (2.4 g), glycerol in vivo and in vitro [249,288,290,291]. In particular, HA oligomers monostearate (1.6 g) and liquid paraffin wax (2.5 g) was topically ad- consisting of 2-10 disaccharides are highly bioactive and capable of ministered to wounds created in rats with streptozotocin-induced dia- stimulating neovascularization in animal wound models [249,292]. Al- betes [249]. Treatment with HA oligomers containing ointment though the size-dependent effects of HA on biological processes are significantly decreased wound size and induced formation of granula- not completely understood, the size of HA can have a considerable influ- tion tissue and re-epithelialization of wound surface by day 14 after ence on receptor activation and its downstream signaling. In an injured wounding [249]. In addition to enhancing wound healing, HA oligomer or disease state, aberrant conditions can cause the degradation of high treatment improved blood flow in diabetic wounds as measured by molecular weight by hyaluronidases and ROS into low molecular weight laser doppler imaging. Histological analysis of wound tissue 14 days HA which further depolymerized to form oligomeric HA [293,294]. post-wounding showed increased numbers of capillaries in the HA olig- Byproducts can then bind several cell surface receptors such as CD44, omer groups in comparison to control wounds [249]. Treatment with 114 A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125

Table 4 Small molecules for enhancing wound healing and angiogenesis in mice.

Treatment Results Mechanism of action Exp. model Ref.

Simvastatin ↑ VEGF mRNA Inhibition of HMG-CoA reductase db/db Mice [262] ↑ Endothelial Cells Pleiotropic Effects ↑ NO products ↑ Collagen synthesis Simvastatin ↑ Endothelial Cells Inhibition of HMG-CoA reductase db/db Mice [246] ↑ LYVE-1 Pleiotropic Effects ↑ Macrophage infiltration ↑ VEGF-C ↑ PDGF- β, eNOS, FGF-2 mRNA Deferoxamine ↑ VEGF protein Iron chelator db/db Mice [271] ↑ Endothelial Cells Increases HIF-1α transactivation Upregulates angiogenic factors ↓ Skin Necrosis/Apoptosis ↓ ROS Prevented Ulcer Formation Asiaticoside ↑ Macrophage Infiltration Inhibition of TGF-β receptors Mice [278] ↑ VEGF, MCP-1, and IL-1β Increased NO production ↑ Cell Proliferation Increased TNF-α production HA oligomers ↑ Endothelial Cells Binds to CD44 receptor promoting pro-mitotic effects on Mice [292] ↑ LYVE-1 endothelial cells and angiogenesis ↑ Fibroblast proliferation ↑ eNOS/procollagen mRNA ↓ MMP-9/MMP-13 mRNA

HA oligomer containing ointment was found to not affect total Src and paracrine effects, immune response modulation, and their ability to dif- ERK1/2 expression in wound tissue, but levels of phospho-Src and ferentiate into vascular lineages. phospho-ERK1/2 increased significantly [249]. Additionally, oligomeric The mode of delivery of cells to the wound can have profound effects HA treatment increased expression of TGF-β1 protein. on the cell state and regenerative activity [298]. Multiple strategies have Overall, hyaluronan has proven to be a versatile bioactive molecule been tested to overcome the general lack of engraftment and cell viabil- playing a significant role in the initiation and progression of biological ity associated with cellular therapeutics [299]. Many researchers have processes and pathological conditions. Shorter HA fragments like oligo- utilized porous scaffolds constructed out of various polymers and mers are crucial mediators of intracellular signaling cascades triggering hydrogels. Primarily, these scaffolds function to improve cellular en- activation of surface receptors such as CD44 to regulate inflammatory graftment so that they can elicit a prolonged effect at the site of injury. and angiogenic responses [287]. Topical administration of oligomeric Ideally, scaffold materials should be non-immunogenic, non-cytotoxic HA is a promising therapeutic strategy to promote wound healing and and facilitate a phenotype in the delivered cells that supports their re- angiogenesis in animal models of chronic and diabetic wounds generative properties. For wound healing applications, resorbable scaf- [249,292]. Evidence suggests treatment with HA oligomers is beneficial folds are often used so that endogenous tissues can grow to replace the in promoting neovascularization in animal wound models, but addi- scaffold area. Furthermore, scaffolds benefit from having a level of po- tional studies are needed to further profile the adverse effects of rosity that can allow for cellular infiltration of their network and migra- prolonged HA oligosaccharide treatment including potential unwanted tion toward the site of injury. To promote this outward migration, cell activation and prolonged inflammation. Tables 4, 5 and 6 summa- integrin-binding/cell adhesion sites can be added to the material sur- rize the results of studies using small molecular therapies to enhance face. While each of these characteristics can help with cellular viability wound healing and angiogenesis. and integration into host tissue, some cellular therapies have found suc- cess without the use of a scaffold. The following section summarizes and highlights several differing approaches for promoting angiogenesis and 5. Stem cell based therapies for wound angiogenesis wound healing through therapeutic cellular delivery.

5.1. Challenges and strategies for therapeutic cell delivery 5.2. Adipose derived stem cells

Stem cell therapies are an emerging and promising approach for Adipose-derived stem cells (ASCs) are a promising therapeutic cell treating diseases not well addressed by conventional therapies. Several type for wound healing as they are multipotent and can achieve a vascu- stem cell types have been explored to facilitate angiogenesis and wound lar phenotype, however their success has been hindered due to poor healing. Adult stem cells including adipose-derived stem cells (ASCs), survivability at the wound microenvironment [299]. Numerous studies bone marrow-derived mesenchymal stem cells (MSCs) and induced have been conducted to optimize ASC delivery and achieve a robust pluripotent stem cells (iPSCs) have been explored for their potential therapeutic effect in various wound healing models (Table 7). Zamora

Table 5 Small molecules for enhancing wound healing and angiogenesis in rabbits.

Exp. model Treatment Results Mechanism of action Ref.

Rabbit Mg-ATP delivered in ↑ Endothelial Cells Regulates production of pro- and anti-angiogenic factors that activate [258] unilamellar lipid vesicles ↑ Inflammatory Cells various receptors such as A2B receptors ↑ Re-epithelialization Diabetic Rabbit Mg-ATP delivered in ↑ Granulation Tissue Regulates production of pro- and anti-angiogenic factors that activate [259] unilamellar lipid vesicles ↑ Re-epithelialization various receptors such as A2B receptors ↑ Inflammatory cells A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125 115

Table 6 Small molecules for enhancing wound healing and angiogenesis in rats.

Exp. Treatment Results Mechanism of action Ref model

Rat Simvastatin loaded ↓ Wound size Inhibition of HMG-CoA reductase [266] chitosan microparticles in Complete epithelialization (d21) Pleiotropic Effects PVA hydrogel Rat Astragaloside IV ↓ Scar Formation Alters activity of VEGF receptors regulating pathways [272] Topically applied ↑ Skin Wound Tensile Strength involving PKB/PI3K ↑ Microvessel Density ↓ Ratio of Type I/III collagen Rat Astragaloside IV ↑ Blood Vessels Alters activity of VEGF receptors regulating pathways [273] Solid lipid nanoparticle in hydrogel ↑ type I collagen deposition involving PKB/PI3K Rat Astragaloside IV in ↓ Scar Formation Alters activity of VEGF receptors regulating pathways [274] Silk fibroin/gelatin electrospun nanofibrous dressing ↑ VEGF involving PKB/PI3K ↑ numbers of blood vessels Type I collagen deposition Rat Asiaticoside in ↑ Blood Vessels Inhibition of TGF-β receptors [280] PVA/ethyl cellulose microspheres Minimal Scarring Increases NO and TNF-α Organized Collagen I Fibers Diabetic Rat +/- DFO ↑ Capillary Density Iron chelator [247] +/- VEGF ↑ HIF-1α, SDF-1α and VEGF Increases HIF-1α transactivation Upregulates Intraperitoneal Injection angiogenic factors Diabetic Rat DFO Ointment ↑ HIF-1α, VEGF, SDF-1α, TGF-β1 and IL-10 Iron chelator [270] ↓ TNF-α, MMP-9, and IL-1β Increases HIF-1α transactivation Upregulates angiogenic factors Diabetic Rat HA Oligomers in ↓ Wound Size Binds to CD44 receptor promoting pro-mitotic effects [249] Ointment ↑ Blood Flow on endothelial cells and angiogenesis ↑ Capillary Density ↑ p-Src, p-ERK and TGF-β1

et al aimed to expedite wound healing through the use of human ASCs on the outer layer of medium sized vessels in the wound bed, ostensibly differentiating into a supportive vascular role on PEG-fibrin (FPEG) gels performing a similar role to that of pericytes or vSMCs. In ASCs grown [300]. ASCs were derived from discarded burn skin samples and cul- on FPEG scaffolds in vitro, VEGFA gene expression was found to increase tured on a 3D PEG-fibrin gel. After three days in culture, ASCs formed over the culture period along with VEGF-A protein expression over the tubule-like structures like those created by endothelial cells, however course of 15 days. These results indicate that, under specificconditions, they lacked PECAM-1 and vWF expression. Gene expression of pericyte ASCs can differentiate into cells that both encourage neovascularization markers including for alpha actin-2 (ACTA2), PDGFRB, neural/glial through VEGF secretion and stabilize vessels through pericyte-like antigen-2 (NG2), angiopoietin-1 (ANGPT1), and angiopoietin-2 interactions. (ANGPT2) were detected via RT-PCR. This result was confirmed via im- While particular cellular treatments are meant to facilitate cell sur- munostaining of NG2 and PDGFR-β, suggesting a supportive vascular vival or differentiation so that they might have a direct effect on phenotype. To test the wound healing potential of ASCs, researchers wound healing, others can have an effect via paracrine signaling. In pa- employed a full-thickness excisional wound model in rats. Although pers by Garg et al and Kosaraju et al, pullulan-collagen hydrogels were wounds appeared to close at similar rates, histological analysis revealed created that increased bone marrow-derived mesenchymal progenitor that vascularization of the wound bed occurred earlier and to a higher cell (BM-MPC) recruitment to injury sites [301,302]. Pullulan is a non- degree in the ASC-FPEG treated group. A human specific mitochondrial immunogenic, hemocompatible, material comprised of polysaccha- antigen was used to determine if these differentiated ASCs played a di- rides. To test progenitor cell recruitment, Kosaraju et al used a parabio- rect role in wound healing and vascularization. Human cells were found sis model in which the vasculature of a GFP-expressing reporter mouse

Table 7 Adipose-derived stem cell therapies for enhancing wound angiogenesis.

Cell type Scaffold Treatment Result Exp. model Ref.

mASCs Pullulan-collagen hydrogel None ↑Recruitment of BM-MPCs Mice [302] ↑Provasculogenic gene expression hASCs Pullulan-collagen hydrogel Capillary Seeding ↑Wound healing speed Mice [301] ↑Vascularity hASCs PEG-fibrin gel None ↑Remodeling speed Rats [300] ↑Vessel formation pASCs Integra collagen scaffold None ↑Vascular density Pigs [330] ↑Number of mature vessels rASCs Cell sheet Ascorbic acid ↑Wound closure speed Rats [331] ↑Vessel density hASCs None Low-Level Light Therapy ↑VEGF and HIF-1α secretion, Mice [332] ↑Cell survival ↑Perfusion recovery rASCs ncl-PADM None ↑Elastic modulus of graft Rats [333] ↑Cellular and vascular infiltration 116 A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125 and a wild-type mouse were connected. This model allows for the trans- as opposed to fibrosis [303–306]. In one study, researchers used a full- fer of cells between mice to be observed. Two dorsal excisional wounds thickness cutaneous wound model in rats to study the wound healing were created on the backs of the wild-type mice. These wounds were potential of rat MSCs [307]. The cells were treated with treated with ASC-seeded hydrogels, an injection of ASCs in PBS, or a and ascorbic acid (vitamin C) and cultured to create a tightly packed, PBS control injection. Cells from the unwounded, GFP expressing, mice confluent, sheet of cells. After 12 days, the cell-aggregate layer began then migrated to the injury sites on the WT mice. BM-MPCs comprised to detach from the culture surface and was used in place of a scaffold of 23%, 8.4%, and 2.1% of the total cells recruited to the injury site in ASC for therapeutic delivery. Evaluation of gene expression showed that hydrogel implant, ASC injection, and PBS control treatments respec- pro-wound healing markers collagen 1 (COL1) and transforming tively. Furthermore, a subpopulation of these BM-MPCs was further growth factor beta (TGF-β)weresignificantly elevated in cells that identified using single cell gene expression profile analysis. This popula- grew in the aggregate form, suggesting superior wound healing poten- tion included the expression of known provasculogenic and remodeling tial. A full thickness cutaneous wound model was performed in rats to genes such as angiogenin (ANG), endosialin (CD248), NANOG, NFKB1, test wound healing. The cell aggregate was placed on the wound bed stem cell antigen-1 (SCA1), and extracellular superoxide dismutase with three layers of sterilized porcine small intestine submucosa to pro- (SOD3). This enhanced wound healing gene profile was expressed by tect the cells from the wound dressing. Wound size after four weeks was a greater amount of BM-MPCs for the ASC hydrogel condition compared significantly smaller (~50% smaller) for the aggregate group compared to ASC injection and PBS control. to the topically delivered cell suspension and intravenously delivered Conditioned media was collected from ASCs cultured on hydrogels cells. Furthermore, capillary density was found to be greatest in the ag- or ASCs cultured using glass or plastic cell culture surfaces[302]. Condi- gregate group. tioned media harvested from hydrogel-cultured ASCs significantly in- Inflammation was assessed for each group through RT-PCR and im- creased migration and proliferation in isolated BM-MPCs. Expression munofluorescence staining. After week two, gene expression showed of genes relating to wound healing, cellular recruitment, and neo- significantly lower TNF-α and IL-1β levels compared to controls. Addi- vascular genes including chemokine ligand 2 (CCL2), matrix tionally, cell aggregate groups showed lower inflammatory marker metalloproteinase-3 (MMP3), hepatocyte growth factor (HGF), hypoxia levels compared to the other two treatment groups. Gene expression inducible factor 1α (HIF1α), and endoglin (ENG) was significantly ele- for inducible nitric oxide synthase (iNOS) was found to be higher for vated in BM-MPC cultures with hydrogel cultured conditioned media the aggregate group compared to both the control and other treatment compared to control media. In addition, they observed increased pro- groups. Taken together, these results indicate a lower level of inflamma- tein levels of HGF and MMP3. Conditioned media from ASCs grown in tion and a pro-healing wound environment. Furthermore, the presence hydrogels increased the tube formation by BM-MPCs in comparison to of CD45+ lymphocytes was reduced and BM-MSC engraftment was im- control groups. Overall, this study showed that ASCs grown on this proved at the wound bed after two weeks. This study did not show that pullulan-collagen scaffold enhanced angiogenesis through paracrine MSCs were directly involved in the formation of new vasculature, in- mechanisms and increased recruitment of circulating BM-MPCs to the stead the authors attribute the improved wound healing to inflamma- wound area. tory regulation. Furthermore, this work represents a departure from a substantial portion of tissue engineering research which utilizes an in- dependent scaffold for cell survival and engraftment by achieving a pos- 5.3. Bone marrow derived mesenchymal stem cells itive effect with a relatively simple cell aggregate approach. This method avoids issues that are normally considered with synthetic scaffolds such Bone marrow derived mesenchymal stem cells (MSCs) are also as degradation/absorption, cytotoxicity, and immune response. While promising candidates as a cellular therapy for enhancing wound healing this technique proved effective in dermal wound healing in this work, (Table 8). MSCs have been found to modulate inflammation at wound further investigation has opened the door to other wound healing sites as well as influence surrounding cells to encourage regeneration arenas such as cardiac tissue repair, with areas soon to follow. [308,309]

Table 8 Cell-based therapies for enhancing wound angiogenesis.

Cell Type Scaffold Treatment Result Ref.

rBM-MSCs Cell Aggregate Matrix Dexamethasone/ Ascorbic Acid ↑ Wound closure [307] ↑ Capillary density ↓ Inflammation mBM-MSCs PEG-PU None ↑ Anti-inflammatory cytokines [310] ↑ Wound closure ↓ Oxidative stress ↓ Pro-inflammatory cytokines ↓ ROS at wound site rBM-MSCs Graphene Foam None ↑ Wound closure [334] ↑ Neo-vascularization ↑ VEGF and FGF-2 mRNA ↓ Fibrotic scar formation hiPSCs HA Hydrogel VEGF and ↑ Cellular integration into vasculature [335] SB-431542 hiPSCs None VEGF, BMP-4, and FGF-2 ↑ Tubule formation [313] ↑ Vascular density ↑ Wound closure PMSCs None None ↑ Wound closure ↑ Vascular density ↑ Cellular integration ↑Integrin β1, Integrin β3, VEGF ↓ICAM SKPs Collagen Sponge None ↑ Capillary regeneration [336] Monocytes iMonogrid Gel Monocytes treated toward M2 phenotype ↑ Vascular density [337] A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125 117

While MSCs have been shown to act as modulators of immune re- embedded in the scaffold. By day 7, the MSC-scaffold group appeared sponse, they are still susceptible to the potentially cytotoxic environ- to have the greatest level of wound closure, with approximately five ment of the wound bed. Geesala et al sought to compliment the anti- times the number cells at the site of injury than the control. Significant inflammatory, pro-regenerative, function of MSCs by including a scaf- upregulation of pro-inflammatory cytokine gene expression was ob- fold composed of polyethylene glycol and polyurethane (PEG-PU) served in the MSC-only group compared to both the scaffold-only and [310]. Reactive oxygen species can impair stem cell function and reduce MSC-scaffold groups, indicating that the scaffold may reduce inflamma- wound healing [311,312]. This group aimed to modulate ROS levels and tion and could protect transplanted cells an immune response. The increase the efficacy of their cellular treatment through the combined presence of ROS in the tissue was evaluated using oxidation sensitive activity of MSCs and a PEG-PU scaffold. The PEG-PU scaffold was synthe- dihydroethidium dye staining. These results indicated that the sized using castor oil, a substance that contains elevated levels of vita- scaffold-only and MSC-scaffold groups had the least ROS activity while min E. In vehicle control studies, vitamin E has shown to increase the the MSC-only group had the greatest. Consistent with these findings, activity of the antioxidant enzyme, Cu/ZnSOD. Preliminary work was gene expression for the antioxidant enzymes catalase and GPX2, a glu- performed to create the interpenetrating polymer network, test scaffold tathione peroxidase, was significantly elevated for the MSC-scaffold biocompatibility, and to ensure cellular migration into and out of the group compared to free MSCs. scaffold. The protective role of the scaffold was assessed by varying con- To test successful engraftment at the wound site, MSC-indicative centrations of hydrogen peroxide to induce oxidative stress in cells that fluorescent markers including CD133-PE and CD90.2-APC were used were cultured either with or without scaffolds. Each of the cell-only on tissue samples. The MSC-scaffold group showed the largest amount groups exhibited a dose dependent drop in proliferation based on an of fluorescence at days 7 and 10, suggesting better engraftment was MTT assay. All scaffold groups showed increased proliferation compared achieved with the scaffold. The MSC-scaffold group sections contained to the cell-only groups when they were MSCs treated with 1 μMand10 a significantly greater amount of PECAM-1 stained cells compared to

μMH2O2. Hoechst staining showed markedly fewer apoptotic nuclei be- the scaffold and MSC-only groups. In addition, there was increased tween the scaffold and non-scaffold groups. gene expression for angiogenesis related genes including those for The scaffolds were then tested in a full-thickness excisional wound VEGFR2, VEGFR3, and Tie-2 from tissues harvested from the wound model in C57BL/6J mice. The study consisted of four treatment groups site. Overall, the study demonstrates a novel and potentially effective including an untreated control, scaffold alone, free MSCs, and MSCs strategy for combining cellular therapeutics with hydrogel scaffolds to

Fig. 7. The angiogenic potential of hiPSC-derived and hESC-derived ECs (red) and vSMCs (green) to HUVEC/vSMC controls in a co-culture tube formation assay were compared. Equal numbers of cells across each group were cultured on Matrigel and allowed to form tubules overnight. Multiple EC to vSMC ratios were tested across all groups including 100:0, 60:40 and 40:60. (A) Fluorescent microscopy images of the tube formation assay. (B) Quantification of the tube area for the cells in the tube formation assay. *p b 0.05 versus 100:0 HUVEC group. Scale bar = 100 μm. Used with permission [313]. 118 A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125 improve the inhospitable area of the wound site to allow for expedited demonstrating expedited closure but also engraftment of PMSCs into healing. the functional tissues at the wound bed. With the relative ease of har- vesting this type of stem cell, PMSCs could seemingly see wider use in 5.4. Induced pluripotent stem cells similar studies and applications in the future. Skin progenitor cells have been explored as a therapy for wound Induced pluripotent stem cells (iPSCs) have recently received signif- healing and angiogenesis [318,319]. Ke et al. sought to examine the ef- icant attention as a cell type for improving angiogenesis and wound fects of skin-derived precursor cells (SKPs) on diabetic wound healing healing. Kim et al differentiated iPSCs into separate populations of endo- in T2D mice. They hypothesized that SKPs could aid in the healing pro- thelial cells and vSMCs [313]. They compared the angiogenic capability cess of chronic wound disease states by improving vascularization in the of these iPSC-derived vascular cells to primary somatic cells of the same area via differentiation and integration of SKPs. A collagen sponge was type. They first performed a co-culture tube formation assay using vas- used as a scaffold to aid the integration of SKPs into the dorsal, full- cular cells differentiated from iPSCs, cells differentiated from ECs, and thickness, skin wound of diabetic mice. Greater wound closure was ob- HUVECs/hvSMCs at varying ratios. Both differentiated cell groups exhib- served for the SKP/collagen as well as the collagen only groups com- ited cooperation between ECs and vSMCs and greater tubule area com- pared to the saline control at day 7 but slight difference was observed pared to the HUVEC/hvSMC group (Fig. 7A, B). They then injected the between the groups at day 14. Furthermore, greater capillary densities cell groups subcutaneously into athymic nu/nu mice suspended in a were observed in the SKP/collagen group via αSMA and isolectin stain- mixture of Matrigel and collagen I. After two weeks, they found in- ing of histology at the wound bed. The authors also observed potential creased perfused vessel area and vascular density in the groups treated neuronal regeneration alongside vascular formation through the addi- with the iPSC-derived cells in comparison to groups treated with the gel tion of nestin staining. SKPs appear to have a vast potential in the alone or somatic endothelial cell/vSMCs in the gel. Through proteomic wound healing arena as they can differentiate towards three crucial profiling of the secreted factors they found that iPSC-derived vascular would heal cell lineages (vSMCs, ECs, and neurons). Paired with an ef- cell produced more pro-angiogenic soluble factors. They also implanted fective scaffold, difficulties involving engraftment and cell survival can the cells in a dermal wound model in nu/nu mice and found that the be mitigated and the effect of SKPs can be enhanced as seen in the pre- iPSC-derived vascular cells were superior in their ability to induce ciously discussed article. Altogether, these cells present another promis- wound healing and angiogenesis in the wound bed. Lee et al found ing option for an easily sourced adult stem cell therapy for wound that iPSC-derived lymphatic endothelial cells also enhanced wound healing. healing and vascular growth [314]. Using ear and dorsal wound models in nu/nu mice, they observed incorporation of the cells into the lym- phatic vessel network and a rise in lymphvasculogenesis and 6. Conclusion lymphangiogenesis. In addition, the iPSC-derived lymphatic endothelial cells induced enhance wound healing in comparison to control (PBS) or The development of effective therapeutics for enhancing wound an- human primary lymphatic endothelial cell-treated wounds. Zhang et al giogenesis, particularly for non-healing or complex wounds, requires also tested the ability of exosomes produced by iPSC-derived MSCs to the ability of the compound to work in the harsh environment of a induced enhanced wound healing and angiogenesis [315]. After poorly healing wound. As discussed in this review, many compounds confirming the activity of the exosomes through proliferation/migration have shown great promise in enhancing wound healing and angiogen- and tube formation assays, exosomes were subcutaneously introduced esis in animal models of wound healing. In the past, many wound in a dorsal cutaneous injury model in rats. Wound closure after two healing treatments that have been successful in preclinical models ulti- weeks was significantly higher for the exosome-treated group com- mately fail to provide benefits to patients when tested in clinical trials. pared to media controls and they observed an increase in collagen and Fundamentally, the field suffers from the lack of a validated preclinical elastin deposition. Furthermore, wound epithelialization and vessel animal model of chronic wound healing that correlates well with density were significantly increased while scar width was significantly human wound healing in the clinic. While many approaches have decreased. Overall, these works support that iPSC-derived cells have po- been used to reduce wound healing in animal models, there is no con- tential for enhancing wound healing and angiogenesis. sensus as to the model with the best correlation to compromised human wound healing [320]. A change in approach and experimental 5.5. Other stem cell populations mindset is needed in how studies are designed for assessing promising wound therapies. In vitro studies should be targeted not solely at A number of other stem cell types have also been investigated in assessing activity in healthy cells but cells either from diseased patients wound healing and angiogenesis. Placenta-derived mesenchymal stem or in combination with treatments that inhibit the angiogenic or wound cells (PMSCs) wound models have been studied in both rats and mice healing response. The baseline assumption should be one of compro- [316,317]. Kong et al. employed a diabetic, Goto-Kakizaki, rat model mised response and therapeutic resistance, rather than healthy healing. with full thickness dorsal skin wounds. Fluorescently labeled PMSCs New treatments should be designed or evaluated in the context of their were injected intradermally at the wound site and monitored until all ability work in compromised wound environments. wounds closed. Compared to non-treated controls, PMSC-treated Ultimately, effective wound healing is a temporal process that can wounds demonstrated expedited wound closure as well as increased potentially be compromised by disease or infection in a number of vascular density as the wound site. Furthermore, PMSCs were con- ways. In diabetic wound healing, studies have identified many firmed to incorporate into the host vascular network visualized through compromising factors for wound healing including the loss of heparan immunofluorescence histology. Similarly, Abd-Allah et al. employed a sulfate proteoglycans [47,48,54], alterations in inflammatory response skin wound model in wild type mice for 7 and 12 days. Like the previous [321], increased proteolysis[322] and alterations in ROS signaling experiment, PMSCs delivered either intraperitoneally or intraregionally [323], to name only a few. It is difficult to imagine therapy with any sin- induced more rapid healing when compared to the untreated group. gle agent to be able to address more than a few of these factors, pointing Additionally, RT-PCR detected higher expression of angiogenic and to the need for combination therapies that address the issues of a pa- wound healing factors including Integrin β1, Integrin β3, and a decrease tient’s particular wound and diagnostic assays to delineate when a par- in ICAM expression. RT-PCR was also used to detect the presence of ticular compound should be used. In this context, the biology and human cells in the form of albumin and GAPDH expression. An ELISA compromised mechanisms of wound healing must be factored into also revealed greater VEGF in the PMSC-treated tissues. Both of these the design and development of treatments and delivery platforms for studies lend merit to PMSCs as wound healing agents by not only addressing chronic wounds. A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125 119

Acknowledgements [27] R. Sridharan, A.R. Cameron, D.J. Kelly, C.J. Kearney, F.J. O'Brien, Biomaterial based modulation of macrophage polarization: a review and suggested design principles, Mater Today 18 (2015) 313–325. The authors gratefully acknowledge funding through the American [28] K.L. Spiller, T.J. Koh, Macrophage-based therapeutic strategies in regenerative med- Heart Association (17IRG33410888), the DOD CDMRP (W81XWH-16- icine, Adv Drug Deliv Rev 122 (2017) 74–83. [29] R. Sridharan, A.R. Cameron, D.J. Kelly, C.J. Kearney, F.J. O’Brien, Biomaterial based 1-0580; W81XWH-16-1-0582) and the National Institutes of Health modulation of macrophage polarization: a review and suggested design principles, (1R21EB023551-01; 1R21EB024147-01A1; 1R01HL141761-01) to Mater Today 18 (2015) 313–325. ABB. The authors also acknowledge support through a graduate fellow- [30] M.G. Tonnesen, X. Feng, R.A. Clark, Angiogenesis in wound healing, J Investig – ship from the NSF (DGE-1610403) to A.S. Dermatol Symp Proc 5 (2000) 40 46. [31] G.L. Wang, B.H. Jiang, E.A. Rue, G.L. Semenza, Hypoxia-inducible factor 1 is a basic- helix-loop-helix-PAS heterodimer regulated by cellular O2 tension, Proc Natl Acad References Sci USA 92 (1995) 5510–5514. [32] A. Ahluwalia, A.S. Tarnawski, Critical role of hypoxia sensor–HIF-1alpha in VEGF [1] R.J. Snyder, R.S. Kirsner, R.A. Warriner, L.A. Lavery, J.R. Hanft, P. Sheehan, Consensus gene activation. Implications for angiogenesis and tissue injury healing, Curr Med recommendations on advancing the standard of care for treating neuropathic foot Chem 19 (2012) 90–97. ulcers in patients with diabetes, Ostomy Wound Manage 56 (2010) S1–24. [33] M.A. Nugent, R.V. Iozzo, -2, Int J Biochem Cell Biol 32 [2] N.J. Park, L. Allen, V.R. Driver, Updating on understanding and managing chronic (2000) 115–120. wound, Dermatol Ther 26 (2013) 236–256. [34] S. Ashikari-Hada, H. Habuchi, Y. Kariya, K. Kimata, Heparin regulates vascular en- [3] C.K. Sen, G.M. Gordillo, S. Roy, R. Kirsner, L. Lambert, T.K. Hunt, F. Gottrup, G.C. dothelial growth factor165-dependent mitogenic activity, tube formation, and its Gurtner, M.T. Longaker, Human skin wounds: a major and snowballing threat receptor phosphorylation of human endothelial cells: Comparison of the effects to public health and the economy, Wound Repair Regenerat. 17 (2009) of heparin and modified heparins, J Biol Chem 280 (2005) 31508–31515. 763–771. [35] L.A. DiPietro, Angiogenesis and wound repair: when enough is enough, J Leukoc [4] A.J. Singer, R.A. Clark, Cutaneous wound healing, New Engl J Med 341 (1999) Biol 100 (2016) 979–984. 738–746. [36] M.S. Wietecha, M.J. Krol, E.R. Michalczyk, L. Chen, P.G. Gettins, L.A. DiPietro, Pig- [5] G. Crovetti, G. Martinelli, M. Issi, M. Barone, M. Guizzardi, B. Campanati, M. Moroni, ment epithelium-derived factor as a multifunctional regulator of wound healing, A. Carabelli, Platelet gel for healing cutaneous chronic wounds, Transf Apheresis Sci Am J Physiol Heart Circ Physiol 309 (2015) H812–H826. 30 (2004) 145–151. [37] M.S. Wietecha, L. Chen, M.J. Ranzer, K. Anderson, C. Ying, T.B. Patel, L.A. DiPietro, [6] H. Brem, O. Stojadinovic, R.F. Diegelmann, H. Entero, B. Lee, I. Pastar, M. Golinko, H. Sprouty2 downregulates angiogenesis during mouse skin wound healing, Am J Rosenberg, M. Tomic-Canic, Molecular markers in patients with chronic wounds to Physiol Heart Circ Physiol 300 (2011) H459–H467. guide surgical debridement, Mol Med 13 (2007) 30–39. [38] G. Bergers, S. Song, The role of pericytes in blood-vessel formation and mainte- [7] D.G. Armstrong, V.A. Kanda, L.A. Lavery, W. Marston, J.L. Mills, A.J. Boulton, Mind nance, Neuro-Oncol 7 (2005) 452–464. the gap: disparity between research funding and costs of care for diabetic foot ul- [39] C. Hellberg, A. Ostman, C.H. Heldin, PDGF and vessel maturation, Recent Results cers, Diabetes Care 36 (2013) 1815–1817. Cancer Res 180 (2010) 103–114. [8] J. Larsson, C.D. Agardh, J. Apelqvist, A. Stenstrom, Long-term prognosis after healed [40] M. Felcht, R. Luck, A. Schering, P. Seidel, K. Srivastava, J. Hu, A. Bartol, Y. Kienast, C. amputation in patients with diabetes, Clin Orthop Relat Res (1998) 149–158. Vettel, E.K. Loos, S. Kutschera, S. Bartels, S. Appak, E. Besemfelder, D. Terhardt, E. [9] E.A. Ayello, What does the wound say? Why determining etiology is essential for Chavakis, T. Wieland, C. Klein, M. Thomas, A. Uemura, S. Goerdt, H.G. Augustin, appropriate wound care, Adv Skin Wound Care 18 (2005) 98–109 quiz 110-101. Angiopoietin-2 differentially regulates angiogenesis through TIE2 and integrin sig- [10] K. Alexiadou, J. Doupis, Management of Diabetic Foot Ulcers, Diabetes Therapy : naling, J Clin Invest 122 (2012) 1991–2005. Research, Treatment and Education of Diabetes and Related Disorders, vol. 3, [41] N.W. Gale, G. Thurston, S.F. Hackett, R. Renard, Q. Wang, J. McClain, C. Martin, C. 2012 4. Witte, M.H. Witte, D. Jackson, C. Suri, P.A. Campochiaro, S.J. Wiegand, G.D. [11] L.K. Branski, G.G. Gauglitz, D.N. Herndon, M.G. Jeschke, A review of gene and stem Yancopoulos, Angiopoietin-2 is required for postnatal angiogenesis and lymphatic cell therapy in cutaneous wound healing, Burns 35 (2009) 171–180. patterning, and only the latter role is rescued by Angiopoietin-1, Dev Cell 3 (2002) [12] B. Buchberger, M. Follmann, D. Freyer, H. Huppertz, A. Ehm, J. Wasem, The evi- 411–423. dence for the use of growth factors and active skin substitutes for the treatment [42] L. Hakanpaa, T. Sipila, V.M. Leppanen, P. Gautam, H. Nurmi, G. Jacquemet, L. Eklund, of non-infected diabetic foot ulcers (DFU): a health technology assessment J. Ivaska, K. Alitalo, P. Saharinen, Endothelial destabilization by angiopoietin-2 via (HTA), Exp Clin Endocrinol Diabetes 119 (2011) 472–479. integrin beta1 activation, Nat Commun 6 (2015) 5962. [13] M. Sonal Sekhar, M.K. Unnikrishnan, K. Vijayanarayana, G.S. Rodrigues, C. [43] M. Teichert, L. Milde, A. Holm, L. Stanicek, N. Gengenbacher, S. Savant, T. Mukhopadhyay, Topical application/formulation of probiotics: will it be a novel Ruckdeschel, Z. Hasanov, K. Srivastava, J. Hu, S. Hertel, A. Bartol, K. Schlereth, treatment approach for ? Med Hypotheses 82 (2014) 86–88. H.G. Augustin, Pericyte-expressed Tie2 controls angiogenesis and vessel matura- [14] K. Futrega, M. King, W.B. Lott, M.R. Doran, Treating the whole not the hole: neces- tion, Nat Commun 8 (2017), 16106. . sary coupling of technologies for diabetic foot ulcer treatment, Trends Mol Med 20 [44] J. Cai, O. Kehoe, G.M. Smith, P. Hykin, M.E. Boulton, The angiopoietin/Tie-2 system (2014) 137–142. regulates pericyte survival and recruitment in diabetic retinopathy, Invest [15] N. Greer, N.A. Foman, R. MacDonald, J. Dorrian, P. Fitzgerald, I. Rutks, T.J. Wilt, Ad- Ophthalmol Vis Sci 49 (2008) 2163–2171. vanced wound care therapies for nonhealing diabetic, venous, and arterial ulcers: a [45] H. Brem, M. Tomic-Canic, Cellular and molecular basis of wound healing in diabe- systematic review, Ann Intern Med 159 (2013) 532–542. tes, J Clin Invest 117 (2007) 1219–1222. [16] N. Papanas, E. Maltezos, Becaplermin gel in the treatment of diabetic neuropathic [46] W. Qi, C. Yang, Z. Dai, D. Che, J. Feng, Y. Mao, R. Cheng, Z. Wang, X. He, T. Zhou, X. foot ulcers, Clin Interv Aging 3 (2008) 233–240. Gu, L. Yan, X. Yang, J.X. Ma, G. Gao, High levels of pigment epithelium-derived fac- [17] N. Papanas, E. Maltezos, Growth factors in the treatment of diabetic foot ulcers: tor in diabetes impair wound healing through suppression of Wnt signaling, Diabe- new technologies, any promises? Int J Low Extrem Wounds 6 (2007) 37–53. tes 64 (2015) 1407–1419. [18] U.A. Okonkwo, L.A. DiPietro, Diabetes and wound angiogenesis, Int J Mol Sci (2017) [47] S. Das, G. Singh, M. Majid, M.B. Sherman, S. Mukhopadhyay, C.S. Wright, P.E. 18. Martin, A.K. Dunn, A.B. Baker, Syndesome therapeutics for enhancing diabetic [19] S. Das, A.B. Baker, Biomaterials and nanotherapeutics for enhancing skin wound wound healing, Adv Healthc Mater 5 (2016) 2248–2260. healing, Front Bioeng Biotechnol 4 (2016) 82. [48] A.J. Monteforte, B. Lam, S. Das, S. Mukhopadhyay, C.S. Wright, P.E. Martin, A.K. [20] C.J. Ferrante, S.J. Leibovich, Regulation of macrophage polarization and wound Dunn, A.B. Baker, Glypican-1 nanoliposomes for potentiating growth factor activity healing, Adv Wound Care (New Rochelle) 1 (2012) 10–16. in therapeutic angiogenesis, Biomaterials 94 (2016) 45–56. [21] D.B. Gurevich, C.E. Severn, C. Twomey, A. Greenhough, J. Cash, A.M. Toye, H. Mellor, [49] S.L. Drinkwater, A. Smith, B.M. Sawyer, K.G. Burnand, Effect of venous ulcer exu- P. Martin, Live imaging of wound angiogenesis reveals macrophage orchestrated dates on angiogenesis in vitro, Br. J. Surg. 89 (2002) 709–713. vessel sprouting and regression, EMBO J. 37 (2018). [50] S.A. Eming, M. Koch, A. Krieger, B. Brachvogel, S. Kreft, L. Bruckner-Tuderman, T. [22] R. Mirza, L.A. DiPietro, T.J. Koh, Selective and specific macrophage ablation is detri- Krieg, J.D. Shannon, J.W. Fox, Differential proteomic analysis distinguishes tissue mental to wound healing in mice, Am J Pathol. 175 (2009) 2454–2462. repair biomarker signatures in wound exudates obtained from normal healing [23] I. Goren, N. Allmann, N. Yogev, C. Schurmann, A. Linke, M. Holdener, A. Waisman, J. and chronic wounds, J. Proteome Res. 9 (2010) 4758–4766. Pfeilschifter, S. Frank, A transgenic mouse model of inducible macrophage deple- [51] G. Lauer, S. Sollberg, M. Cole, I. Flamme, J. Sturzebecher, K. Mann, T. Krieg, S.A. tion effects of diphtheria toxin-driven lysozyme M-specific cell lineage ablation Eming, Expression and proteolysis of vascular endothelial growth factor is in- on wound inflammatory, angiogenic, and contractive processes, Am J Pathol 175 creased in chronic wounds, J Invest Dermatol 115 (2000) 12–18. (2009) 132–147. [52] S.A. Eming, G. Lauer, M. Cole, S. Jurk, H. Christ, C. Hornig, T. Krieg, H.A. Weich, In- [24] N. Jetten, S. Verbruggen, M.J. Gijbels, M.J. Post, M.P. De Winther, M.M. Donners, creased levels of the soluble variant of the vascular endothelial growth factor re- Anti-inflammatory M2, but not pro-inflammatory M1 macrophages promote an- ceptor VEGFR-1 are associated with a poor prognosis in wound healing, J Invest giogenesis in vivo, Angiogenesis 17 (2014) 109–118. Dermatol 123 (2004) 799–802. [25] E. Zajac, B. Schweighofer, T.A. Kupriyanova, A. Juncker-Jensen, P. Minder, J.P. [53] S. Barrientos, O. Stojadinovic, M.S. Golinko, H. Brem, M. Tomic-Canic, Growth fac- Quigley, E.I. Deryugina, Angiogenic capacity of M1- and M2-polarized macro- tors and cytokines in wound healing, Wound Repair Regenerat 16 (2008) phages is determined by the levels of TIMP-1 complexed with their secreted 585–601. proMMP-9, Blood 122 (2013) 4054–4067. [54] S. Das, G. Singh, A.B. Baker, Overcoming disease-induced growth factor resistance [26] J.M. Anderson, A. Rodriguez, D.T. Chang, Foreign body reaction to biomaterials, in therapeutic angiogenesis using recombinant co-receptors delivered by a liposo- Semin Immunol 20 (2008) 86–100. mal system, Biomaterials 35 (2014) 196–205. 120 A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125

[55] M.C. Robson, T.A. Mustoe, T.K. Hunt, The future of recombinant growth factors in [84] L. Wang, X. Wu, T. Shi, L. Lu, Epidermal growth factor (EGF)-induced corneal epi- wound healing, Am. J. Surg. 176 (1998) 80S–82S. thelial wound healing through nuclear factor kappaB subtype-regulated CCCTC [56] W.A. Woodley, T. David, DeClerck Brittany, Chen Mei, Li Wei, Keratinocyte migra- binding factor (CTCF) activation, J. Biol. Chem. 288 (2013) 24363–24371. tion and a hypothetical new role for extracellular heat shock protein 90 alpha in [85] N. Fullard, A. Moles, S. O'Reilly, J.M. van Laar, D. Faini, J. Diboll, N.J. Reynolds, D.A. orchestrating skin wound healing, Adv Wound Care 4 (2015) 203–212. Mann, J. Reichelt, F. Oakley, The c-Rel subunit of NF-kappaB regulates epidermal [57] T.J. Wieman, Clinical efficacy of becaplermin (rhPDGF-BB) gel. Becaplermin Gel homeostasis and promotes skin fibrosis in mice, Am. J. Pathol. 182 (2013) Studies Group, Am. J. Surg. 176 (1998) 74S–79S. 2109–2120. [58] R.C. Fang, R.D. Galiano, A review of becaplermin gel in the treatment of diabetic [86] J.W. Park, S.R. Hwang, I.S. Yoon, Advanced growth factor delivery systems in neuropathic foot ulcers, Biol. Ther. Dent. 2 (2008) 1–12. wound management and skin regeneration, Molecules 22 (2017). [59] D.L. Steed, Clinical evaluation of recombinant human platelet-derived growth fac- [87] E.S. Gil, B. Panilaitis, E. Bellas, D.L. Kaplan, Functionalized silk biomaterials for tor for the treatment of lower extremity diabetic ulcers. Diabetic Ulcer Study wound healing, Adv Healthc Mater 2 (2013) 206–217. Group, J. Vasc. Surg. 21 (1995) 71–78 discussion 79-81. [88] A. Schneider, X.Y. Wang, D.L. Kaplan, J.A. Garlick, C. Egles, Biofunctionalized [60] T.J. Wieman, J.M. Smiell, Y. Su, Efficacy and safety of a topical gel formulation of re- electrospun silk mats as a topical bioactive dressing for accelerated wound healing, combinant human platelet-derived growth factor-BB (becaplermin) in patients Acta Biomater. 5 (2009) 2570–2578. with chronic neuropathic diabetic ulcers. A phase III randomized placebo- [89] S. Kondo, Y. Kuroyanagi, Development of a wound dressing composed of controlled double-blind study, Diabetes Care 21 (1998) 822–827. hyaluronic acid and collagen sponge with epidermal growth factor, J. Biomater. [61] J.M. Smiell, T.J. Wieman, D.L. Steed, B.H. Perry, A.R. Sampson, B.H. Schwab, Effi- Sci. Polym. Ed. 23 (2012) 629–643. cacy and safety of becaplermin (recombinant human platelet-derived growth [90] J.T. Hardwicke, J. Hart, A. Bell, R. Duncan, D.W. Thomas, R. Moseley, The effect of factor-BB) in patients with nonhealing, lower extremity diabetic ulcers: a com- dextrin-rhEGF on the healing of full-thickness, excisional wounds in the (db/db) binedanalysisoffourrandomizedstudies, Wound Repair Regeneration 7 diabetic mouse, J Control Release 152 (2011) 411–417. (1999) 335–346. [91] T. Imamura, Physiological functions and underlying mechanisms of fibroblast [62] R.C. Fang, R.D. Galiano, A review of becaplermin gel in the treatment of diabetic growth factor (FGF) family members: recent findings and implications for their neuropathic foot ulcers, Biologics (2008) 1–12. pharmacological application, Biol. Pharm. Bull. 37 (2014) 1081–1089. [63] United States Food and Drug Administration, Product Description Regranex., [92] A. Komi-Kuramochi, M. Kawano, Y. Oda, M. Asada, M. Suzuki, J. Oki, T. Accessed 3/16/2018. Imamura, Expression of fibroblast growth factors and their receptors during [64] A.M. Gilligan, C.R. Waycaster, C.T. Milne, Cost effectiveness of becaplermin gel on full-thickness skin wound healing in young and aged mice, J. Endocrinol. 186 wound closure for the treatment of pressure injuries, Wounds 30 (6) (2018) (2005) 273–289. 197–204. [93] S. Werner, K.G. Peters, M.T. Longaker, F. Fuller-Pace, M.J. Banda, L.T. Williams, Large [65] G. Bergers, S. Song, The role of pericytes in blood-vessel formation and mainte- induction of keratinocyte growth factor expression in the dermis during wound nance1, Neuro Oncol (2005) 452–464. healing, Proc. Natl. Acad. Sci. U. S. A. 89 (1992) 6896–6900. [66] A.N. Stratman, A.E. Schwindt, K.M. Malotte, G.E. Davis, Endothelial-derived PDGF- [94] R. Ronca, A. Giacomini, M. Rusnati, M. Presta, The potential of fibroblast growth fac- BB and HB-EGF coordinately regulate pericyte recruitment during vasculogenic tor/fibroblast signaling as a therapeutic target in tumor an- tube assembly and stabilization, Blood 116 (2010) 4720–4730. giogenesis, Expert Opin. Ther. Targets 19 (2015) 1361–1377. [67] R. Cao, E. Brakenhielm, R. Pawliuk, D. Wariaro, M.J. Post, E. Wahlberg, P. Leboulch, [95] S. Barrientos, H. Brem, O. Stojadinovic, M. Tomic-Canic, Clinical application of Y. Cao, Angiogenic synergism, vascular stability and improvement of hind-limb is- growth factors and cytokines in wound healing, Wound Repair Regener. 22 chemia by a combination of PDGF-BB and FGF-2, Nat. Med. 9 (2003) 604–613. (2014) 569–578. [68] Z. Xie, C.B. Paras, H. Weng, P. Punnakitikashem, L.C. Su, K. Vu, L. Tang, J. Yang, K.T. [96] S. Werner, H. Smola, X. Liao, M.T. Longaker, T. Krieg, P.H. Hofschneider, L.T. Nguyen, Dual growth factor releasing multi-functional nanofibers for wound Williams, The function of KGF in morphogenesis of epithelium and healing, Acta Biomater. 9 (2013) 9351–9359. reepithelialization of wounds, Science 266 (1994) 819–822. [69] J. Ishihara, A. Ishihara, K. Fukunaga, K. Sasaki, M.J.V. White, P.S. Briquez, J.A. [97] S.I. Blaber, J. Diaz, M. Blaber, Accelerated healing in NONcNZO10/LtJ type 2 diabetic Hubbell, Laminin heparin-binding peptides bind to several growth factors and en- mice by FGF-1, Wound Repair Regen. 23 (2015) 538–549. hance diabetic wound healing, Nat. Commun. 9 (2018) 2163. [98] M.A. Seeger, A.S. Paller, The roles of growth factors in keratinocyte migration, Adv [70] H.J. Lai, C.H. Kuan, H.C. Wu, J.C. Tsai, T.M. Chen, D.J. Hsieh, T.W. Wang, Tailored de- Wound Care (New Rochelle), United States (2015) 213–224. sign of electrospun composite nanofibers with staged release of multiple angio- [99] C. Peng, B. Chen, H.K. Kao, G. Murphy, D.P. Orgill, L. Guo, Lack of FGF-7 further de- genic growth factors for chronic wound healing, Acta Biomater. 10 (2014) lays cutaneous wound healing in diabetic mice, Plast. Reconstr. Surg. 128 (2011) 4156–4166. 673e–684e. [71] S. Das, M. Majid, A.B. Baker, Syndecan-4 enhances PDGF-BB activity in diabetic [100] Y. Qu, C. Cao, Q. Wu, A. Huang, Y. Song, H. Li, Y. Zuo, C. Chu, J. Li, Y. Man, The dual wound healing, Acta Biomater. 42 (2016) 56–65. delivery of KGF and bFGF by collagen membrane to promote skin wound healing, J. [72] G.L. Brown, L.B. Nanney, J. Griffen, A.B. Cramer, J.M. Yancey, L.J. Curtsinger 3rd, L. Tissue Eng. Regen. Med. 12 (2018) 1508–1518. Holtzin, G.S. Schultz, M.J. Jurkiewicz, J.B. Lynch, Enhancement of wound healing [101] V.C. Lees, T.P. Fan, A freeze-injured skin graft model for the quantitative study of by topical treatment with epidermal growth factor, N. Engl. J. Med. 321 (1989) basic fibroblast growth factor and other promoters of angiogenesis in wound 76–79. healing, Br. J. Plast. Surg. 47 (1994) 349–359. [73] V. Falanga, W.H. Eaglstein, B. Bucalo, M.H. Katz, B. Harris, P. Carson, Topical use of [102] H.J. Park, S. Lee, K.H. Kang, C.Y. Heo, J.H. Kim, H.S. Yang, B.S. Kim, Enhanced random human recombinant epidermal growth factor (h-EGF) in venous ulcers, J Dermatol skin flap survival by sustained delivery of fibroblast growth factor 2 in rats, ANZ J. Surg Oncol 18 (1992) 604–606. Surg. 83 (2013) 354–358. [74] R.J. Bodnar, Epidermal growth factor and epidermal growth factor receptor: The [103] T. Bizenjima, F. Seshima, Y. Ishizuka, T. Takeuchi, T. Kinumatsu, A. Saito, Fibroblast Yin and Yang in the treatment of cutaneous wounds and cancer, Adv Wound growth factor-2 promotes healing of surgically created periodontal defects in rats Care (New Rochelle) 2 (2013) 24–29. with early, streptozotocin-induced diabetes via increasing cell proliferation and [75] J.P. Hong, H.D. Jung, Y.W. Kim, Recombinant human epidermal growth factor (EGF) regulating angiogenesis, J. Clin. Periodontol. 42 (2015) 62–71. to enhance healing for diabetic foot ulcers, Ann. Plast. Surg. 56 (2006) 394–398 dis- [104] X. Fu, Z. Shen, Y. Chen, J. Xie, Z. Guo, M. Zhang, Z. Sheng, Randomised placebo- cussion 399-400. controlled trial of use of topical recombinant bovine basic fibroblast growth factor [76] V.K. Mohan, Recombinant human epidermal growth factor (REGEN-D 150): effect for second-degree burns, Lancet 352 (1998) 1661–1664. on healing of diabetic foot ulcers, Diabetes Res. Clin. Pract. 78 (2007) 405–411. [105] X. Fu, Z. Shen, Y. Chen, J. Xie, Z. Guo, M. Zhang, Z. Sheng, Recombinant bovine basic [77] J. Berlanga, J.I. Fernandez, E. Lopez, P.A. Lopez, A. del Rio, C. Valenzuela, J. fibroblast growth factor accelerates wound healing in patients with burns, donor Baldomero, V. Muzio, M. Raices, R. Silva, B.E. Acevedo, L. Herrera, Heberprot-P: a sites and chronic dermal ulcers, Chin. Med. J. 113 (2000) 367–371. novel product for treating advanced diabetic foot ulcer, MEDICC Rev 15 (2013) [106] S. Akita, K. Akino, T. Imaizumi, A. Hirano, Basic fibroblast growth factor accelerates 11–15. and improves second-degree burn wound healing, Wound Repair Regen. 16 [78] A.B. Schreiber, M.E. Winkler, R. Derynck, Transforming growth factor-alpha - a (2008) 635–641. more potent angiogenic mediator than epidermal growth-factor, Science 232 [107] M. Sakamoto, N. Morimoto, S. Ogino, C. Jinno, T. Taira, S. Suzuki, Efficacy of gelatin (1986) 1250–1253. gel sheets in sustaining the release of basic fibroblast growth factor for murine skin [79] H.L. Tuyet, T.T. Nguyen Quynh, H. Vo Hoang Minh, D.N. Thi Bich, T. Do Dinh, D. Le defects, J. Surg. Res. 201 (2016) 378–387. Tan, H.L. Van, T. Le Huy, H. Doan Huu, T.N. Tran Trong, The efficacy and safety of [108] M. Fujita, M. Ishihara, M. Shimizu, K. Obara, S. Nakamura, Y. Kanatani, Y. Morimoto, epidermal growth factor in treatment of diabetic foot ulcers: the preliminary re- B. Takase, T. Matsui, M. Kikuchi, T. Maehara, Therapeutic angiogenesis induced by sults, Int. Wound J. 6 (2009) 159–166. controlled release of fibroblast growth factor-2 from injectable chitosan/non- [80] V.B. Mehta, G.E. Besner, HB-EGF promotes angiogenesis in endothelial cells via PI3- anticoagulant heparin hydrogel in a rat hindlimb ischemia model, Wound Repair kinase and MAPK signaling pathways, Growth Factors 25 (2007) 253–263. Regener 15 (2007) 58–65. [81] H. van Cruijsen, G. Giaccone, K. Hoekman, Epidermal growth factor receptor and [109] S. Das, A.J. Monteforte, G. Singh, M. Majid, M.B. Sherman, A.K. Dunn, A.B. Baker, angiogenesis: Opportunities for combined anticancer strategies, Int. J. Cancer 117 Syndecan-4 enhances therapeutic angiogenesis after hind limb ischemia in mice (2005) 883–888. with type 2 diabetes, Adv Healthc Mater 5 (2016) 1008–1013. [82] N.R. Johnson, Y. Wang, Controlled delivery of heparin-binding EGF-like growth fac- [110] E. Jang, H. Albadawi, M.T. Watkins, E.R. Edelman, A.B. Baker, Syndecan-4 proteoli- tor yields fast and comprehensive wound healing, J Control Release 166 (2013) posomes enhance fibroblast growth factor-2 (FGF-2)-induced proliferation, migra- 124–129. tion, and neovascularization of ischemic muscle, Proc. Natl. Acad. Sci. U. S. A. 109 [83] Y. Shirakata, R. Kimura, D. Nanba, R. Iwamoto, S. Tokumaru, C. Morimoto, K. (2012) 1679–1684. Yokota, M. Nakamura, K. Sayama, E. Mekada, S. Higashiyama, K. Hashimoto, [111] J. Zhang, L.M. Postovit, D. Wang, R.B. Gardiner, R. Harris, M. Abdul, A. Thomas, In Heparin-binding EGF-like growth factor accelerates keratinocyte migration and situ loading of basic fibroblast growth factor within porous silica nanoparticles skin wound healing, J. Cell Sci. 118 (2005) 2363–2370. for a prolonged release, Nanoscale Res. Lett. 4 (2009) 1297–1302. A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125 121

[112] W.Li,Y.Lan,R.Guo,Y.Zhang,W.Xue,Y.Zhang,Invitroandinvivoevaluationofa [137] S. Dhall, J.P. Silva, Y. Liu, M. Hrynyk, M. Garcia, A. Chan, J. Lyubovitsky, R.J. Neufeld, novel collagen/cellulose nanocrystals scaffold for achieving the sustained release of M. Martins-Green, Release of insulin from PLGA-alginate dressing stimulates re- basic fibroblast growth factor, J. Biomater. Appl. 29 (2015) 882–893. generative healing of burn wounds in rats, Clin. Sci. (Lond.) 129 (2015) [113] C.Y. Chu, J. Deng, L. Liu, Y.B. Cao, X.W. Wei, J.D. Li, Y. Man, Nanoparticles combined 1115–1129. with growth factors: recent progress and applications, RSC Adv. 6 (2016) [138] Y. Liu, M. Petreaca, M. Martins-Green, Cell and molecular mechanisms of insulin- 90856–90872. induced angiogenesis, J. Cell. Mol. Med. 13 (2009) 4492–4504. [114] N. Ferrara, H.P. Gerber, J. LeCouter, The biology of VEGF and its receptors, Nat. Med. [139] M.H.M. Lima, A.M. Caricilli, L.L. de Abreu, E.P. Araújo, F.F. Pelegrinelli, A.C.P. Thirone, 9(2003)669–676. D.M. Tsukumo, A.F.M. Pessoa, M.F. dos Santos, M.A. de Moraes, J.B.C. Carvalheira, [115] M. Simons, J.A. Ware, Therapeutic angiogenesis in cardiovascular disease, Nat. Rev. L.A. Velloso, M.J.A. Saad, Topical insulin accelerates wound healing in diabetes by Drug Discov. 2 (2003) 863–871. enhancing the AKT and ERK pathways: A double-blind placebo-controlled clinical [116] J.M. Isner, A. Pieczek, R. Schainfeld, R. Blair, L. Haley, T. Asahara, K. Rosenfield, S. trial, PLoS One 7 (5) (2012) e36974. Razvi, K. Walsh, J.F. Symes, Clinical evidence of angiogenesis after arterial gene [140] P. Giri, S. Ebert, U.D. Braumann, M. Kremer, S. Giri, H.G. Machens, A. Bader, Skin re- transfer of phVEGF165 in patient with ischaemic limb, Lancet 348 (1996) 370–374. generation in deep second-degree scald injuries either by infusion pumping or [117] I. Baumgartner, A. Pieczek, O. Manor, R. Blair, M. Kearney, K. Walsh, J.M. Isner, Con- topical application of recombinant human erythropoietin gel, Drug Des Dev Ther stitutive expression of phVEGF165 after intramuscular gene transfer promotes col- 9 (2015) 2565–2579. lateral vessel development in patients with critical limb ischemia, Circulation 97 [141] A. De Luisi, L. Binetti, R. Ria, S. Ruggieri, S. Berardi, I. Catacchio, V. Racanelli, V. (1998) 1114–1123. Pavone, B. Rossini, A. Vacca, D. Ribatti, Erythropoietin is involved in the angiogenic [118] T.D. Henry, K. Rocha-Singh, J.M. Isner, D.J. Kereiakes, F.J. Giordano, M. Simons, D.W. potential of bone marrow macrophages in multiple myeloma, Angiogenesis 16 Losordo, R.C. Hendel, R.O. Bonow, S.M. Eppler, T.F. Zioncheck, E.B. Holmgren, E.R. (2013) 963–973. McCluskey, Intracoronary administration of recombinant human vascular endo- [142] C.T. Wen, T. He, Y.Q. Xing, Erythropoietin promotes retinal angiogenesis in a mouse thelial growth factor to patients with coronary artery disease, Am. Heart J. 142 model, Mol. Med. Rep. 10 (2014) 2979–2984. (2001) 872–880. [143] B. Zhang, L. Xia, B. Hu, Y. Chen, J. Zhang, C. Zhu, B. Chen, Y. Wang, B. Wang, Effects of [119] T.D. Henry, B.H. Annex, G.R. McKendall, M.A. Azrin, J.J. Lopez, F.J. Giordano, P.K. recombinant human erythropoietin on angiogenesis in chronic ischemic porcine Shah, J.T. Willerson, R.L. Benza, D.S. Berman, C.M. Gibson, A. Bajamonde, A.C. myocardium, Zhonghua wai ke za zhi 52 (2014) 366–369. Rundle, J. Fine, E.R. McCluskey, V. Investigators, The VIVA trial: Vascular endothe- [144] C.M. McVicar, L.M. Colhoun, J.L. Abrahams, C.L. Kitson, R. Hamilton, R.J. Medina, D. lial growth factor in Ischemia for Vascular Angiogenesis, Circulation 107 (2003) Durga, T.A. Gardiner, P.M. Rudd, A.W. Stitt, Differential modulation of angiogenesis 1359–1365. by erythropoiesis-stimulating agents in a mouse model of ischaemic retinopathy, [120] N.N. Nissen, P.J. Polverini, A.E. Koch, M.V. Volin, R.L. Gamelli, L.A. DiPietro, Vascular PLoS One 5 (7) (2010) e11870. endothelial growth factor mediates angiogenic activity during the proliferative [145] N. Irrera, A. Bitto, G. Pizzino, M. Vaccaro, F. Squadrito, M. Galeano, F. Stagno phase of wound healing, Am. J. Pathol. 152 (1998) 1445–1452. d'Alcontres, F. Stagno d'Alcontres, M. Buemi, L. Minutoli, M.R. Colonna, D. [121] P. Losi, E. Briganti, A. Magera, D. Spiller, C. Ristori, B. Battolla, M. Balderi, S. Kull, A. Altavilla, Epoetin alpha and epoetin zeta: A comparative study on stimulation of Balbarini, R. Di Stefano, G. Soldani, Tissue response to poly(ether)urethane-polydi- angiogenesis and wound repair in an experimental model of burn injury, Biomed. methylsiloxane-fibrin composite scaffolds for controlled delivery of pro- Res. Int. 2015 (2015), 968927. . angiogenic growth factors, Biomaterials 31 (2010) 5336–5344. [146] K. Nakamatsu, Y. Nishimura, M. Suzuki, S. Kanamori, O. Maenishi, Y. Yasuda, [122] F. Cianfarani, R. Tommasi, C.M. Failla, M.T. Viviano, G. Annessi, M. Papi, G. Erythropoietin/erythropoietin-receptor system as an angiogenic factor in chemi- Zambruno, T. Odorisio, Granulocyte/macrophage colony-stimulating factor treat- cally induced murine hepatic tumors, Int. J. Clin. Oncol. 9 (2004) 184–188. ment of human chronic ulcers promotes angiogenesis associated with de novo [147] T. Okazaki, S. Ebihara, M. Asada, S. Yamanda, K. Niu, H. Arai, Erythropoietin pro- vascular endothelial growth factor transcription in the ulcer bed, Br. J. Dermatol. motes the growth of tumors lacking its receptor and decreases survival of 154 (2006) 34–41. tumor-bearing mice by enhancing angiogenesis, Neoplasia (New York, N.Y.) 10 [123] R.D. Galiano, O.M. Tepper, C.R. Pelo, K.A. Bhatt, M. Callaghan, N. Bastidas, S. (2008) 932–939. Bunting, H.G. Steinmetz, G.C. Gurtner, Topical vascular endothelial growth factor [148] J. Deshane, S. Chen, S. Caballero, A. Grochot-Przeczek, H. Was, S. Li Calzi, R. accelerates diabetic wound healing through increased angiogenesis and by mobi- Lach, T.D. Hock, B. Chen, N. Hill-Kapturczak, G.P. Siegal, J. Dulak, A. lizing and recruiting bone marrow-derived cells, Am. J. Pathol. 164 (2004) Jozkowicz, M.B. Grant, A. Agarwal, Stromal cell-derived factor 1 promotes an- 1935–1947. giogenesis via a heme oxygenase 1-dependent mechanism, J. Exp. Med. 204 [124] H.K. Makadia, S.J. Siegel, Poly lactic-co-glycolic acid (PLGA) as biodegradable con- (2007) 605–618. trolled drug delivery carrier, Polymers (Basel) 3 (2011) 1377–1397. [149] Y. Zhu, R. Hoshi, S. Chen, J. Yi, C. Duan, R.D. Galiano, H.F. Zhang, G.A. Ameer, [125] K.K. Chereddy, A. Lopes, S. Koussoroplis, V. Payen, C. Moia, H. Zhu, P. Sonveaux, P. Sustained release of stromal cell derived factor-1 from an antioxidant Carmeliet, A. des Rieux, G. Vandermeulen, V. Preat, Combined effects of PLGA thermoresponsive hydrogel enhances dermal wound healing in diabetes, J Control and vascular endothelial growth factor promote the healing of non-diabetic and di- Release 238 (2016) 114–122. abetic wounds, Nanomedicine 11 (2015) 1975–1984. [150] M.A. Olekson, R. Faulknor, A. Bandekar, M. Sempkowski, H.C. Hsia, F. Berthiaume, [126] J. Huang, J. Ren, G. Chen, Z. Li, Y. Liu, G. Wang, X. Wu, Tunable sequential drug de- SDF-1 liposomes promote sustained cell proliferation in mouse diabetic wounds, livery system based on chitosan/hyaluronic acid hydrogels and PLGA microspheres Wound Repair Regenerat 23 (2015) 711–723. for management of non-healing infected wounds, Mater. Sci. Eng. C Mater. Biol. [151] B.L. Hoh, K. Hosaka, D.P. Downes, K.W. Nowicki, E.N. Wilmer, G.J. Velat, E.W. Scott, Appl. 89 (2018) 213–222. Stromal cell-derived factor-1 promoted angiogenesis and inflammatory cell infil- [127] P. Losi, E. Briganti, C. Errico, A. Lisella, E. Sanguinetti, F. Chiellini, G. Soldani, Fibrin- tration in aneurysm walls, J. Neurosurg. 120 (2014) 73–86. based scaffold incorporating VEGF- and bFGF-loaded nanoparticles stimulates [152] J.R. Krieger, M.E. Ogle, J. McFaline-Figueroa, C.E. Segar, J.S. Temenoff, E.A. Botchwey, wound healing in diabetic mice, Acta Biomater. 9 (2013) 7814–7821. Spatially localized recruitment of anti-inflammatory monocytes by SDF-1alpha- [128] Y.M. Elcin, V. Dixit, G. Gitnick, Extensive in vivo angiogenesis following controlled releasing hydrogels enhances microvascular network remodeling, Biomaterials release of human vascular endothelial cell growth factor: implications for tissue 77 (2016) 280–290. engineering and wound healing, Artif. Organs 25 (2001) 558–565. [153] T. Tang, H. Jiang, Y. Yu, F. He, S.Z. Ji, Y.Y. Liu, Z.S. Wang, S.C. Xiao, C. Tang, G.Y. Wang, [129] A. Tellechea, E.A. Silva, J. Min, E.C. Leal, M.E. Auster, L. Pradhan-Nabzdyk, W. Shih, Z.F. Xia, A new method of wound treatment: targeted therapy of skin wounds with D.J. Mooney, A. Veves, Alginate and DNA gels are suitable delivery systems for di- reactive oxygen species-responsive nanoparticles containing SDF-1alpha, Int. J. abetic wound healing, Int J Low Extrem Wounds 14 (2015) 146–153. Nanomedicine 10 (2015) 6571–6585. [130] T. Odorisio, F. Cianfarani, C.M. Failla, G. Zambruno, The placenta growth factor in [154] Y. Li, S. Chang, W. Li, G. Tang, Y. Ma, Y. Liu, F. Yuan, Z. Zhang, G.Y. Yang, Y. Wang, skin angiogenesis, J. Dermatol. Sci. 41 (2006) 11–19. cxcl12-engineered endothelial progenitor cells enhance neurogenesis and angio- [131] F. Cianfarani, G. Zambruno, L. Brogelli, F. Sera, P.M. Lacal, M. Pesce, M.C. Capogrossi, genesis after ischemic brain injury in mice, Stem Cell Res Ther 9 (1) (2018) C.M. Failla, M. Napolitano, T. Odorisio, Placenta growth factor in diabetic wound 139–153. healing, altered expression and therapeutic potential, Am J Pathol 169 (2006) [155] E. Vagesjo, E. Ohnstedt, A. Mortier, H. Lofton, F. Huss, P. Proost, S. Roos, M. 1167–1182. Phillipson, Accelerated wound healing in mice by on-site production and delivery [132] B. Hendrickx, K. Verdonck, S. Van den Berge, S. Dickens, E. Eriksson, J.J. Vranckx, A. of CXCL12 by transformed lactic acid bacteria, Proc. Natl. Acad. Sci. U. S. A. 115 Luttun, Integration of blood outgrowth endothelial cells in dermal fibroblast sheets (2018) 1895–1900. promotes full thickness wound healing, Stem Cells 28 (2010) 1165–1177. [156] M.M. Moisenovich, N.V. Malyuchenko, A.Y. Arkhipova, M.S. Kotlyarova, L.I. [133] T. Odorisio, C. Schietroma, M.L. Zaccaria, F. Cianfarani, C. Tiveron, L. Tatangelo, C.M. Davydova, A.V. Goncharenko, O.I. Agapova, M.S. Drutskaya, V.G. Bogush, I.I. Failla, G. Zambruno, Mice overexpressing placenta growth factor exhibit increased Agapov, V.G. Debabov, M.P. Kirpichnikov, Novel 3D-microcarriers from recombi- vascularization and vessel permeability, J. Cell Sci. 115 (2002) 2559–2567. nant spidroin for regenerative medicine, Dokl. Biochem. Biophys. 463 (2015) [134] S.J. Park, K.J. Kim, W.U. Kim, C.S. Cho, Interaction of mesenchymal stem cells with 232–235. fibroblast-like synoviocytes via cadherin-11 promotes angiogenesis by enhanced [157] D. Ti, H. Hao, L. Xia, C. Tong, J. Liu, L. Dong, S. Xu, Y. Zhao, H. Liu, X. Fu, W. Han, Con- secretion of , J. Immunol. 192 (2014) 3003–3010. trolled release of thymosin beta 4 using a collagen-chitosan sponge scaffold aug- [135] M.M. Martino, P.S. Briquez, E. Guc, F. Tortelli, W.W. Kilarski, S. Metzger, J.J. Rice, G.A. ments cutaneous wound healing and increases angiogenesis in diabetic rats with Kuhn, R. Muller, M.A. Swartz, J.A. Hubbell, Growth factors engineered for super- hindlimb ischemia, Tissue Eng. A 21 (2015) 541–549. affinity to the extracellular matrix enhance tissue healing, Science 343 (2014) [158] R. Kikuchi, K. Nakamura, S. MacLauchlan, D.T. Ngo, I. Shimizu, J.J. Fuster, Y. 885–888. Katanasaka, S. Yoshida, Y. Qiu, T.P. Yamaguchi, T. Matsushita, T. Murohara, N. [136] M. Hrynyk, M. Martins-Green, A.E. Barron, R.J. Neufeld, Sustained prolonged topical Gokce, D.O. Bates, N.M. Hamburg, K. Walsh, An antiangiogenic isoform of VEGF- delivery of bioactive human insulin for potential treatment of cutaneous wounds, A contributes to impaired vascularization in peripheral artery disease, Nat. Med. Int. J. Pharm. 398 (2010) 146–154. 20 (2014) 1464–1471. 122 A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125

[159] R. Brooks, R. Williamson, M. Bass, Syndecan-4 Independently Regulates Multiple in treating methicillin-resistant Staphylococcus aureus burn wound infections, ACS Small GTPases to Promote Fibroblast Migration During Wound Healing, Small Infectious Dis 3 (2017) 820–832. GTPases, United States, 2012 73–79. [186] S.T. Grambart, Sports medicine and platelet-rich plasma: nonsurgical therapy, Clin [160] A. Monteforte, B. Lam, M.B. Sherman, K. Henderson, A.D. Sligar, A. Spencer, B. Tang, Podiat Med Surg 32 (2015) 99–107. A.K. Dunn, A.B. Baker, Glioblastoma exosomes for therapeutic angiogenesis in pe- [187] J.P. Frechette, I. Martineau, G. Gagnon, Platelet-rich plasmas: growth factor content ripheral ischemia, Tissue Eng. A 23 (2017) 1251–1261. and roles in wound healing, J. Dent. Res. 84 (2005) 434–439. [161] N.M. Kofler, M. Simons, Angiogenesis versus Arteriogenesis: Neuropilin 1 Modula- [188] K.M. Lacci, A. Dardik, Platelet-rich plasma: support for its use in wound healing, tion of VEGF Signaling, 2015. Yale J Biol Med 83 (2010) 1–9. [162] F. Mac Gabhann, A.S. Popel, Targeting neuropilin-1 to inhibit VEGF signaling in can- [189] M.J. Martinez-Zapata, A.J. Marti-Carvajal, I. Sola, J.A. Exposito, I. Bolibar, L. cer: Comparison of therapeutic approaches, PLoS Comput. Biol. 2 (2006), e180. . Rodriguez, J. Garcia, C. Zaror, Autologous platelet-rich plasma for treating chronic [163] I. Zachary, Neuropilins: role in signalling, angiogenesis and disease, Chemical im- wounds, Cochrane Database Syst. Rev. (5) (2016), CD006899. . munology and allergy 99 (2014) 37–70. [190] . Y.R. Jeon, E.H. Kang, C.E. Yang, I.S. Yun, W.J. Lee, D.H. Lew, The effect of platelet- [164] P. Dutta, S. Das, Mammalian antimicrobial peptides: Promising therapeutic targets rich plasma on composite graft survival, Plast. Reconstr. Surg. 134 (2014) against infection and chronic inflammation, Curr. Top. Med. Chem. 16 (2016) 239–246. 99–129. [191] M. Qiu, D. Chen, C. Shen, J. Shen, H. Zhao, Y. He, Platelet-rich plasma-loaded poly(d, [165] A. Pfalzgraff, K. Brandenburg, G. Weindl, Antimicrobial peptides and their thera- l-lactide)-poly(ethylene glycol)-poly(d,l-lactide) hydrogel dressing promotes full- peutic potential for bacterial skin infections and wounds, Front. Pharmacol. 9 thickness skin wound healing in a rodent model, Int. J. Mol. Sci. 17 (2016). (2018) 281. [192] I. Andia, E. Rubio-Azpeitia, N. Maffulli, Platelet-rich plasma modulates the secretion [166] B. Bommarius, H. Jenssen, M. Elliott, J. Kindrachuk, M. Pasupuleti, H. Gieren, K.E. of inflammatory/angiogenic proteins by inflamed tenocytes, Clin Orthopaed Re- Jaeger, R.E. Hancock, D. Kalman, Cost-effective expression and purification of anti- lated Res (2015) 1624–1634. microbial and host defense peptides in Escherichia coli, Peptides 31 (2010) [193] . J. Leotot, L. Coquelin, G. Bodivit, P. Bierling, P. Hernigou, H. Rouard, N. 1957–1965. Chevallier, Platelet lysate coating on scaffolds directly and indirectly enhances [167] K.K. Chereddy, C.H. Her, M. Comune, C. Moia, A. Lopes, P.E. Porporato, J. Vanacker, cell migration, improving bone and blood vessel formation, Acta Biomater. 9 M.C. Lam, L. Steinstraesser, P. Sonveaux, H. Zhu, L.S. Ferreira, G. Vandermeulen, V. (2013) 6630–6640. Preat, PLGA nanoparticles loaded with host defense peptide LL37 promote [194] P. Xie, S. Jia, R. Tye, C. Chavez-Munoz, M. Vracar-Grabar, S.J. Hong, R. Galiano, T.A. wound healing, J Control Release 194 (2014) 138–147. Mustoe, Systemic administration of hemoglobin improves ischemic wound [168] I. Garcia-Orue, G. Gainza, C. Girbau, R. Alonso, J.J. Aguirre, J.L. Pedraz, M. Igartua, healing, J. Surg. Res. 194 (2015) 696–705. R.M. Hernandez, LL37 loaded nanostructured lipid carriers (NLC): A new strategy [195] A.T. Kawaguchi, Y. Okamoto, Y. Kise, S. Takekoshi, C. Murayama, H. Makuuchi, Ef- for the topical treatment of chronic wounds, Eur J Pharm Biopharm 108 (2016) fects of liposome-encapsulated hemoglobin on gastric wound healing in the rat, 310–316. Artif. Organs 38 (2014) 641–649. [169] G.H. Gudmundsson, B. Agerberth, J. Odeberg, T. Bergman, B. Olsson, R. Salcedo, The [196] J. Banerjee, Y.C. Chan, C.K. Sen, MicroRNAs in skin and wound healing, Physiol. Ge- human gene FALL39 and processing of the cathelin precursor to the antibacterial nomics 43 (2011) 543–556. peptide LL-37 in granulocytes, Eur. J. Biochem. 238 (1996) 325–332. [197] D.P. Bartel, MicroRNAs: genomics, biogenesis, mechanism, and function, Cell 116 [170] M. Kittaka, H. Shiba, M. Kajiya, K. Ouhara, K. Takeda, K. Kanbara, T. Fujita, H. (2004) 281–297. Kawaguchi, H. Komatsuzawa, H. Kurihara, Antimicrobial peptide LL37 promotes [198] E.J. Mulholland, N. Dunne, H.O. McCarthy, MicroRNA as therapeutic targets for vascular endothelial growth factor-A expression in human periodontal chronic wound healing, Mol Ther Nucleic Acids 8 (2017) 46–55. cells, J. Periodontal Res. 48 (2013) 228–234. [199] K. Deonarine, M.C. Panelli, M.E. Stashower, P. Jin, K. Smith, H.B. Slade, C. Norwood, [171] D. Vandamme, B. Landuyt, W. Luyten, L. Schoofs, A comprehensive summary of E. Wang, F.M. Marincola, D.F. Stroncek, Gene expression profiling of cutaneous LL-37, the factotum human cathelicidin peptide, Cell. Immunol. 280 (2012) wound healing, J. Transl. Med. 5 (2007) 11. 22–35. [200] J.M. Wang, J. Tao, D.D. Chen, J.J. Cai, K. Irani, Q. Wang, H. Yuan, A.F. Chen, MicroRNA [172] R. Ramos, J.P. Silva, A.C. Rodrigues, R. Costa, L. Guardao, F. Schmitt, R. Soares, M. miR-27b rescues bone marrow-derived angiogenic cell function and accelerates Vilanova, L. Domingues, M. Gama, Wound healing activity of the human antimicro- wound healing in type 2 diabetes mellitus, Arterioscler. Thromb. Vasc. Biol. 34 bial peptide LL37, Peptides 32 (2011) 1469–1476. (2014) 99–109. [173] G. Sheehan, G. Bergsson, N.G. McElvaney, E.P. Reeves, K. Kavanagh, The human [201] M.H. Lu, C.J. Hu, L. Chen, X. Peng, J. Chen, J.Y. Hu, M. Teng, G.P. Liang, miR-27b re- cathelicidin antimicrobial peptide LL-37 promotes the growth of the pulmonary presses migration of mouse MSCs to burned margins and prolongs wound repair pathogen Aspergillus fumigatus, Infect. Immun. 86 (2018) e00097-18. through silencing SDF-1a, PLoS One 8 (2013), e68972. . [174] K.B. Christopher, Chapter 117 - Vitamin D and critically ill intensive care unit pa- [202] Y. Zhu, Z. Li, Y. Wang, L. Li, D. Wang, W. Zhang, L. Liu, H. Jiang, J. Yang, J. Cheng, tients, in: D. Feldman (Ed.), Vitamin D, Fourth EditionAcademic Press 2018, Overexpression of miR-29b reduces collagen biosynthesis by inhibiting heat pp. 1177–1194. shock protein 47 during skin wound healing, Transl Res 178 (2016) 38–53 e36. [175] X.B. Wei, R.J. Wu, D.Y. Si, X.D. Liao, L.L. Zhang, R.J. Zhang, Novel hybrid peptide [203] J.K. Park, H. Peng, W. Yang, J. Katsnelson, O. Volpert, R.M. Lavker, miR-184 exhibits cecropin A (1–8)-LL37 (17–30) with potential antibacterial activity, Int J Mol Sci angiostatic properties via regulation of Akt and VEGF signaling pathways, FASEB J 17 (7) (2016) 983–994. 31 (2017) 256–265. [176] M. Yang, M. Tang, X. Ma, L. Yang, J. He, X. Peng, G. Guo, L. Zhou, N. Luo, Z. Yuan, A. [204] H.H. Kwok, L.S. Chan, P.Y. Poon, P.Y. Yue, R.N. Wong, Ginsenoside-Rg1 induces an- Tong, AP-57/C10orf99 is a new type of multifunctional antimicrobial peptide, giogenesis by the inverse regulation of MET receptor expression Biochem. Biophys. Res. Commun. 457 (2015) 347–352. through miR-23a, Toxicol. Appl. Pharmacol. 287 (2015) 276–283. [177] X. Li, R. Fan, A. Tong, M. Yang, J. Deng, L. Zhou, X. Zhang, G. Guo, In situ gel-forming [205] S. Dangwal, B. Stratmann, C. Bang, J.M. Lorenzen, R. Kumarswamy, J. Fiedler, C.S. AP-57 peptide delivery system for cutaneous wound healing, Int. J. Pharm. 495 Falk, C.J. Scholz, T. Thum, D. Tschoepe, Impairment of wound healing in patients (2015) 560–571. with type 2 diabetes mellitus influences circulating microRNA patterns via inflam- [178] J.P. Silva, S. Dhall, M. Garcia, A. Chan, C. Costa, M. Gama, M. Martins-Green, Im- matory cytokines, Arterioscler. Thromb. Vasc. Biol. 35 (2015) 1480–1488. proved burn wound healing by the antimicrobial peptide LLKKK18 released from [206] T. Wang, Y. Feng, H. Sun, L. Zhang, L. Hao, C. Shi, J. Wang, R. Li, X. Ran, Y. Su, Z. Zou, conjugates with dextrin embedded in a carbopol gel, Acta Biomater. 26 (2015) miR-21 regulates skin wound healing by targeting multiple aspects of the healing 249–262. process, Am. J. Pathol. 181 (2012) 1911–1920. [179] I. Nagaoka, S. Hirota, F. Niyonsaba, M. Hirata, Y. Adachi, H. Tamura, S. Tanaka, D. [207] O. Tornavaca, M. Chia, N. Dufton, L.O. Almagro, D.E. Conway, A.M. Randi, M.A. Heumann, Augmentation of the lipopolysaccharide-neutralizing activities of Schwartz, K. Matter, M.S. Balda, ZO-1 controls endothelial adherens junctions, human cathelicidin CAP18/LL-37-derived antimicrobial peptides by replacement cell-cell tension, angiogenesis, and barrier formation, J. Cell Biol. 208 (2015) with hydrophobic and cationic amino acid residues, Clin. Diagn. Lab. Immunol. 9 821–838. (2002) 972–982. [208] Y.C. Chan, S. Roy, S. Khanna, C.K. Sen, Downregulation of endothelial microRNA- [180] Y. Wang, Z. Chen, G. Luo, W. He, K. Xu, R. Xu, Q. Lei, J. Tan, J. Wu, M. Xing, In-situ- 200b supports cutaneous wound angiogenesis by desilencing GATA binding pro- generated vasoactive intestinal peptide loaded microspheres in mussel-inspired tein 2 and vascular endothelial growth factor receptor 2, Arterioscler. Thromb. polycaprolactone nanosheets creating spatiotemporal releasing microenvironment Vasc. Biol. 32 (2012) 1372–1382. to promote wound healing and angiogenesis, ACS Appl. Mater. Interfaces 8 (2016) [209] A. Bitto, L. Minutoli, M.R. Galeano, D. Altavilla, F. Polito, T. Fiumara, M. Calo, P. Lo 7411–7421. Cascio, L. Zentilin, M. Giacca, F. Squadrito, Angiopoietin-1 gene transfer improves [181] J. Yang, Q.D. Shi, T.B. Song, G.F. Feng, W.J. Zang, C.H. Zong, L. Chang, Vasoactive in- impaired wound healing in genetically diabetic mice without increasing VEGF ex- testinal peptide increases VEGF expression to promote proliferation of brain vascu- pression, Clin. Sci. (Lond.) 114 (2008) 707–718. lar endothelial cells via the cAMP/PKA pathway after ischemic insult in vitro, [210] J. Krutzfeldt, N. Rajewsky, R. Braich, K.G. Rajeev, T. Tuschl, M. Manoharan, M. Peptides 42 (2013) 105–111. Stoffel, Silencing of microRNAs in vivo with 'antagomirs', Nature 438 (2005) [182] H. Liu, Z. Duan, J. Tang, Q. Lv, M. Rong, R. Lai, A short peptide from frog skin accel- 685–689. erates diabetic wound healing, FEBS J. 281 (2014) 4633–4643. [211] J. Stenvang, A. Petri, M. Lindow, S. Obad, S. Kauppinen, Inhibition of microRNA [183] L. Mu, J. Tang, H. Liu, C. Shen, M. Rong, Z. Zhang, R. Lai, A potential wound-healing- function by antimiR oligonucleotides, Silence 3 (1) (2012). promoting peptide from salamander skin, FASEB J (2014) 3919–3929. [212] T. Lucas, F. Schafer, P. Muller, S.A. Eming, A. Heckel, S. Dimmeler, Light-inducible [184] H. Tomioka, H. Nakagami, A. Tenma, Y. Saito, T. Kaga, T. Kanamori, N. Tamura, antimiR-92a as a therapeutic strategy to promote skin repair in healing-impaired K. Tomono, Y. Kaneda, R. Morishita, Novel anti-microbial peptide SR-0379 ac- diabetic mice, Nat. Commun. 8 (2017), 15162. . celerates wound healing via the PI3 kinase/Akt/mTOR pathway, PLoS One 9 [213] A. Bonauer, G. Carmona, M. Iwasaki, M. Mione, M. Koyanagi, A. Fischer, J. (2014), e92597. . Burchfield, H. Fox, C. Doebele, K. Ohtani, E. Chavakis, M. Potente, M. Tjwa, C. [185] Z. Ma, J. Han, B. Chang, L. Gao, Z. Lu, F. Lu, H. Zhao, C. Zhang, X. Bie, Membrane-ative Urbich, A.M. Zeiher, S. Dimmeler, MicroRNA-92a controls angiogenesis and func- amphipathic peptide WRL3 with in vitro antibiofilm capability and in vivo efficacy tional recovery of ischemic tissues in mice, Science 324 (2009) 1710–1713. A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125 123

[214] B. Icli, C.S. Nabzdyk, J. Lujan-Hernandez, M. Cahill, M.E. Auster, A.K. Wara, X. Sun, D. [241] J. Ko, H. Jun, H. Chung, C. Yoon, T. Kim, M. Kwon, S. Lee, S. Jung, M. Kim, J.H. Park, Ozdemir, G. Giatsidis, D.P. Orgill, M.W. Feinberg, Regulation of impaired angiogen- Comparison of EGF with VEGF non-viral gene therapy for cutaneous wound esis in diabetic dermal wound healing by microRNA-26a, J. Mol. Cell. Cardiol. 91 healing of streptozotocin diabetic mice, Diabetes Metab. J. 35 (2011) 226–235. (2016) 151–159. [242] M.A. Fonder, G.S. Lazarus, D.A. Cowan, B. Aronson-Cook, A.R. Kohli, A.J. Mamelak, [215] M. Sinha, S. Ghatak, S. Roy, C.K. Sen, microRNA-200b as a Switch for Inducible Adult Treating the chronic wound: A practical approach to the care of nonhealing Angiogenesis, Antioxid. Redox Signal. 22 (2015) 1257–1272. wounds and wound care dressings, J. Am. Acad. Dermatol. 58 (2008) 185–206. [216] G. Pizzino, N. Irrera, F. Galfo, G. Pallio, F. Mannino, A. D'Amore, E. Pellegrino, A. Ieni, [243] T.A. Mustoe, K. O'Shaughnessy, O. Kloeters, Chronic wound pathogenesis and cur- G.T. Russo, M. Calapai, D. Altavilla, F. Squadrito, A. Bitto, Effects of the antagomiRs rent treatment strategies: a unifying hypothesis, Plast. Reconstr. Surg. 117 (2006) 15b and 200b on the altered healing pattern of diabetic mice, Br. J. Pharmacol. 35S–41S. 175 (2018) 644–655. [244] S.M. Bauer, R.J. Bauer, O.C. Velazquez, Angiogenesis, vasculogenesis, and induction [217] S. Daya, K.I. Berns, Gene therapy using adeno-associated virus vectors, Clin. of healing in chronic wounds, Vasc. Endovasc. Surg. 39 (2005) 293–306. Microbiol. Rev. 21 (2008) 583–593. [245] K.G. Harding, H.L. Morris, G.K. Patel, Science, medicine and the future: healing [218] T. Friedmann, Progress toward human gene therapy, Science 244 (1989) chronic wounds, BMJ 324 (2002) 160–163. 1275–1281. [246] J. Asai, H. Takenaka, S. Hirakawa, J. Sakabe, A. Hagura, S. Kishimoto, K. Maruyama, [219] M. Galeano, B. Deodato, D. Altavilla, D. Cucinotta, N. Arsic, H. Marini, V. Torre, M. K. Kajiya, S. Kinoshita, Y. Tokura, N. Katoh, Topical simvastatin accelerates wound Giacca, F. Squadrito, Adeno-associated viral vector-mediated human vascular en- healing in diabetes by enhancing angiogenesis and lymphangiogenesis, Am. J. dothelial growth factor gene transfer stimulates angiogenesis and wound healing Pathol. 181 (2012) 2217–2224. in the genetically diabetic mouse, Diabetologia 46 (2003) 546–555. [247] Z. Hou, C. Nie, Z. Si, Y. Ma, Deferoxamine enhances neovascularization and acceler- [220] H. Kampfer, J. Pfeilschifter, S. Frank, Expressional regulation of angiopoietin-1 and ates wound healing in diabetic rats via the accumulation of hypoxia-inducible -2 and the tie-1 and -2 receptor tyrosine kinases during cutaneous wound healing: factor-1alpha, Diabetes Res. Clin. Pract. 101 (2013) 62–71. a comparative study of normal and impaired repair, Lab. Investig. 81 (2001) [248] J.G. Merrell, S.W. McLaughlin, L. Tie, C.T. Laurencin, A.F. Chen, L.S. Nair, Curcumin- 361–373. loaded poly(epsilon-caprolactone) nanofibres: diabetic wound dressing with anti- [221] T. Namba, H. Koike, K. Murakami, M. Aoki, H. Makino, N. Hashiya, T. Ogihara, Y. oxidant and anti-inflammatory properties, Clin. Exp. Pharmacol. Physiol. 36 (2009) Kaneda, M. Kohno, R. Morishita, Angiogenesis induced by endothelial nitric oxide 1149–1156. synthase gene through vascular endothelial growth factor expression in a rat [249] Y. Wang, G. Han, B. Guo, J. Huang, Hyaluronan oligosaccharides promote diabetic hindlimb ischemia model, Circulation 108 (2003) 2250–2257. wound healing by increasing angiogenesis, Pharmacol. Rep. 68 (2016) 1126–1132. [222] M.R. Schaffer, U. Tantry, P.A. Efron, G.M. Ahrendt, F.J. Thornton, A. Barbul, Diabetes- [250] G. Han, L.N. Nguyen, C. Macherla, Y. Chi, J.M. Friedman, J.D. Nosanchuk, L.R. impaired healing and reduced wound nitric oxide synthesis: a possible pathophys- Martinez, Nitric oxide-releasing nanoparticles accelerate wound healing by pro- iologic correlation, Surgery 121 (1997) 513–519. moting fibroblast migration and collagen deposition, Am. J. Pathol. 180 (2012) [223] A.M. Breen, P. Dockery, T. O'Brien, A.S. Pandit, The use of therapeutic gene eNOS de- 1465–1473. livered via a fibrin scaffold enhances wound healing in a compromised wound [251] B.B. Childress, J.K. Stechmiller, Role of nitric oxide in wound healing, Biol Res Nurs model, Biomaterials 29 (2008) 3143–3151. 4(2002)5–15. [224] A. Breen, P. Dockery, T. O'Brien, A. Pandit, Fibrin scaffold promotes adenoviral gene [252] A. Schwentker, Y. Vodovotz, R. Weller, T.R. Billiar, Nitric oxide and wound repair: transfer and controlled vector delivery, J. Biomed. Mater. Res. A 89 (2009) role of cytokines? Nitric Oxide 7 (2002) 1–10. 876–884. [253] M.B. Witte, A. Barbul, Role of nitric oxide in wound repair, Am. J. Surg. 183 (2002) [225] M. Kulkarni, A. O'Loughlin, R. Vazquez, K. Mashayekhi, P. Rooney, U. Greiser, E. 406–412. O'Toole, T. O'Brien, M.M. Malagon, A. Pandit, Use of a fibrin-based system for en- [254] J.D. Luo, A.F. Chen, Nitric oxide: a newly discovered function on wound healing, hancing angiogenesis and modulating inflammation in the treatment of hypergly- Acta Pharmacol. Sin. 26 (2005) 259–264. cemic wounds, Biomaterials 35 (2014) 2001–2010. [255] S. Chien, Intracellular ATP delivery using highly fusogenic liposomes, Methods Mol. [226] B.C. Dash, D. Thomas, M. Monaghan, O. Carroll, X. Chen, K. Woodhouse, T. O'Brien, Biol. 605 (2010) 377–391. A. Pandit, An injectable elastin-based gene delivery platform for dose-dependent [256] B. Sangiuliano, N.M. Perez, D.F. Moreira, J.E. Belizario, Cell death-associated modulation of angiogenesis and inflammation for critical limb ischemia, Biomate- molecular-pattern molecules: inflammatory signaling and control, Mediat. rials 65 (2015) 126–139. Inflamm. (2014) (2014) 821043. [227] K.W. Moore, R. de Waal Malefyt, R.L. Coffman, A. O'Garra, -10 and the [257] B. Chiang, E. Essick, W. Ehringer, S. Murphree, M.A. Hauck, M. Li, S. Chien, Enhanc- interleukin-10 receptor, Annu. Rev. Immunol. 19 (2001) 683–765. ing skin wound healing by direct delivery of intracellular adenosine triphosphate, [228] S. Kobsa, N.J. Kristofik, A.J. Sawyer, A.L. Bothwell, T.R. Kyriakides, W.M. Saltzman, Am. J. Surg. 193 (2007) 213–218. An electrospun scaffold integrating nucleic acid delivery for treatment of full- [258] J. Wang, Q. Zhang, R. Wan, Y. Mo, M. Li, M.T. Tseng, S. Chien, Intracellular adenosine thickness wounds, Biomaterials 34 (2013) 3891–3901. triphosphate delivery enhanced skin wound healing in rabbits, Ann. Plast. Surg. 62 [229] R. Guo, S. Xu, L. Ma, A. Huang, C. Gao, Enhanced angiogenesis of gene-activated (2009) 180–186. dermal equivalent for treatment of full thickness incisional wounds in a porcine [259] J. Wang, R. Wan, Y. Mo, M. Li, Q. Zhang, S. Chien, Intracellular delivery of adenosine model, Biomaterials 31 (2010) 7308–7320. triphosphate enhanced healing process in full-thickness skin wounds in diabetic [230] R. Guo, S. Xu, L. Ma, A. Huang, C. Gao, The healing of full-thickness burns treated by rabbits, Am. J. Surg. 199 (2010) 823–832. using plasmid DNA encoding VEGF-165 activated collagen-chitosan dermal equiv- [260] M. Khaidakov, W. Wang, J.A. Khan, B.Y. Kang, P.L. Hermonat, J.L. Mehta, Statins and alents, Biomaterials 32 (2011) 1019–1031. angiogenesis: is it about connections? Biochem. Biophys. Res. Commun. 387 [231] M. Thiersch, M. Rimann, V. Panagiotopoulou, E. Ozturk, T. Biedermann, M. Textor, (2009) 543–547. T.C. Luhmann, H. Hall, The angiogenic response to PLL-g-PEG-mediated HIF- [261] Y. Kureishi, Z. Luo, I. Shiojima, A. Bialik, D. Fulton, D.J. Lefer, W.C. Sessa, K. Walsh, 1alpha plasmid DNA delivery in healthy and diabetic rats, Biomaterials 34 The HMG-CoA reductase inhibitor simvastatin activates the protein kinase Akt (2013) 4173–4182. and promotes angiogenesis in normocholesterolemic animals, Nat. Med. 6 [232] D. Trentin, H. Hall, S. Wechsler, J.A. Hubbell, Peptide-matrix-mediated gene trans- (2000) 1004–1010. fer of an oxygen-insensitive hypoxia-inducible factor-1alpha variant for local in- [262] A. Bitto, L. Minutoli, D. Altavilla, F. Polito, T. Fiumara, H. Marini, M. Galeano, M. Calo, duction of angiogenesis, Proc. Natl. Acad. Sci. U. S. A. 103 (2006) 2506–2511. P. Lo Cascio, M. Bonaiuto, A. Migliorato, A.P. Caputi, F. Squadrito, Simvastatin en- [233] G.L. Semenza, Oxygen sensing, hypoxia-inducible factors, and disease pathophysi- hances VEGF production and ameliorates impaired wound healing in experimental ology, Annu. Rev. Pathol. 9 (2014) 47–71. diabetes, Pharmacol. Res. 57 (2008) 159–169. [234] M. Tomizawa, F. Shinozaki, Y. Motoyoshi, T. Sugiyama, S. Yamamoto, M. Sueishi, [263] M. Rezvanian, C.K. Tan, S.F. Ng, Simvastatin-loaded lyophilized wafers as a poten- Sonoporation: Gene transfer using ultrasound, World J Methodol 3 (2013) 39–44. tial dressing for chronic wounds, Drug Dev. Ind. Pharm. 42 (2016) 2055–2062. [235] W.G. Marshall Jr., B.A. Boone, J.D. Burgos, S.I. Gografe, M.K. Baldwin, M.L. Danielson, [264] G. Dabiri, E. Damstetter, T. Phillips, Choosing a wound dressing based on common M.J. Larson, D.R. Caretto, Y. Cruz, B. Ferraro, L.C. Heller, K.E. Ugen, M.J. Jaroszeski, R. wound characteristics, Adv Wound Care (New Rochelle) 5 (2016) 32–41. Heller, Electroporation-mediated delivery of a naked DNA plasmid expressing [265] S.F. Ng, N. Jumaat, Carboxymethyl cellulose wafers containing antimicrobials: a VEGF to the porcine heart enhances protein expression, Gene Ther. 17 (2010) modern drug delivery system for wound infections, Eur. J. Pharm. Sci. 51 (2014) 419–423. 173–179. [236] J. Gehl, Electroporation: theory and methods, perspectives for drug delivery, gene [266] S. Yasasvini, R.S. Anusa, B.N. VedhaHari, P.C. Prabhu, D. RamyaDevi, Topical hydro- therapy and research, Acta Physiol. Scand. 177 (2003) 437–447. gel matrix loaded with Simvastatin microparticles for enhanced wound healing ac- [237] P.Y. Lee, S. Chesnoy, L. Huang, Electroporatic delivery of TGF-beta1 gene works syn- tivity, Mater. Sci. Eng. C Mater. Biol. Appl. 72 (2017) 160–167. ergistically with electric therapy to enhance diabetic wound healing in db/db mice, [267] R. Cheng, J.X. Ma, Angiogenesis in diabetes and obesity, Rev. Endocr. Metab. Disord. J Invest Dermatol 123 (2004) 791–798. 16 (2015) 67–75. [238] B.W. Bigger, O. Tolmachov, J.M. Collombet, M. Fragkos, I. Palaszewski, C. Coutelle, [268] S.K. Kota, L.K. Meher, S. Jammula, S.K. Kota, S.V. Krishna, K.D. Modi, Aberrant angio- An araC-controlled bacterial cre expression system to produce DNA minicircle genesis: The gateway to diabetic complications, Indian J Endocrinol Metab 16 vectors for nuclear and mitochondrial gene therapy, J. Biol. Chem. 276 (2001) (2012) 918–930. 23018–23027. [269] W. Li, Y. Li, S. Guan, J. Fan, C.F. Cheng, A.M. Bright, C. Chinn, M. Chen, D.T. Woodley, [239] P. Mayrhofer, M. Blaesen, M. Schleef, W. Jechlinger, Minicircle-DNA production by Extracellular heat shock protein-90alpha: linking hypoxia to skin cell motility and site specific recombination and protein-DNA interaction chromatography, The wound healing, EMBO J. 26 (2007) 1221–1233. journal of gene medicine 10 (2008) 1253–1269. [270] M. Ram, V. Singh, S. Kumawat, D. Kumar, M.C. Lingaraju, T. Uttam Singh, A. Rahal, [240] C.S. Yoon, H.S. Jung, M.J. Kwon, S.H. Lee, C.W. Kim, M.K. Kim, M. Lee, J.H. Park, S.K. Tandan, D. Kumar, Deferoxamine modulates cytokines and growth factors to Sonoporation of the minicircle-VEGF(165) for wound healing of diabetic mice, accelerate cutaneous wound healing in diabetic rats, Eur. J. Pharmacol. 764 Pharm. Res. 26 (2009) 794–801. (2015) 9–21. 124 A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125

[271] D. Duscher, E. Neofytou, V.W. Wong, Z.N. Maan, R.C. Rennert, M. Inayathullah, M. [299] A. King, S. Balaji, S.G. Keswani, T.M. Crombleholme, The role of stem cells in wound Januszyk, M. Rodrigues, A.V. Malkovskiy, A.J. Whitmore, G.G. Walmsley, M.G. angiogenesis, Adv Wound Care (New Rochelle) 3 (2014) 614–625. Galvez, A.J. Whittam, M. Brownlee, J. Rajadas, G.C. Gurtner, Transdermal deferox- [300] D.O. Zamora, S. Natesan, S. Becerra, N. Wrice, E. Chung, L.J. Suggs, R.J. Christy, En- amine prevents pressure-induced diabetic ulcers, Proc. Natl. Acad. Sci. U. S. A. hanced wound vascularization using a dsASCs seeded FPEG scaffold, Angiogenesis 112 (2015) 94–99. 16 (2013) 745–757. [272] X. Chen, L.H. Peng, N. Li, Q.M. Li, P. Li, K.P. Fung, P.C. Leung, J.Q. Gao, The healing and [301] R.K. Garg, R.C. Rennert, D. Duscher, M. Sorkin, R. Kosaraju, L.J. Auerbach, J. Lennon, anti-scar effects of astragaloside IV on the wound repair in vitro and in vivo, J. M.T. Chung, K. Paik, J. Nimpf, J. Rajadas, M.T. Longaker, G.C. Gurtner, Capillary force Ethnopharmacol. 139 (2012) 721–727. seeding of hydrogels for adipose-derived stem cell delivery in wounds, Stem Cells [273] X. Chen, L.H. Peng, Y.H. Shan, N. Li, W. Wei, L. Yu, Q.M. Li, W.Q. Liang, J.Q. Gao, Transl. Med. 3 (2014) 1079–1089. Astragaloside IV-loaded nanoparticle-enriched hydrogel induces wound [302] R. Kosaraju, R.C. Rennert, Z.N. Maan, D. Duscher, J. Barrera, A.J. Whittam, M. healing and anti-scar activity through topical delivery, Int. J. Pharm. 447 Januszyk, J. Rajadas, M. Rodrigues, G.C. Gurtner, Adipose-derived stem cell- (2013) 171–181. seeded hydrogels increase endogenous progenitor cell recruitment and neovascu- [274] Y.H.Shan,L.H.Peng,X.Liu,X.Chen,J.Xiong,J.Q.Gao,Silkfibroin/gelatin larization in wounds, Tissue Eng. A 22 (2016) 295–305. electrospun nanofibrous dressing functionalized with astragaloside IV induces [303] P. Luz-Crawford, C. Jorgensen, F. Djouad, Mesenchymal stem cells direct the immu- healing and anti-scar effects on burn wound, Int. J. Pharm. 479 (2015) 291–301. nological fate of macrophages, Results Probl. Cell Differ. 62 (2017) 61–72. [275] W. Bylka, P. Znajdek-Awizen, E. Studzinska-Sroka, A. Danczak-Pazdrowska, M. [304] M.J. Regulski, Mesenchymal stem cells: Guardians of inflammation, Wounds 29 Brzezinska, Centella asiatica in dermatology: an overview, Phytother. Res. 28 (2017) 20–27. (2014) 1117–1124. [305] R. Lim, S.D. Ricardo, W. Sievert, Cell-based therapies for tissue fibrosis, Front. [276] T.K. Biswas, B. Mukherjee, Plant medicines of Indian origin for wound healing ac- Pharmacol. 8 (2017) 633. tivity: a review, Int J Low Extrem Wounds 2 (2003) 25–39. [306] K. Rodgers, S.S. Jadhav, The application of mesenchymal stem cells to treat thermal [277] B. Kumar, M. Vijayakumar, R. Govindarajan, P. Pushpangadan, and radiation burns, Adv. Drug Deliv. Rev. 123 (2018) 75–81. Ethnopharmacological approaches to wound healing–exploring medicinal plants [307] Y. An, W. wei, H. Jing, L. Ming, S. Liu, Y. Jin, Bone marrow mesenchymal stem cell of India, J. Ethnopharmacol. 114 (2007) 103–113. aggregate: an optimal cell therapy for full-layer cutaneous wound vascularization [278] Y. Kimura, M. Sumiyoshi, K. Samukawa, N. Satake, M. Sakanaka, Facilitating action and regeneration, Sci. Rep. 5 (2015), 17036. . of asiaticoside at low doses on burn wound repair and its mechanism, Eur. J. [308] H. Sekine, T. Shimizu, I. Dobashi, K. Matsuura, N. Hagiwara, M. Takahashi, E. Pharmacol. 584 (2008) 415–423. Kobayashi, M. Yamato, T. Okano, Cardiac cell sheet transplantation improves dam- [279] T. Phaechamud, K. Yodkhum, J. Charoenteeraboon, Y. Tabata, Chitosan-aluminum aged heart function via superior cell survival in comparison with dissociated cell monostearate composite sponge dressing containing asiaticoside for wound injection, Tissue Eng Part A 17 (2011) 2973–2980. healing and angiogenesis promotion in chronic wound, Mater. Sci. Eng. C Mater. [309] C.C. Huang, H.W. Tsai, W.Y. Lee, W.W. Lin, D.Y. Chen, Y.W. Hung, J.W. Chen, S.M. Biol. Appl. 50 (2015) 210–225. Hwang, Y. Chang, H.W. Sung, A translational approach in using cell sheet frag- [280] C.Z. Zhang, J. Niu, Y.S. Chong, Y.F. Huang, Y. Chu, S.Y. Xie, Z.H. Jiang, L.H. Peng, Po- ments of autologous bone marrow-derived mesenchymal stem cells for cellular rous microspheres as promising vehicles for the topical delivery of poorly soluble cardiomyoplasty in a porcine model, Biomaterials 34 (2013) 4582–4591. asiaticoside accelerate wound healing and inhibit scar formation in vitro [310] R. Geesala, N. Bar, N.R. Dhoke, P. Basak, A. Das, Porous polymer scaffold for on-site &in vivo, Eur. J. Pharm. Biopharm. 109 (2016) 1–13. delivery of stem cells–Protects from oxidative stress and potentiates wound tissue [281] X.F. Zheng, X.Y. Lu, Measurement andcorrelation of solubilities of asiaticoside in repair, Biomaterials 77 (2016) 1–13. water, methanol, ethanol, n-propanol, n-butanol, and a methanol plus water mix- [311] M. Schafer, S. Werner, Oxidative stress in normal and impaired wound repair, ture from (278.15 to 343.15) K, J. Chem. Eng. Data 56 (2011) 674–677. Pharmacol. Res. 58 (2008) 165–171. [282] G. Sun, X. Zhang, Y.I. Shen, R. Sebastian, L.E. Dickinson, K. Fox-Talbot, M. Reinblatt, [312] J. Case, D.A. Ingram, L.S. Haneline, Oxidative stress impairs endothelial progenitor C. Steenbergen, J.W. Harmon, S. Gerecht, Dextran hydrogel scaffolds enhance an- cell function, Antioxid. Redox Signal. 10 (2008) 1895–1907. giogenic responses and promote complete skin regeneration during burn wound [313] K.L. Kim, S.H. Song, K.S. Choi, W. Suh, Cooperation of endothelial and smooth mus- healing, Proc. Natl. Acad. Sci. U. S. A. 108 (2011) 20976–20981. cle cells derived from human induced pluripotent stem cells enhances neovascu- [283] Y.I. Shen, H.G. Song, A. Papa, J. Burke, S.W. Volk, S. Gerecht, Acellular hydrogels for larization in dermal wounds, Tissue Eng Part A 19 (2013) 2478–2485. regenerative burn wound healing: Translation from a porcine model, J Invest [314] S.J. Lee, C. Park, J.Y. Lee, S. Kim, P.J. Kwon, W. Kim, Y.H. Jeon, E. Lee, Y.S. Yoon, Gen- Dermatol 135 (2015) 2519–2529. eration of pure lymphatic endothelial cells from human pluripotent stem cells and [284] S. Zhao, L. Li, H. Wang, Y. Zhang, X. Cheng, N. Zhou, M.N. Rahaman, Z. Liu, W. their therapeutic effects on wound repair, Sci. Rep. 5 (2015), 11019. . Huang, C. Zhang, Wound dressings composed of copper-doped borate bioactive [315] J. Zhang, J. Guan, X. Niu, G. Hu, S. Guo, Q. Li, Z. Xie, C. Zhang, Y. Wang, Exosomes glass microfibers stimulate angiogenesis and heal full-thickness skin defects in a released from human induced pluripotent stem cells-derived MSCs facilitate cuta- rodent model, Biomaterials 53 (2015) 379–391. neous wound healing by promoting collagen synthesis and angiogenesis, J. Transl. [285] B.P. Toole, Hyaluronan: from extracellular glue to pericellular cue, Nat. Rev. Cancer Med. 13 (2015) 49. 4(2004)528–539. [316] P. Kong, X. Xie, F. Li, Y. Liu, Y. Lu, Placenta mesenchymal stem cell accelerates [286] J.R. Fraser, T.C. Laurent, U.B. Laurent, Hyaluronan: its nature, distribution, functions wound healing by enhancing angiogenesis in diabetic Goto-Kakizaki (GK) rats, and turnover, J. Intern. Med. 242 (1997) 27–33. Biochem. Biophys. Res. Commun. 438 (2013) 410–419. [287] E.L. Pardue, S. Ibrahim, A. Ramamurthi, Role of hyaluronan in angiogenesis and its [317] S.H. Abd-Allah, A.S. El-Shal, S.M. Shalaby, E. Abd-Elbary, N.F. Mazen, R.R. Abdel utility to angiogenic tissue engineering, Organ 4 (2008) 203–214. Kader, The role of placenta-derived mesenchymal stem cells in healing of induced [288] R. Deed, P. Rooney, P. Kumar, J.D. Norton, J. Smith, A.J. Freemont, S. Kumar, Early- full-thickness skin wound in a mouse model, IUBMB Life 67 (2015) 701–709. response gene signalling is induced by angiogenic oligosaccharides of hyaluronan [318] T. Ke, M. Yang, D. Mao, M. Zhu, Y. Che, D. Kong, C. Li, Co-transplantation of skin- in endothelial cells. Inhibition by non-angiogenic, high-molecular-weight derived precursors and collagen sponge facilitates diabetic wound healing by pro- hyaluronan, Int. J. Cancer 71 (1997) 251–256. moting local vascular regeneration, Cell Physiol Biochem 37 (2015) 1725–1737. [289] J.E. Rayahin, J.S. Buhrman, Y. Zhang, T.J. Koh, R.A. Gemeinhart, High and low molec- [319] Y. Iwata, H. Akamatsu, Y. Hasebe, S. Hasegawa, K. Sugiura, Skin-resident stem cells ular weight hyaluronic acid differentially influence macrophage activation, ACS and wound healing, Nihon Rinsho Meneki Gakkai Kaishi 40 (2017) 1–11. Biomater Sci Eng 1 (2015) 481–493. [320] R. Nunan, K.G. Harding, P. Martin, Clinical challenges of chronic wounds: searching [290] A.C. Petrey, C.A. de la Motte, Hyaluronan, a crucial regulator of inflammation, Front. for an optimal animal model to recapitulate their complexity, Dis. Model. Mech. 7 Immunol. 5 (2014) 101. (2014) 1205–1213. [291] M. Litwiniuk, A. Krejner, M.S. Speyrer, A.R. Gauto, T. Grzela, Hyaluronic acid in in- [321] A.E. Boniakowski, A.S. Kimball, B.N. Jacobs, S.L. Kunkel, K.A. Gallagher, Macrophage- flammation and tissue regeneration, Wounds 28 (2016) 78–88. mediated inflammation in normal and diabetic wound healing, J. Immunol. 199 [292] F. Gao, Y. Liu, Y. He, C. Yang, Y. Wang, X. Shi, G. Wei, Hyaluronan oligosaccharides (2017) 17–24. promote excisional wound healing through enhanced angiogenesis, Matrix Biol. 29 [322] M. Muller, C. Trocme, B. Lardy, F. Morel, S. Halimi, P.Y. Benhamou, Matrix metallo- (2010) 107–116. proteinases and diabetic foot ulcers: the ratio of MMP-1 to TIMP-1 is a predictor of [293] K.S. Girish, K. Kemparaju, The magic glue hyaluronan and its eraser hyaluronidase: wound healing, Diabet. Med. 25 (2008) 419–426. a biological overview, Life Sci. 80 (2007) 1921–1943. [323] C. Dunnill, T. Patton, J. Brennan, J. Barrett, M. Dryden, J. Cooke, D. Leaper, N.T. [294] L. Soltes, R. Mendichi, G. Kogan, J. Schiller, M. Stankovska, J. Arnhold, Degradative Georgopoulos, Reactive oxygen species (ROS) and wound healing: the functional action of reactive oxygen species on hyaluronan, Biomacromolecules 7 (2006) role of ROS and emerging ROS-modulating technologies for augmentation of the 659–668. healing process, Int. Wound J. 14 (2017) 89–96. [295] A.W. Griffioen, M.J.H. Coenen, C.A. Damen, S.M.M. Hellwig, D.H.J. vanWeering, W. [324] F.A. Abreu, C.L. Ferreira, G.A. Silva, O. Paulo Cde, M.N. Miziara, F.F. Silveira, J.B. Alves, Vooys, G.H. Blijham, G. Groenewegen, CD44 is involved in tumor angiogenesis; Effect of PDGF-BB, IGF-I growth factors and their combination carried by liposomes an activation antigen on human endothelial cells, Blood 90 (1997) 1150–1159. in tooth socket healing, Braz. Dent. J. 24 (2013) 299–307. [296] G. Cao, R.C. Savani, M. Fehrenbach, C. Lyons, L. Zhang, G. Coukos, H.M. Delisser, In- [325] F. Garcia-Esteo, G. Pascual, A. Gallardo, J. San-Roman, J. Bujan, J.M. Bellon, A biode- volvement of endothelial CD44 during in vivo angiogenesis, Am. J. Pathol. 169 gradable copolymer for the slow release of growth hormone expedites scarring in (2006) 325–336. diabetic rats, J Biomed Mater Res B Appl Biomater 81 (2007) 291–304. [297] R.C. Savani, G. Cao, P.M. Pooler, A. Zaman, Z. Zhou, H.M. DeLisser, Differential in- [326] I. Struman, F. Bentzien, H. Lee, V. Mainfroid, G. D'Angelo, V. Goffin, R.I. Weiner, J.A. volvement of the hyaluronan (HA) receptors CD44 and receptor for HA- Martial, Opposing actions of intact and N-terminal fragments of the human prolac- mediated motility in endothelial cell function and angiogenesis, J. Biol. Chem. tin/growth hormone family members on angiogenesis: an efficient mechanism for 276 (2001) 36770–36778. the regulation of angiogenesis, Proc. Natl. Acad. Sci. U. S. A. 96 (1999) 1246–1251. [298] F.M. Chen, L.A. Wu, M. Zhang, R. Zhang, H.H. Sun, Homing of endogenous stem/pro- [327] Y. Zhu, Z. Li, Y. Wang, L. Li, D. Wang, W. Zhang, L. Liu, H. Jiang, J. Yang, J. Cheng, genitor cells for in situ tissue regeneration: Promises, strategies, and translational Overexpression of miR-29b reduces collagen biosynthesis by inhibiting heat perspectives, Biomaterials 32 (2011) 3189–3209. shock protein 47 during skin wound healing, Transl. Res. 178 (2016) 38–53 e36. A.P. Veith et al. / Advanced Drug Delivery Reviews 146 (2019) 97–125 125

[328] T. Lucas, F. Schafer, P. Muller, S.A. Eming, A. Heckel, S. Dimmeler, Light-inducible [333] T.S. Iyyanki, L.W. Dunne, Q. Zhang, J. Hubenak, K.C. Turza, C.E. Butler, Adipose- antimiR-92a as a therapeutic strategy to promote skin repair in healing-impaired derived stem-cell-seeded non-cross-linked porcine acellular dermal matrix in- diabetic mice, Nat. Commun. 8 (2017), 15162. . creases cellular infiltration, vascular infiltration, and mechanical strength of ven- [329] Y.C. Chan, S. Roy, Y. Huang, S. Khanna, C.K. Sen, The microRNA miR-199a-5p down- tral hernia repairs, Tissue Eng. A 21 (2015) 475–485. regulation switches on wound angiogenesis by derepressing the v-ets erythroblas- [334] Z. Li, H. Wang, B. Yang, Y. Sun, R. Huo, Three-dimensional graphene foams loaded tosis virus E26 oncogene homolog 1-matrix metalloproteinase-1 pathway, J. Biol. with bone marrow derived mesenchymal stem cells promote skin wound healing Chem. 287 (2012) 41032–41043. with reduced scarring, Mater. Sci. Eng. C Mater. Biol. Appl. 57 (2015) 181–188. [330] P. Foubert, S. Barillas, A.D. Gonzalez, Z. Alfonso, S. Zhao, I. Hakim, C. Meschter, M. [335] X.Y. Chan, R. Black, K. Dickerman, J. Federico, M. Levesque, J. Mumm, S. Gerecht, Tenenhaus, J.K. Fraser, Uncultured adipose-derived regenerative cells (ADRCs) Three-dimensional vascular network assembly from diabetic patient-derived in- seeded in collagen scaffold improves dermal regeneration, enhancing early vascu- duced pluripotent stem cells, Arterioscler. Thromb. Vasc. Biol. 35 (2015) larization and structural organization following thermal burns, Burns 41 (2015) 2677–2685. 1504–1516. [336] T. Ke, M. Yang, D. Mao, M. Zhu, Y. Che, D. Kong, C. Li, Co-tansplantation of skin- [331] Y. Kato, T. Iwata, S. Morikawa, M. Yamato, T. Okano, Y. Uchigata, Allogeneic trans- derived precursors and collagen sponge facilitates diabetic wound healing by pro- plantation of an adipose-derived stem cell sheet combined with artificial skin ac- moting local vascular regeneration, Cell Physiol Biochem 37 (2015) 1725–1737. celerates wound healing in a rat wound model of type 2 diabetes and obesity, [337] H.S. Milch, S.Y. Schubert, S. Hammond, J.H. Spiegel, Enhancement of ischemic Diabetes 64 (2015) 2723–2734. wound healing by inducement of local angiogenesis, Laryngoscope 120 (2010) [332] I.S. Park, A. Mondal, P.S. Chung, J.C. Ahn, Vascular regeneration effect of adipose- 1744–1748. derived stem cells with light-emitting diode phototherapy in ischemic tissue, La- sers Med. Sci. 30 (2015) 533–541.