Next Article in Journal
Composition of Transcription Machinery and Its Crosstalk with Nucleoid-Associated Proteins and Global Transcription Factors
Next Article in Special Issue
The Effects of Hyperbaric Oxygenation on Oxidative Stress, Inflammation and Angiogenesis
Previous Article in Journal
Kaempferol Ameliorates Oxygen-Glucose Deprivation/Reoxygenation-Induced Neuronal Ferroptosis by Activating Nrf2/SLC7A11/GPX4 Axis
Previous Article in Special Issue
Oxygen: The Rate-Limiting Factor for Episodic Memory Performance, Even in Healthy Young Individuals
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Commentary

Molecular Biomarkers of Oxygen Therapy in Patients with Diabetic Foot Ulcers

by
Alisha R. Oropallo
1,*,
Thomas E. Serena
2,
David G. Armstrong
3 and
Mark Q. Niederauer
4
1
Comprehensive Wound Healing Center and Hyperbarics, Department of Vascular Surgery, Zucker School of Medicine Hofstra/Northwell, Hempstead, NY 11549, USA
2
Serena Group Research Foundation, Cambridge, MA 02140, USA
3
Limb Preservation Program, Department of Surgery, Keck School of Medicine of University of Southern California, Los Angeles, CA 90033, USA
4
EO2 Concepts, San Antonio, TX 78249, USA
*
Author to whom correspondence should be addressed.
Biomolecules 2021, 11(7), 925; https://doi.org/10.3390/biom11070925
Submission received: 30 May 2021 / Revised: 13 June 2021 / Accepted: 19 June 2021 / Published: 22 June 2021
(This article belongs to the Special Issue Oxygen Therapy)

Abstract

:
Hyperbaric oxygen therapy (HBOT) and topical oxygen therapy (TOT) including continuous diffuse oxygen therapy (CDOT) are often utilized to enhance wound healing in patients with diabetic foot ulcerations. High pressure pure oxygen assists in the oxygenation of hypoxic wounds to increase perfusion. Although oxygen therapy provides wound healing benefits to some patients with diabetic foot ulcers, it is currently performed from clinical examination and imaging. Data suggest that oxygen therapy promotes wound healing via angiogenesis, the creation of new blood vessels. Molecular biomarkers relating to tissue inflammation, repair, and healing have been identified. Predictive biomarkers can be used to identify patients who will most likely benefit from this specialized treatment. In diabetic foot ulcerations, specifically, certain biomarkers have been linked to factors involving angiogenesis and inflammation, two crucial aspects of wound healing. In this review, the mechanism of how oxygen works in wound healing on a physiological basis, such as cell metabolism and growth factor signaling transduction is detailed. Additionally, observable clinical outcomes such as collagen formation, angiogenesis, respiratory burst and cell proliferation are described. The scientific evidence for the impact of oxygen on biomolecular pathways and its relationship to the outcomes in clinical research is discussed in this narrative review.

1. Introduction

Lower extremity complications in people with diabetes constitute a large worldwide burden within in already burdened population [1,2]. Every 1.2 s, someone with diabetes develops a foot ulcer [1]. More than half of these wounds become infected [3,4,5], leading to a high rate of emergency department visits, hospitalizations and ultimately amputation [6]. Every 20 s, someone with diabetes undergoes an amputation somewhere in the world [7,8,9]. Patients with diabetic foot ulcers are at a near three-fold greater risk for death in the year following wounding than patients with diabetes without foot ulcers [10]. This increases with additional comorbidities. Following ulceration, Charcot arthropathy, development of chronic limb threatening ischemia or amputation, 5-year mortality is comparable to most cancers [11,12,13]. Additionally, the costs for care for patients with diabetic foot ulcers exceed the cost of care most individual cancers [11,14,15].
The role of oxygen in wound healing has long garnered interest among researchers and clinicians alike. This interest has only increased as modalities for delivery of oxygen have evolved from large hyperbaric chambers to portable, direct topical application using localized chambers and more recently to handheld, wearable systems which continuously diffuse oxygen directly into the wound bed. There are distinct differences and advantages of each modality for oxygen delivery. Although oxygen therapy can be used for a variety of etiologies, the focus of oxygen therapy discussed revolves around its most common usage which is diabetic foot ulcerations due to its element of ischemia, either involving a macro or micro circulatory component. In hyperbaric oxygen therapy (HBOT) a contained chamber is pressurized with 100% oxygen to 2.0–2.4 atmospheres absolute for 90 min 5–7 days per week. HBOT relies on respiration and the circulatory system to deliver oxygen to the wound bed [16]. The increased pressure supersaturates the plasma; however, oxygen delivery relies on local capillary structure to reach injured tissues. Deficient or absent capillary beds may impede the oxygen delivery to ischemic tissues. Traditional TOT uses high flow oxygen concentrators coupled with chambers or bags placed directly over or around the wound. TOT applies oxygen directly to the wound, allowing the oxygen to diffuse directly into the wound and is therefore not reliant on underlying capillary structures. TOT follows an intermittent treatment regimen similar to HBOT. Currently, most recent devices are wearable and continuously generate pure, humidified oxygen from surrounding air using electrochemical oxygen generators. There is no need for an external oxygen source. They continuously diffuse oxygen (CDO) directly into wounds (24 h a day, 7 days a week) using an oxygen diffuser or oxygen diffusion dressings. CDOT, like TOT, does not depend on the underlying capillary structure of the wound bed, however, unlike TOT, the continuous application of oxygen resembles physiologic oxygen delivery. The biomolecular evidence for the effects of oxygen in wound healing including all modalities of delivery are presented. Although delivery mechanisms differ, the effect of oxygen at a cellular level is consistent.
However, the availability of oxygen to injured tissues will depend on the method of delivery. For HBOT, which relies on inspired oxygen, the availability depends on arterial pO2, vascular supply, local capillary structures and the diffusion distance for the oxygen from the capillaries to the cells. Both edema and necrotic debris increase the diffusion distance. If the local structures are impaired or vasoconstriction is present, wound perfusion can be significantly impaired such that little to no increase in wound pO2 levels occurs despite breathing supplemental oxygen [17,18,19]. Hence, there is a need for determining local vascular adequacy using methods such as transcutaneous oxygen pressure measurement prior to initiating HBOT. Modalities that use direct application of oxygen to the wound, such as TOT and CDOT, still require adequate vascular sufficiency, yet are significantly less dependent on local capillary structures. Necrotic tissue increases the diffusion distance to the wound, so debridement is an important step to ensure optimal diffusion of oxygen into the wound bed for topically applied oxygen. Debridement has been shown to have significant benefit when applied to standard moist wound therapies [20,21,22]. The importance of debridement in topically applied oxygen was recently demonstrated in a double blind, placebo-controlled clinical study, where CDOT showed dramatically higher wound closure rates and overall closure for wounds that were debrided frequently versus those that were not [23] in patients with diabetic foot ulcers.
The molecular processes discussed herein are oxygen dependent and do not occur without oxygen. The reactions are catalyzed by enzymes which typically have about 50% maximum speed at normal tissue pO2 levels (40–80 mm Hg) and reach 90% of maximum speed at levels varying between about 150 mm Hg to over 400 mm Hg [24,25]. These higher levels can only be achieved with supplemental oxygen. An interesting finding regarding the positive correlation between oxygen concentration and functionality is that the more oxygen there is, the faster and better the outcomes are compared to normal wound healing. The differences are even greater when compared to ischemic wounds which are hypoxic. The definitions of hypoxia and hyperoxia are relative. In the context of this review, they are relative to the levels normally found in healthy tissue surrounding a wound (40–80 mm Hg).

2. Cell Metabolism and Energy

Oxygen plays a crucial role in energy production and cell metabolism. In this role, oxygen is required for intracellular processes such as biosynthesis and transport, not to mention cell survival [26]. Oxygen dependent enzymes include adenosine triphosphate (ATP) for chemical energy and nicotinamide adenine dinucleotide phosphate (NADPH) oxygenase for respiratory burst (reactive oxygen species release). ATP fuels most active cellular processes and the increased energy demand of tissue that is undergoing healing leads to a hypermetabolic state wherein additional energy is generated from oxidative metabolism [27,28,29,30]. Other metabolic processes such as aerobic glycolysis, ß-oxidation of fatty acids and the citric acid cycle are tightly attached to the energy acquisition by oxidative phosphorylation and are, therefore, oxygen dependent [31]. Conversely, when tissue oxygen levels are consistently too low (<20 mmHg pO2), cells convert to anaerobic metabolism and go into survival mode in which wound healing activities such as mitotic cell division, and, therefore, re-epithelialization with collagen production are impaired [32,33,34]. Prolonged exposure to extremely low oxygen levels, if not alleviated by oxygen, can result in cell death and tissue necrosis due to the inability of the cells to repair the spontaneous decay of cell components (DNA, RNA and proteins) and inability to maintain calcium pumps which require ATP to function [35,36].

3. Molecular Biomarkers in Growth Factor Signaling Transduction

Reactive oxygen species (ROS) are essential for the signaling processes of growth factors and processes such as leukocyte recruitment, cell motility, angiogenesis and extracellular matrix formation involved in wound healing [37]. The rate-limiting substrate for ROS production is oxygen. In a wound site, almost all wound-related cells can generate ROS using the enzyme nicotinamide adenine dinucleotide phosphate (NADPH) oxidase. The functionality of NADPH oxidase correlates positively to pO2 levels, with the maximal function of NADPH oxidase observed at pO2 > 300 mm Hg, levels only achievable with supplemental oxygen. In wounds deficient of oxygen, such as diabetic or ischemic wounds, NADPH oxidase ceases to function at pO2 levels below 20 mm Hg. There have been no noted adverse effects or increased reports of safety issues associated with high concentrations of oxygen in wound care. The increased ROS levels appear to accelerate the signaling processes without causing any damage at the cellular level. On a clinical level, studies have shown comparable or decreased adverse events and hospitalizations compared to standard of care with no supplemental oxygen [38,39,40,41].
Signal transduction of growth factors and cytokines is stimulated by ROS [42]. ROS, such as superoxide and hydrogen peroxide, increase vascular endothelial growth factor (VEGF) production in macrophages and keratinocytes [43,44]. ROS are also required for platelet-derived growth factor (PDGF) to regulate cell growth and division [45]. Like VEGF, PDGF plays a significant role in blood vessel formation (angiogenesis) [46]. ROS have effects on other processes such as cytokine action, cell motility and extracellular matrix formation [47]. Conversely, tissue hypoxia will limit redox signaling and disable the function of several growth factors such as PDGF, VEGF, keratinocyte growth factor, insulin-like growth factor one (IGF-1), transforming growth factor beta (TFG-β) and numerous molecular mechanisms (e.g., leukocyte recruitment, cell motility and integrin function) which rely on redox signaling [37,48,49]. This positive correlation between pO2 levels, ROS production, and growth factor promotion of cytokine expression explains why ischemic diabetic wounds, having little to no ROS, fail to heal and why wounds supplemented with oxygen heal faster. Typical molecular biomarkers that are indicative of wound healing are shown in Table 1 along with the processes that they are associated with. These biomarkers and their effects on wound healing will be discussed in greater detail throughout this review.
The impact of continuous diffusion of oxygen therapy (CDOT) on wound cytokines and growth factors was recently demonstrated in a prospective study of 23 patients with diabetic foot ulcers below the malleolus [50]. Results showed significant increases in growth factors, cytokines and transcutaneous oxygen pressure measurement levels after application of CDOT. Growth factors significantly increased from 280% to 820% of base levels in the first week and decreased in subsequent weeks [50] (Figure 1). Cytokines increased significantly (up to 680% compared to baseline levels) in the first two weeks and then decreased. Significant increases in transcutaneous oxygen pressure measurement indicated increased oxygen perfusion in the wound periphery. This is evidence that the topically applied oxygen not only saturated the wound bed, yet also elevated the levels of oxygen in the surrounding tissues.

4. Collagen Formation

Oxygen is essential to make and properly organize collagen, which is the primary component of skin, accounting for 70–80% of dry weight and acts as the primary structural scaffold of skin and structures the matrix for angiogenesis. Organized collagen is bundled into fibers, which are interwoven and can be stretched in multiple directions without tearing. At the biomolecular level, oxygen is required for the hydroxylation of proline and lysine in procollagen [51]. Several posttranslational steps in collagen synthesis are oxygen dependent. The enzymes prolyl hydroxylase, lysyl hydroxylase and lysyl oxidase all require oxygen [46,52,53]. The formation of cross-linked triple-helices occurs via the oxygen-dependent enzyme prolyl hydroxylase and are excreted as collagen fibers. Collagen fibers are arranged into linear fibrils via cross-linking by lysyl hydroxylase. Linear fibrils are cross-linked by lysyl oxidase—a necessary step to achieve the necessary tensile strength for healed wounds.
Higher oxygen concentrations increase the amount of collagen deposition [54] and tensile strength [55,56,57]. The rate limiting step is the rate of prolyl hydroxylation [52,53]. The oxygen level required for optimal prolyl hydryoxlase activity is at oxygen levels approaching 250 mmHg, exceeding those present in normal wounds and only achievable using oxygen therapy treatment [58,59]. It has been shown that increasing oxygen concentrations above normal physiologic levels enhances collagen synthesis and tensile strength in both animal and human subjects [55,56,57] and can increase the level of collagen organization [60]. Correction of vasoconstriction and hypoxia can result in a 10-fold increase in collagen deposition in wound repair [17,54,56,61]. The rates of collagen deposition increase as oxygen levels increase, with optimal activity at levels higher than 250 mmHg [62]. Conversely, hypoxic wounds as in patients with diabetes deposit collagen poorly and become infected easily [51,54],
In a study using supplemental oxygen at a rate of 4 L/min through nasal cannula for 12 h a day for 3 days, it was found that three times as much collagen was deposited in patients with well-perfused and oxygenated wounds compared with those with lower oxygenation and perfusion scores [54]. A separate study using direct topical oxygen on chronic diabetic foot ulcers showed significant increases in the expression of genes associated with collagen production (TGF-β, VEGF and IL-6) during weekly follow-up visits after application of CDO in patients with diabetic foot ulcers [50].

5. Angiogenesis Biomarkers

The creation of new blood vessels, angiogenesis, is essential to the growth and survival of repair tissue. Oxygen levels directly affect not only the rate, yet also the quality of new blood vessel growth. Sufficient oxygen levels are required for correct collagen synthesis (posttranslational hydroxylation) [63] without which the new capillary tubes assemble poorly and remain fragile [62,64,65]. Supplemental oxygen has been shown to accelerate blood vessel growth [66]. Moderate hyperoxia increases the appearance of new blood vessels in wounds [67]. Similar to ROS activity, the rate of angiogenesis has been shown to be directly proportional to oxygen levels in damaged tissues [62], with maximum activity levels at pO2 levels exceeding 250 mm Hg.
VEGF has been shown to be a major long-term angiogenic stimulus at the wound site and is believed to be most prevalent and efficacious signal for angiogenesis. Oxygen treatment induces VEGF mRNA levels in endothelial cells and macrophages [68,69,70] and VEGF 121/165 protein expression in wounds [71]. Oxygen has also been shown to facilitate the release of VEGF165 from cell-associated stores [72].
Hyperbaric and topical oxygen therapy have been shown to increase VEGF expression in wounds [73] and induce angiogenesis [74]. More recently, a clinical study on gene expression of multiple factors involved in angiogenesis (VEGF, TGF-β, IL-6 and CXCL8) showed significant increases upon continuous application of oxygen (Figure 1) [50]. The expression levels over time are similar to gross observations of their effect in the field: increased redness within the first week and exudate levels that peak within the first two weeks and then subside, both indicators of new capillary formation. Furthermore, the curved response shown in Figure 1 reflects what would be expected of a chronic wound “reawakening” and entering the inflammatory stage.

6. Respiratory Burst Process and Cytokine Production

Oxygen is essential for respiratory burst, or the production of reactive oxygen species (ROS), used by phagocytes such as neutrophils and macrophages in bactericidal activity and the removal of necrotic cellular debris. NADPH oxidase, also known as leukocyte oxidase, has been shown to support macrophage survival, a delay of apoptosis and enables dead cell cleansing by phagocytosis [75]. NADPH oxidase in wound phagocytes, such as neutrophils and macrophages, produces superoxides (O2- and H2O2) for bactericidal activities [76]. It has been shown that up to 98% of oxygen consumed by these cells is used to produce ROS during phagocytosis [24]. Leukocyte activity, which involves the production of ROS which enables oxidative killing, is directly proportional to local oxygen concentration [77,78]. The optimal ROS production is seen at oxygen levels of greater than 300 mmHg, levels which can only be achieved with supplemental oxygen [79]. ROS activity is not restricted to phagocytes. At the wound site, ROS are generated by almost all wound-related cells [46]. The efficacy of supplemental oxygen has been shown to be similar to antibiotic administration and has additive effects when used together [80,81].
Interleukins are a type of cytokine protein that play important roles in the differentiation/activation of immune cells in addition to their proliferation, maturation, migration and adhesion. [https://www.ncbi.nlm.nih.gov/books/NBK499840/ StatPearls Publishing; 31 January 2021]. The addition of continuous oxygen therapy directly to a wound has been shown to increaseIL-6 significantly (up to 680% relative to baseline) in clinical studies [50,82]. IL-6 has been shown to induce chemotaxis of leukocytes into a wound [83,84]. As inflammation progresses, IL-6 signaling is responsible for the switch to a reparative environment.

7. Cell Proliferation Molecular Markers

Increasing oxygen levels results in faster cell proliferation, re-epithelialization and collagen formation. Fibroblast proliferation and protein production have been reported to be optimal at 160 mmHg, i.e., at pO2 levels two-fold to three-fold higher than those found in healthy tissues [85], indicating that supplemental oxygen increases the rate of wound repair. Endothelial progenitor cells (EPCs) are essential in wound healing, but their circulating and wound level numbers are decreased in diabetes. Elevated oxygen levels (hyperoxia) reverse the diabetic defect in EPC mobilization [86]. EPC mobilization into circulation is triggered by hyperoxia through induction of nitric oxide with resulting enhancement in ischemic limb perfusion and diabetic wound healing [87,88,89].
Matrix metalloproteinases (MMPs) are a group of enzymes responsible for degrading a majority of extracellular matrix proteins during tissue development, growth and turnover [90,91]. MMPs have diagnostic, predictive and indicative power for wound healing and can be measured from wound fluid. They are required for a wound to heal properly, at a suitable level, in the correct position and for a certain length of time. Excess activity may lead to a chronic non healing wound. Chronically increased levels of MMPs and reduced levels of TIMPs (MMP regulators), or just abnormalities in their ratio, are associated with non-healing. Studies show that medical interventions which aid in lowering MMP activity will promote the healing of stalled wounds and that decreasing MMP-2 tissue levels will result in wound healing. In one study elevated MMP-1 and TIMP-1 levels were noted in on oxygen treatment group, yet not in the control group [82].
At the clinical level, the cumulative effects of oxygen in all the various aspects discussed herein result in significant real-world results. In a clinical study which analyzed VEGF expression versus wound size reduction using TOT, a significant correlation between wound closure and VEGF expression was found [73]. Recent results using topically applied oxygen therapy, both continuously and intermittently, on diabetic foot wounds has been shown to increase the rate of wound closure, by as much as 460% relative to moist wound therapy in several double-blinded trials, two of which had placebo control groups [23,38,39,40,41]. As would be predicted by the positive correlation of various mechanisms of action to the relative concentration of oxygen, wounds that were larger, deeper, more chronic and weight-bearing had improved responses relative to controls than those that were smaller, shallower, less chronic or non-weight-bearing, respectively [38].

8. Summary

Biomolecular pathways associated with wound care have been shown to be positively correlated to local tissue oxygen concentration. Maximal activity levels of the related enzymes, growth factors and cytokines, as well as the associated physiological processes, are significantly above the levels normally found in healthy tissues. The positive, differential effect from supplemental oxygen has been shown to be even higher for tissues with compromised blood supply leading to ischemic diabetic wounds. Increasing the levels of oxygen in afflicted tissues significantly increases not only the rate, yet also the quality of tissue repair. These elevated levels of oxygen can only be achieved through supplemental oxygen therapy, whether it be respiratory based (HBOT) or directly applied to the wound (TOT, CDOT), all of which are reliant on diffusion gradients. The recent research on the scientific basis and clinical outcomes of oxygen therapy lays a foundation for further research in molecular biomarkers utilizing oxygen therapy.

Author Contributions

Conceptualization, A.R.O., M.Q.N.; resources, A.R.O., T.E.S., D.G.A. and M.Q.N.; data curation, A.R.O., T.E.S., D.G.A. and M.Q.N.; writing—original draft preparation, A.R.O. and M.Q.N.; writing—review and editing, A.R.O., T.E.S., D.G.A. and M.Q.N. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data available on PubMed.

Conflicts of Interest

Authors ARO and TES s declare no conflict of interest. MQN is an employee of EO2 Concepts. DGA is a member of the Scientific Advisory Board of EO2 concepts.

References

  1. Armstrong, D.G.; Boulton, A.J.M.; Bus, S.A. Diabetic Foot Ulcers and Their Recurrence. N. Engl. J. Med. 2017, 376, 2367–2375. [Google Scholar] [CrossRef] [PubMed]
  2. Zhang, Y.; Lazzarini, P.A.; McPhail, S.M.; Van Netten, J.J.; Armstrong, D.G.; Pacella, R.E. Global Disability Burdens of Diabetes-Related Lower-Extremity Complications in 1990 and 2016. Diabetes Care 2020, 43, 964–974. [Google Scholar] [CrossRef] [PubMed]
  3. Lipsky, B.A.; Senneville, É.; Urbančič-Rovan, V.; Peters, E.J.; on behalf of the International Working Group on the Diabetic Foot(IWGDF); Abbas, Z.G.; Aragón-Sánchez, J.; Diggle, M.; Embil, J.M.; Kono, S.; et al. Guidelines on the diagnosis and treatment of foot infection in persons with diabetes (IWGDF 2019 update). Diabet. Metab. Res. Rev. 2020, 36 (Suppl. 1), e3280. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Armstrong, D.G.; A Lavery, L.; Frykberg, R.G.; Wu, S.C.; Boulton, A.J. Validation of a diabetic foot surgery classification. Int. Wound J. 2006, 3, 240–246. [Google Scholar] [CrossRef]
  5. Skrepnek, G.H.; Armstrong, D.G.; Mills, J.L. 2,500,000 Troubled Soles: Ten-Year Analysis of Diabetic Foot Infections in the United States. J. Vasc. Surg. 2013, 58, 558. [Google Scholar] [CrossRef] [Green Version]
  6. Skrepnek, G.H.; Mills, J.L., Sr.; Armstrong, D.G. A Diabetic Emergency One Million Feet Long: Disparities and Burdens of Illness among Diabetic Foot Ulcer Cases within Emergency Departments in the United States, 2006–2010. PLoS ONE 2015, 10, e0134914. [Google Scholar] [CrossRef] [Green Version]
  7. Armstrong, D.G.; A Lavery, L.; Harkless, L.B.; Van Houtum, W.H. Amputation and reamputation of the diabetic foot. J. Am. Podiatr. Med. Assoc. 1997, 87, 255–259. [Google Scholar] [CrossRef]
  8. Lipsky, B.A.; Berendt, A.R.; Cornia, P.B.; Pile, J.C.; Peters, E.J.G.; Armstrong, D.G.; Deery, H.G.; Embil, J.M.; Joseph, W.S.; Karchmer, A.W.; et al. Executive Summary: 2012 Infectious Diseases Society of America Clinical Practice Guideline for the Diagnosis and Treatment of Diabetic Foot Infections. Clin. Infect. Dis. 2012, 54, 1679–1684. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  9. Boulton, A.J.; Armstrong, D.G.; Kirsner, R.S.; Attinger, C.E.; Lavery, L.A.; Lipsky, B.A.; Mills, J.L.; Steinberg, J.S. Diagnosis and Management of Diabetic Foot Complications; American Diabetes Association: Arlington, VA, USA, 2018. [Google Scholar] [CrossRef]
  10. Saluja, S.; Anderson, S.G.; Hambleton, I.; Shoo, H.; Livingston, M.; Jude, E.B.; Lunt, M.; Dunn, G.; Heald, A.H. Foot ulceration and its association with mortality in diabetes mellitus: A meta-analysis. Diabet. Med. 2020, 37, 211–218. [Google Scholar] [CrossRef] [Green Version]
  11. Armstrong, D.G.; Swerdlow, M.A.; Armstrong, A.A.; Conte, M.S.; Padula, W.V.; Bus, S.A. Five year mortality and direct costs of care for people with diabetic foot complications are comparable to cancer. J. Foot Ankle Res. 2020, 13, 16. [Google Scholar] [CrossRef] [Green Version]
  12. Conte, M.S.; Bradbury, A.W.; Kolh, P.; White, J.V.; Dick, F.; Fitridge, R.; Mills, J.L.; Ricco, J.-B.; Suresh, K.R.; Murad, M.H.; et al. Global vascular guidelines on the management of chronic limb-threatening ischemia. J. Vasc. Surg. 2019, 69, 3S–125S.e40. [Google Scholar] [CrossRef] [Green Version]
  13. Mills, J.L., Sr.; Conte, M.S.; Armstrong, D.G.; Pomposelli, F.B.; Schanzer, A.; Sidawy, A.N.; Andros, G. Society for Vascular Surgery Lower Extremity Guidelines Committee. The Society for Vascular Surgery Lower Extremity Threatened Limb Classification System: Risk stratification based on Wound, Ischemia, and foot Infection (WIfI). J. Vasc. Surg. 2014, 59, 220–234.e2. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Nussbaum, S.R.; Carter, M.J.; Fife, C.E.; DaVanzo, J.; Haught, R.; Nusgart, M.; Cartwright, D. An Economic Evaluation of the Impact, Cost, and Medicare Policy Implications of Chronic Nonhealing Wounds. Value Health 2018, 21, 27–32. [Google Scholar] [CrossRef] [Green Version]
  15. Barshes, N.R.; Sigireddi, M.; Wrobel, J.S.; Mahankali, A.; Robbins, J.M.; Kougias, P.; Armstrong, D.G. The system of care for the diabetic foot: Objectives, outcomes, and opportunities. Diabet. Foot Ankle 2013, 4, 21847. [Google Scholar] [CrossRef] [PubMed]
  16. Howard, M.A.; Asmis, R.; Evans, K.K.; Mustoe, T.A. Oxygen and wound care: A review of current therapeutic modalities and future direction. Wound Repair Regen. 2013, 21, 503–511. [Google Scholar] [CrossRef]
  17. Hopf, H.W.; Hunt, T.K.; West, J.M.; Blomquist, P.; Goodson, W.; Jensen, J.; Jonsson, K.; Paty, P.B.; Rabkin, J.M.; Upton, R.A.; et al. Wound Tissue Oxygen Tension Predicts the Risk of Wound Infection in Surgical Patients. Arch. Surg. 1997, 132, 997–1004. [Google Scholar] [CrossRef] [PubMed]
  18. Gottrup, F.; Firmin, R.; Rabkin, J.; Halliday, B.J.; Hunt, T.K. Directly measured tissue oxygen tension and arterial oxygen tension assess tissue perfusion. Crit. Care Med. 1987, 15, 1030–1036. [Google Scholar] [CrossRef]
  19. Hopf, H.; West, J.; Hunt, T. Clonidine increases tissue oxygen in patients with local tissue hypoxia in non-healing wounds. Wound Repair Regen. 1996, 4, A129. [Google Scholar]
  20. Edwards, J.; Stapley, S. Debridement of diabetic foot ulcers. Cochrane Database Syst. Rev. 2010, CD003556. [Google Scholar] [CrossRef]
  21. Armstrong, D.G.; Lavery, L.A.; Nixon, B.P.; Boulton, A.J.M. It’s Not What You Put On, but What You Take Off: Techniques for Debriding and Off-Loading the Diabetic Foot Wound. Clin. Infect. Dis. 2004, 39, S92–S99. [Google Scholar] [CrossRef]
  22. Lebrun, E.; Tomic-Canic, M.; Kirsner, R.S. The role of surgical debridement in healing of diabetic foot ulcers. Wound Repair Regen. 2010, 18, 433–438. [Google Scholar] [CrossRef] [PubMed]
  23. Lavery, L.A.; Niederauer, M.Q.; Papas, K.K.; Armstrong, D.G. Does Debridement Improve Clinical Outcomes in People with DFU Ulcers Treated with CDO? Wounds 2019, 31, 246–251. [Google Scholar] [PubMed]
  24. Allen, D.B.; Maguire, J.J.; Mahdavian, M.; Wicke, C.; Marcocci, L.; Scheuenstuhl, H.; Chang, M.; Le, A.X.; Hopf, H.W.; Hunt, T.K. Wound Hypoxia and Acidosis Limit Neutrophil Bacterial Killing Mechanisms. Arch. Surg. 1997, 132, 991–996. [Google Scholar] [CrossRef] [PubMed]
  25. Hopf, H.W.; Humphrey, L.M.; Puzziferri, N.; West, J.M.; Attinger, C.E.; Hunt, T.K. Adjuncts to preparing wounds for closure: Hyperbaric oxygen, growth factors, skin substitutes, negative pressure wound therapy (vacuum-assisted closure). Foot Ankle Clin. 2001, 6, 661–682. [Google Scholar] [CrossRef]
  26. Sen, C.K. Wound healing essentials: Let there be oxygen. Wound Repair Regen. 2009, 17, 1–18. [Google Scholar] [CrossRef] [Green Version]
  27. Gupta, A.; Raghubir, R. Energy metabolism in the granulation tissue of diabetic rats during cutaneous wound healing. Mol. Cell. Biochem. 2005, 270, 71–77. [Google Scholar] [CrossRef]
  28. Hohn, D.C.; Ponce, B.; Burton, R.W.; Hunt, T.K. Antimicrobial systems of the surgical wound. I. A comparison of oxidative metabolism and microbicidal capacity of phagocytes from wounds and from peripheral blood. Am. J. Surg. 1977, 133, 597–600. [Google Scholar] [CrossRef]
  29. Matsuda, T.; Clark, N.; Hariyani, G.D.; Bryant, R.S.; Hanumadass, M.L.; Kagan, R.J. The Effect of Burn Wound Size on Resting Energy Expenditure. J. Trauma Inj. Infect. Crit. Care 1987, 27, 115–118. [Google Scholar] [CrossRef]
  30. Im, M.J.; Hoopes, J.E. Energy metabolism in healing skin wounds. J. Surg. Res. 1970, 10, 459–464. [Google Scholar] [CrossRef]
  31. Tandara, A.A.; Mustoe, T.A. Oxygen in Wound Healing? More than a Nutrient. World J. Surg. 2004, 28, 294–300. [Google Scholar] [CrossRef]
  32. LaVan, F.B.; Hunt, T.K. Oxygen and Wound Healing. Clin. Plast. Surg. 1990, 17, 463–472. [Google Scholar] [CrossRef]
  33. Hess, C.L.; Howard, M.A.; Attinger, C.E. A Review of Mechanical Adjuncts in Wound Healing: Hydrotherapy, Ultrasound, Negative Pressure Therapy, Hyperbaric Oxygen, and Electrostimulation. Ann. Plast. Surg. 2003, 51, 210–218. [Google Scholar] [CrossRef]
  34. Hunt, T.K. Basic Principles of Wound Healing. J. Trauma Inj. Infect. Crit. Care 1990, 30, 122–128. [Google Scholar] [CrossRef]
  35. Milton, S.L.; Prentice, H.M. Beyond anoxia: The physiology of metabolic downregulation and recovery in the anoxia-tolerant turtle. Comp. Biochem. Physiol. Part A Mol. Integr. Physiol. 2007, 147, 277–290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Stys, P.K.; Ransom, B.R.; Waxman, S.G.; Davis, P.K. Role of extracellular calcium in anoxic injury of mammalian central white matter. Proc. Natl. Acad. Sci. USA 1990, 87, 4212–4216. [Google Scholar] [CrossRef] [Green Version]
  37. Sen, C.K.; Roy, S. Redox signals in wound healing. Biochim. Biophys. Acta (BBA) Gen. Subj. 2008, 1780, 1348–1361. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  38. Niederauer, M.Q.; Michalek, J.E.; Liu, Q.; Papas, K.K.; Lavery, L.A.; Armstrong, D.G. Continuous diffusion of oxygen improves diabetic foot ulcer healing when compared with a placebo control: A randomised, double-blind, multicentre study. J. Wound Care 2018, 27, S30–S45. [Google Scholar] [CrossRef]
  39. Frykberg, R.G.; Franks, P.; Edmonds, M.; Brantley, J.N.; Téot, L.; Wild, T.; Garoufalis, M.G.; Lee, A.M.; Thompson, J.A.; Reach, G.; et al. A Multinational, Multicenter, Randomized, Double-Blinded, Placebo-Controlled Trial to Evaluate the Efficacy of Cyclical Topical Wound Oxygen (TWO2) Therapy in the Treatment of Chronic Diabetic Foot Ulcers: The TWO2 Study. Diabetes Care 2019, 43, 616–624. [Google Scholar] [CrossRef] [Green Version]
  40. Serena, T.E.; Bullock, N.M.; Cole, W.; Lantis, J.; Li, L.; Moore, S.; Patel, K.; Sabo, M.; Wahab, N.; Price, P. Topical oxygen therapy in the treatment of diabetic foot ulcers: A multicentre, open, randomised controlled clinical trial. J. Wound Care 2021, 30, S7–S14. [Google Scholar] [CrossRef]
  41. Yu, J.; Lu, S.; McLaren, A.-M.; Perry, J.A.; Cross, K.M. Topical oxygen therapy results in complete wound healing in diabetic foot ulcers. Wound Repair Regen. 2016, 24, 1066–1072. [Google Scholar] [CrossRef]
  42. Sundaresan, M.; Yu, Z.-X.; Ferrans, V.J.; Sulciner, D.J.; Gutkind, J.S.; Irani, K.; Goldschmidt-Clermont, P.J.; Finkel, T. Regulation of reactive-oxygen-species generation in fibroblasts by Rac 1. Biochem. J. 1996, 318, 379–382. [Google Scholar] [CrossRef] [Green Version]
  43. Sen, C.K.; Khanna, S.; Babior, B.M.; Hunt, T.K.; Ellison, E.C.; Roy, S. Oxidant-induced Vascular Endothelial Growth Factor Expression in Human Keratinocytes and Cutaneous Wound Healing. J. Biol. Chem. 2002, 277, 33284–33290. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Cho, M.; Hunt, T.K.; Hussain, M.Z. Hydrogen peroxide stimulates macrophage vascular endothelial growth factor release. Am. J. Physiol. Circ. Physiol. 2001, 280, H2357–H2363. [Google Scholar] [CrossRef]
  45. Sundaresan, M.; Yu, Z.-X.; Ferrans, V.J.; Irani, K.; Finkel, T. Requirement for generation of H2O2 for platelet-derived growth factor signal transduction. Science 1995, 270, 296–299. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  46. Gordillo, G.M.; Sen, C.K. Revisiting the essential role of oxygen in wound healing. Am. J. Surg. 2003, 186, 259–263. [Google Scholar] [CrossRef]
  47. Sen, C.K. The general case for redox control of wound repair. Wound Repair Regen. 2003, 11, 431–438. [Google Scholar] [CrossRef] [PubMed]
  48. Roy, S.; Khanna, S.; Sen, C.K. Redox regulation of the VEGF signaling path and tissue vascularization: Hydrogen peroxide, the common link between physical exercise and cutaneous wound healing. Free Radic. Biol. Med. 2008, 44, 180–192. [Google Scholar] [CrossRef]
  49. Roy, S.; Khanna, S.; Rink, C.; Biswas, S.; Sen, C.K. Characterization of the acute temporal changes in excisional murine cutaneous wound inflammation by screening of the wound-edge transcriptome. Physiol. Genom. 2008, 34, 162–184. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  50. Lavery, L.A.; Killeen, A.L.; Farrar, D.; Akgul, Y.; Crisologo, P.A.; Malone, M.; Davis, K.E. The effect of continuous diffusion of oxygen treatment on cytokines, perfusion, bacterial load, and healing in patients with diabetic foot ulcers. Int. Wound J. 2020, 17, 1986–1995. [Google Scholar] [CrossRef]
  51. Hunt, T.K.; Zederfeldt, B.; Goldstick, T.K. Oxygen and healing. Am. J. Surg. 1969, 118, 521–525. [Google Scholar] [CrossRef]
  52. Prockop, D.; Kivirikko, K.; Tuderman, L.; Guzman, N. The biosynthesis of collagen and its disorders (part 1). N. Engl. J. Med. 1979, 301, 13–23. [Google Scholar] [CrossRef]
  53. Prockop, D.; Kivirikko, K.; Tuderman, L.; Guzman, N. The biosynthesis of collagen and its disorder (part 2). N. Engl. J. Med. 1979, 301, 77–85. [Google Scholar] [CrossRef] [PubMed]
  54. Jonsson, K.; Jensen, J.; Goodson, W.; Scheuenstuhl, H.; West, J.; Hopf, H.; Hunt, T. Tissue oxygenation, anemia, and perfusion in relation to wound healing in surgical patients. Ann. Surg. 1991, 214, 605–613. [Google Scholar] [CrossRef] [PubMed]
  55. Niinikoski, J. Effect of oxygen supply on wound healing and formation of experimental granulation tissue. Acta Physiol. Scand. Suppl. 1969, 334, 1–72. [Google Scholar]
  56. Hunt, T.K.; Pai, M.P. The effect of varying ambient oxygen tensions on wound metabolism and collagen synthesis. Surg. Gynecol. Obstet. 1972, 135, 561–567. [Google Scholar] [PubMed]
  57. Stephens, F.O.; Hunt, T.K. Effect of Changes in Inspired Oxygen and Carbon Dioxide Tensions on Wound Tensile Strength. Ann. Surg. 1971, 173, 515–519. [Google Scholar] [CrossRef]
  58. Hutton, J.J.; Tappel, A.; Udenfriend, S. Cofactor and substrate requirements of collagen proline hydroxylase. Arch. Biochem. Biophys. 1967, 118, 231–240. [Google Scholar] [CrossRef]
  59. Myllyla, R.; Tuderman, L.; Kivirikko, K. Mechanism of the prolyl hydroxlase reaction. 2. Kinetic analysis of the reaction se-quence. Eur. J. Biochem. 1977, 80, 349–357. [Google Scholar] [CrossRef]
  60. Asmis, R.; Qiao, M.; Zhao, Q. Low-Flow Oxygenation of Full-Excisional Skin Wounds on Diabetic Mice Improves Wound Heal-ing by Accelerating Wound Closure and Reepithelialization. Int. Wound J. 2010, 7, 349–357. [Google Scholar] [CrossRef]
  61. Hartmann, M.; Jönsson, K.; Zederfeldt, B. Effect of tissue perfusion and oxygenation on accumulation of collagen in healing wounds. Randomized study in patients after major abdominal operations. Eur. J. Surg. 1992, 158, 521–526. [Google Scholar]
  62. Hopf, H.W.; Gibson, J.J.; Angeles, A.P.; Constant, J.S.; Feng, J.J.; Rollins, M.D.; Zamirul Hussain, M.; Hunt, T.K. Hyperoxia and angiogenesis. Wound Repair Regen. 2005, 13, 558–564. [Google Scholar] [CrossRef] [PubMed]
  63. Mussini, E.; Hutton, J.J., Jr.; Udenfriend, S. Collagen proline hydroxylase in wound healing, granuloma formation, scurvy, and growth. Science 1967, 157, 927–929. [Google Scholar] [CrossRef]
  64. Berthod, F.; Germain, L.; Tremblay, N.; Auger, F.A. Extracellular matrix deposition by fibroblasts is necessary to promote capil-lary-like tube formation in vitro. J. Cell. Physiol. 2006, 207, 491–498. [Google Scholar] [CrossRef] [PubMed]
  65. Hunt, T.K.; Aslam, R.S.; Beckert, S.; Wagner, S.; Ghani, Q.P.; Hussain, M.Z.; Roy, S.; Sen, C.K. Aerobically derived lactate stimulates re-vascularization and tissue repair via redox mechanisms. Antioxid. Redox Signal. 2007, 9, 1115–1124. [Google Scholar] [CrossRef] [Green Version]
  66. Knighton, D.R.; Silver, I.A.; Hunt, T.K. Regulation of wound-healing angiogenesis-effect of oxygen gradients and inspired oxygen concentration. Surgery 1981, 90, 262–270. [Google Scholar] [PubMed]
  67. Sheikh, A.Y.; Rollins, M.D.; Hopf, H.W.; Hunt, T.K. Hyperoxia improves microvascular perfusion in a murine wound model. Wound Repair Regen. 2005, 13, 303–308. [Google Scholar] [CrossRef]
  68. Maniscalco, W.M.; Watkins, R.H.; Finkelstein, J.N.; Campbell, M.H. Vascular endothelial growth factor mRNA increases in alveolar epithelial cells during recovery from oxygen injury. Am. J. Respir. Cell Mol. Biol. 1995, 13, 377–386. [Google Scholar] [CrossRef]
  69. Deaton, P.R.; McKellar, C.T.; Culbreth, R.; Veal, C.F.; Cooper, J.A. Hyperoxia stimulates interleukin-8 release from alveolar macrophages and U937 cells: Attenuation by dexamethasone. Am. J. Physiol. Lung Cell. Mol. Physiol. 1994, 267, 187–192. [Google Scholar] [CrossRef]
  70. Darrington, R.S.; Godden, D.J.; Park, M.S.; Ralston, S.H.; Wallace, H.M. The effect of hyperoxia on the expression of cytokine mRNA in endothelial cells. Biochem. Soc. Trans. 1997, 25, 292S. [Google Scholar] [CrossRef] [Green Version]
  71. Sheikh, A.Y.; Gibson, J.J.; Rollins, M.D.; Hopf, H.W.; Hussain, Z.; Hunt, T.K. Effect of Hyperoxia on Vascular Endothelial Growth Factor Levels in a Wound Model. Arch. Surg. 2000, 135, 1293–1297. [Google Scholar] [CrossRef]
  72. Shenberger, J.S.; Zhang, L.; Powell, R.J.; Barchowsky, A. Hyperoxia enhances VEGF release from A549 cells via post-transcriptional processes. Free Radic. Biol. Med. 2007, 43, 844–852. [Google Scholar] [CrossRef] [Green Version]
  73. Gordillo, G.M.; Roy, S.; Khanna, S.; Schlanger, R.; Khandelwal, S.; Phillips, G.; Sen, C.K. Topical oxygen therapy induces vascular endothelial growth factor expression and im-proves closure of clinically presented chronic wounds. Clin. Exp. Pharmacol. Physiol. 2008, 35, 957–964. [Google Scholar] [CrossRef] [Green Version]
  74. Heng, M.C.; Harker, J.; Csathy, G.; Marshall, C.; Brazier, J.; Sumampong, S.; Paterno Gomez, E. Angiogenesis in necrotic ulcers treated with hyperbaric oxygen. Ostomy Wound Manag. 2000, 46, 18–28. [Google Scholar]
  75. Brown, J.R.; Goldblatt, D.; Buddle, J.; Morton, L.; Thrasher, A.J. Diminished production of anti-inflammatory mediators during neutrophil apoptosis and macrophage phagocytosis in chronic granulomatous disease (CGD). J. Leukoc. Biol. 2003, 73, 591–599. [Google Scholar] [CrossRef]
  76. Babior, B.M. Oxygen-dependent microbial killing by phagocytes (first of two parts). N. Engl. J. Med. 1978, 298, 659–668. [Google Scholar] [CrossRef] [PubMed]
  77. Rabkin, J.M.; Hunt, T.K. Infection and oxygen. In Problem Wounds: The Role of Oxygen; Davis, J.C., Hunt, T.K., Eds.; Elsevier: New York, NY, USA, 1988; pp. 1–16. [Google Scholar]
  78. Cianci, P. Advances in the treatment of the diabetic foot: Is there a role for adjunctive hyperbaric oxygen therapy? Wound Repair Regen. 2004, 12, 2–10. [Google Scholar] [CrossRef]
  79. Wattel, F.; Mathieu, D. Oxygen and wound healing. Bull. Acad. Natl. Med. 2005, 189, 853–864. [Google Scholar]
  80. Knighton, D.; Halliday, B.; Hunt, T. Oxygen as an antibiotic: The effect of inspired oxygen on infection. Arch. Surg. 1984, 119, 199–204. [Google Scholar] [CrossRef] [PubMed]
  81. Knighton, D.R.; Halliday, B.; Hunt, T.K. Oxygen as an Antibiotic. Arch. Surg. 1986, 121, 191–195. [Google Scholar] [CrossRef]
  82. Driver, V.R.; Yao, M.; Kantarci, A.; Gu, G.; Park, N.; Hasturk, H. A prospective, randomized clinical study evaluating the effect of transdermal continuous oxygen therapy on biological processes and foot ulcer healing in persons with diabetes mellitus. Ostomy Wound Manag. 2013, 59, 19–26. [Google Scholar]
  83. Weissenbach, M.; Clahsen, T.; Weber, C.; Spitzer, D.; Wirth, D.; Vestweber, D.; Heinrich, P.C.; Schaper, F. Interleukin-6 is a direct mediator of T cell migration. Eur. J. Immunol. 2004, 34, 2895–2906. [Google Scholar] [CrossRef]
  84. Wright, H.; Cross, A.L.; Edwards, S.W.; Moots, R.J. Effects of IL-6 and IL-6 blockade on neutrophil function in vitro and in vivo. Rheumatology 2014, 53, 1321–1331. [Google Scholar] [CrossRef] [Green Version]
  85. Pandit, A.S.; Faldman, D.S. Effect of oxygen treatment and dressing oxygen permeability on wound healing. Wound Repair Re-gen. 1994, 2, 130–137. [Google Scholar] [CrossRef] [PubMed]
  86. Gallagher, K.A.; Liu, Z.-J.; Xiao, M.; Chen, H.; Goldstein, L.J.; Buerk, D.G.; Nedeau, A.; Thom, S.R.; Velazquez, O.C. Diabetic impairments in NO-mediated endothelial progenitor cell mobilization and homing are reversed by hyperoxia and SDF-1α. J. Clin. Investig. 2007, 117, 1249–1259. [Google Scholar] [CrossRef] [Green Version]
  87. Goldstein, L.J.; Gallagher, K.A.; Bauer, S.M.; Bauer, R.J.; Baireddy, V.; Liu, Z.J.; Buerk, D.G.; Thom, S.R.; Velazquez, O.C. Endothelial pro-genitor cell release into circulation is triggered by hyperoxia-induced increases in bone marrow nitric oxide. Stem Cells 2006, 24, 2309–2318. [Google Scholar] [CrossRef] [PubMed]
  88. Gallagher, K.A.; Goldstein, L.J.; Thom, S.R.; Velazquez, O.C. Hyperbaric Oxygen and Bone Marrow–Derived Endothelial Progenitor Cells in Diabetic Wound Healing. Vascular 2006, 14, 328–337. [Google Scholar] [CrossRef]
  89. Thom, S.R.; Bhopale, V.M.; Velazquez, O.C.; Goldstein, L.J.; Thom, L.H.; Buerk, D.G. Stem cell mobilization by hyperbaric oxygen. Am. J. Physiol. Heart Circ. Physiol. 2006, 290, H1378–H1386. [Google Scholar] [CrossRef] [Green Version]
  90. Armstrong, D.G.; Jude, E.B. The Role of Matrix Metalloproteinases in Wound Healing. J. Am. Podiatr. Med. Assoc. 2002, 92, 12–18. [Google Scholar] [CrossRef] [PubMed]
  91. Armstrong, D.G.; Gurtner, G.C. A histologically hostile environment made more hospitable? Nat. Rev. Endocrinol. 2018, 14, 511–512. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Impact of CDO on gene expression for various cytokines and growth factors at each visit Asterisks indicate significant increase from baseline [50]. Adapted with permission.
Figure 1. Impact of CDO on gene expression for various cytokines and growth factors at each visit Asterisks indicate significant increase from baseline [50]. Adapted with permission.
Biomolecules 11 00925 g001
Table 1. Molecular Biomarkers and Clinical Impact in Wound Healing.
Table 1. Molecular Biomarkers and Clinical Impact in Wound Healing.
Growth Factors
IGF-1protein production and cell proliferation and migration
PDGFcell growth and division and chemotaxis
TGF-βangiogenesis, fibroblast proliferation, collagen synthesis and deposition, extracellular matrix (ECM) remodeling, tissue remodeling, granulation tissue stimulant and anti-inflammatory mediator
VEGFangiogenesis and collagen deposition and epithelialization
IGF-1protein production and cell proliferation and migration
cytokines
CXCL8angiogenesis, epithelialization, fibroblast migration and inflammatory mediator
IL-6leukocyte infiltration, angiogenesis, collagen accumulation, anti-inflammatory, granulation tissue stimulant and mitogenic
TNF-αleukocyte recruitment, cell regulator, ECM synthesis and inflammatory mediator
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Oropallo, A.R.; Serena, T.E.; Armstrong, D.G.; Niederauer, M.Q. Molecular Biomarkers of Oxygen Therapy in Patients with Diabetic Foot Ulcers. Biomolecules 2021, 11, 925. https://doi.org/10.3390/biom11070925

AMA Style

Oropallo AR, Serena TE, Armstrong DG, Niederauer MQ. Molecular Biomarkers of Oxygen Therapy in Patients with Diabetic Foot Ulcers. Biomolecules. 2021; 11(7):925. https://doi.org/10.3390/biom11070925

Chicago/Turabian Style

Oropallo, Alisha R., Thomas E. Serena, David G. Armstrong, and Mark Q. Niederauer. 2021. "Molecular Biomarkers of Oxygen Therapy in Patients with Diabetic Foot Ulcers" Biomolecules 11, no. 7: 925. https://doi.org/10.3390/biom11070925

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop