From Wikipedia, the free encyclopedia
Plasma Gelsolin
Crystal structure of the cytoplasmic form of human Gelsolin ( 3FFN​)
Identifiers
SymbolPlasma Gelsolin
Pfam PF00626
Pfam clan CL0092
InterPro IPR007123
SCOP2 1vil / SCOPe / SUPFAM
Available protein structures:
Pfam   structures / ECOD  
PDB RCSB PDB; PDBe; PDBj
PDBsum structure summary
PDB 2FGH​, 1H1V​, 3CIP​, 6QW3​, 1D0N​, 1KCQ​, 1P8X​, 3C15​, 3FFN​, 5DD2​, 5UBO​, 5ZZ0​, 6LJK​, 1C0F​, 1EQY​, 1ESV​, 1RGI​, 1YAG​, 2FH1​, 2FH2​, 2FH3​, 2FH4​, 2LLF​, 3FFK​, 3TU5​, 5FAE​, 5FAF​, 5H3M​, 5H3N​, 5O2Z​, 6JCO​, 6JEG​, 6JEH​, 6LJE​, 6Q9R​, 6Q9Z​, 6QBF​, 1SVY​, 4CBX​, 1D4X​, 1MDU​, 1NLV​, 1NM1​, 1NMD​, 1NPH​, 1P8Z​, 1SOL​, 1YVN​, 2FF3​, 2FF6​, 3A5L​, 3A5M​, 3A5N​, 3A5O​, 3CJB​, 3CJC​, 5MVV​, 1C0G

Plasma gelsolin (pGSN) is an 83 kDa abundant protein constituent of normal plasma and an important component of the innate immune system. The identification of pGSN in Drosophila melanogaster [1] and C. elegans [2] points to an ancient origin early in evolution. [3] Its extraordinary structural conservation reflects its critical regulatory role in multiple essential functions. [4] Its roles include the breakdown of filamentous actin released from dead cells, activation of macrophages, and localization of the inflammatory response. Substantial decreases in plasma levels are observed in acute and chronic infection and injury in both animal models and in humans. Supplementation therapies with recombinant human pGSN have been shown effective in more than 20 animal models.

pGSN has a cytoplasmic isoform (cGSN) known to be an actin-binding protein controlling cytoskeletal dynamics. cGSN is expressed from the same gene, and is identical to pGSN except for its lack of a 24 amino acid N-terminal extension.

History

The cellular isoform of Gelsolin was discovered in 1979 in the lab of Thomas P. Stossel. Its name comes from observed calcium-dependent reversible gel-sol transitions of macrophage cytoplasmic extract. [5] Around the same time a similarly sized plasma protein was discovered and shown to depolymerize actin; it was named Brevin, due to its ability to shorten actin filaments. [6] [7] [8] [9] [10] In 1986 it was demonstrated that Brevin was identical to cellular Gelsolin except for a 24 AA N-terminal extension, and was renamed Plasma Gelsolin. [11]

Structure

A solution phase representation of pGSN in the presence of Ca2+ adapted from 3FFN​ and low-resolution SAXS information. [12] The 24 AA N-terminal extension unique to the plasma isoform was manually added (left, light blue); no structural information for it is known nor represented. Colors represent the six domains of Gelsolin. [13] [14]

Plasma Gelsolin is a 755 AA, 83 kDa plasma protein made up of six "gelsolin domains," each composed of a 5-6 strand β-sheet between one long and one short α-helix. [15] It exhibits a weak homology between domains S1 and S4, S2 and S5, and S3 and S6, and is identical to the cytoplasmic form of the protein except for the addition of a 24 AA N-terminal extension. Additionally a 27 AA N-terminal signal peptide is cleaved prior to pGSN's secretion from the cell. Both forms of the protein are encoded by highly conserved genes on chromosome 9 in humans, but are under the control of different promoters. [11] There is a single disulfide bond formed on the second domain of the plasma protein, [15] there are no documented natural post-translational modifications, and the pI ≈ 6. [16] [17]

Isoforms and mutations

Aside from the cellular form, the only other known isoform is Gelsolin-3, an identical non-secreted protein containing an 11 AA, rather than 24 AA, N-terminal extension. It has been found in brain, testes, and lung oligodendrocytes, and is reportedly involved in myelin remodeling during spiralization around the axon. [18]

Plasma Gelsolin is highly conserved, [4] and its only known mutations are single point mutations. One of several such mutations leads to Finnish Familial Amyloidosis, a disorder in which pGSN becomes more conformationally flexible and susceptible to enzymatic cleavage resulting in accumulation of peptide fragments into amyloid fibrils. D187N/Y is the most common mutation with additional reports of G167R, N184K, P432R, A551P, and Ala7fs in the medical literature. [19] In addition to this several mutations as well as down-regulation of the protein are associated with breast cancer. [20]

Ca2+

At moderate pH in the absence of Ca2+ pGSN is compact and globular. Low pH or the presence of >nM Ca2+ is associated with an elongated structure with greater backbone flexibility. [12] This flexibility exposes the actin binding sites. [13] Since physiological levels of Ca2+ are ~2 mM, pGSN is natively elongated and able to bind to leaked actin from cellular damage.

Functions

Binding

Plasma Gelsolin is a sticky protein known to bind to a number of peptides and proteins: Actin (see: Relationships with actin), [5] [21] [22] Apo-H, [23] , [24] [25] α-Synuclein, [26] Integrin, [23] Tcp-1, [27] Fibronectin, [28] Syntaxin-4, [29] Tropomyosin, [30] fatty acids and phospholipids (see: Binding and inactivation of diverse inflammatory mediators): LPA, [31] [32] [33] LPS (endotoxin), [33] [34] [35] LTA, [35] PAF, [36] S1P, [37] polyphosphoinositides including PIP2; [38] [39] [40] and nucleic acids: Ap3A, [41] ATP, [42] [43] ADP. [44] PIP2, a phospholipid component of cell membranes, competes with ATP and actin for pGSN binding, [45] and will dissociate F-Actin-capped pGSN. [46] [47]

Relationships with actin

Actin toxicity and removal

Actin is the most abundant cellular protein, and its release into extracellular fluid and circulation following cellular injury from disease [4] [48] or injury [49] leads to increased blood viscosity, [4] hindered microcirculation, [50] and activation of platelets. [51] [52] Hemodialysis patients with low levels of pGSN and high levels of actin in blood had markedly higher mortality. [53] Actin is a major component of biofilms that accumulate at local sites of injury and infection, impeding access of host immune components and therapeutics such as antibiotics. Biofilms are particularly pathogenic in the setting of foreign bodies like indwelling catheters and tissue implants. [54]

Actin exchanges between monomeric (G) and filamentous (F) forms according to the concentrations of it, ATP, and cations. [55] pGSN along with Vitamin D-binding protein (DBP) bind and clear monomeric actin. [46] DBP binds with greater affinity to G-actin, leaving pGSN available to sever F-actin. [56] Furthermore, DBP is capable of removing one actin from a 2:1 actin-pGSN complex, restoring its ability to sever F-actin. [57] F-actin, severed and capped by pGSN, is removed by sinusoidal endothelial cells of the liver. [58] pGSN removes 60% of actin trapped in fibrin clots in vitro leading to an increased rate of clot lysis. [59]

Severing, capping, nucleation, and polymerization

Although pGSN is capable of initiating the polymerization of actin through nucleation, its primary relationship with it in blood is depolymerization through filament severing. [4] Actin severing occurs rapidly in the presence of pGSN and Ca2+. [46] pGSN wraps around filaments, non-enzymatically cleaving them. [15] It remains attached, "capping" the barbed/plus end of the severed filament and inducing a torsional twist that is cooperative through its length. [60] [61] Capping has a reported binding affinity <250 pM in the presence of Ca2+ that is substantially weakened in its absence. Capping also blocks further polymerization at the fast growing, barbed end. [62]

While no evidence exists for nucleating/polymerizing of G-actin by pGSN in vivo, the ability of it to do so in vitro is well documented. [63] [64] Actin polymerization is initiated by the production of an actin trimer nucleus. [65] Formation of nuclei is energetically disfavored, but dimers and/or trimers can be catalyzed/stabilized by a number of cellular proteins. [66] In excess of a 2:1 actin:gelsolin stoichiometry and in the presence of Ca2+, gelsolin will bind three actin monomers. [67] A monomer adds to the trimer creating a tetramer that undergoes an internal conversion to an active tetramer witnessed by a concentration-independent lag phase. Subsequent fibrilization proceeds by monomer addition. [68] Gelsolin remains attached to the fast-growing (barbed/plus) end of actin, producing short, slow-growing fibrils. [69]

These actions are similar to those of cytoplasmic form of pGSN, cGSN, which contributes to structural changes of cells through both nucleating/polymerizing and severing/capping. [15]

Amyloid prevention and clearance

pGSN may play an important role in the prevention and management of amyloidosis in several diseases. It is found in complex with in plasma [25] and reported to both inhibit amyloid formation and defibrillize preformed fibrils in vitro. [24] Mice with an Alzheimer's disease model given pGSN showed a 5-fold decrease in progression of Cerebral Amyloid Angiopathy. [70] pGSN has also been found in Lewy Bodies, amyloid containing protein aggregates associated with Parkinson's disease and Dementia with Lewy bodies. [71] [72]

Role in inflammation

Macrophage stimulation

MARCO receptor

Macrophage receptor MARCO is responsible for pathogen recognition and phagocytosis. Macrophages incubated with actin at concentrations consistent with lung injury showed decreased uptake of bacteria. Uptake was restored when actin was administered in the presence of pGSN. [73]

NOS3

NOS3 is an enzyme that is protective against systemic inflammation and myocardial dysfunction. [74] [75] pGSN activates phosphorylation of Ser1177 in NOS3 and Ser473 in Akt. [76] NOS3 is known to be activated by phosphorylation of Akt. [77] Mouse macrophage uptake and killing of bacteria in vitro was enhanced by pGSN, and no significant enhancement was found for NOS3-/- macrophages. In vivo, mice showed 15-fold improvement in bacterial clearance when given pGSN, and no significant enhancement was found for NOS3-/- mice. [76]

Inflammatory mediators

pGSN has been shown to bind to the fatty acid inflammatory mediators LPA, [31] [32] [33] LPS (endotoxin), [33] [34] [35] LTA, [35] PAF, [36] S1P, [37] and polyphosphoinositides including PIP2. [46] [39] [40] Mediators of inflammation, the body's innate healing mechanism, accumulate at the site of the injury to begin the processes of defense and repair, [78] [79] [80] and the depletion of local pGSN allows them to do their work. [81]

See Binding and inactivation of diverse inflammatory mediators

Therapeutic potential

The broad therapeutic potential of pGSN supplementation resides in the fact that the molecule embodies a multifunctional system contributing importantly to innate immunity rather than a pharmacologic intervention with selective and specific activities.

Plasma gelsolin's primary function is to keep inflammation local and enhance the function of the innate immune system. It functions through a pleiotropic mechanism of action; severing toxic filamentous actin ( F-actin), binding inflammatory mediators, and enhancing pathogen clearance. These mechanisms are quite distinct from other anti-inflammatory agents that function as antagonists of individual mediators or inhibitors of specific enzymes, and work to ablate inflammation. Most systemic anti-inflammatory agents also suppress the immune system [82] [83] and often require caution in administration because they increase the risk of infection. [84] Plasma gelsolin is unique in that it has also been demonstrated to enhance the antimicrobial action of macrophages, [73] which engulf and digest cellular debris and pathogens, boosting immunity against both gram positive and gram negative bacterial infections. [76]

Mechanisms of action

Plasma gelsolin plays a central role in the body's innate immune system and is responsible for localizing inflammation—a mechanism so central to species survival that it has been highly conserved by evolution. [4] Experimental and epidemiology data suggest that pGSN performs the role of a buffer or shield that modulates the inflammatory response to injury or infection. [85] The system accomplishes this goal in three key ways described below:

Debridement

Plasma gelsolin binds and severs filamentous actin exposed from cells damaged by injury, [6] [7] [86] including both infectious and sterile injury. Actin has been reported to activate platelets, [52] interfere with fibrinolysis, [59] [87] damage endothelial cells, [88] and to function as a danger signal ( DAMP). [89] Administration of large quantities of filamentous actin to rats resulted in lethal pulmonary hemorrhage and thrombosis. [50]

Another key “toxicity” of exposed actin is the fact that it is a major component of biofilms that accumulate at local sites of injury and infection, and that it impedes the access of host immune components and therapeutics such as antibiotics. [54] [90] Biofilms are particularly pathogenic in the setting of foreign bodies like indwelling catheters and tissue implants. [54] As a result of actin exposure at the local site of injury, the local level of plasma gelsolin around the site of the injury initially becomes depleted as it “debrides” the local involved site. [36] Mediators of inflammation, the body's innate healing mechanism, accumulate at the site of the injury to begin the processes of defense and repair, and the depletion of local plasma gelsolin allows them to do their work. [36] While local pGSN levels are depressed, the presence of this abundant protein in the circulation ensures that the inflammatory process stays local, and that stores of plasma gelsolin are available to address further injury so that the overall immune response remains intact.

Augmentation of macrophage antimicrobial activity

pGSN has antimicrobial activity in vitro and in vivo. Administration of pGSN subcutaneously or by inhalation to mice challenged with lethal inocula of S. pneumoniae or even more lethal combinations of influenza virus and bacteria markedly diminished the number of viable bacteria in the animals’ airways and significantly reduced mortality. The number of inflammation-inducing neutrophils was also considerably reduced, presumably as a result of enhanced bacterial clearance. This is true for contemporaneous or delayed administration of recombinant pGSN. [76] [91]

A basis of pGSN's antimicrobial action is that it enhances the ability of cultivated lung macrophages to ingest gram positive and gram negative bacteria. This has been demonstrated in vitro. [76] Improved phagocytosis is the product of pGSN debriding actin bound to macrophage scavenger receptors preventing their function. [73] pGSN also increases the ability of macrophages to kill ingested microorganisms by inducing macrophage nitric oxide synthase activity. [76]

Binding and inactivation of diverse inflammatory mediators

pGSN binds to a number of inflammatory mediators and signaling agents. Binding to LPA occurs at the same site on the molecule that ligates actin and interacts with polyphosphoinositides. [31] Subsequent studies showed that gelsolin alters the effector function of LPA's receptor binding. [32] [36] Binding to inflammatory mediators, and in some cases inhibition of their effector function, has been shown for platelet-activating factor, [36] lipopolysaccharide endotoxin, [34] sphingosine-1-phosphate, [35] and lipoteichoic acid [37] and small molecule purinergic agonists including ATP and ADP. [41] [43] [44] [42] The binding of pGSN to Alzheimer peptide has also been well documented. [24] [25] [92]

Role of mediators which bind to plasma gelsolin
Mediator Role
LPA [31] [32] [33] A phospholipid derivative that can act as a signaling molecule and activates G protein coupled receptors. It has been associated with cell proliferation.
LPS/endotoxin [33] [34] [35] Found in the outer membrane of Gram-negative bacteria, it elicits a strong immune response in animals.
PAF [36] A potent phospholipid activator and mediator of many leukocyte functions, including platelet aggregation, inflammation, and anaphylaxis. It is produced in response to specific stimuli by a variety of cell types, including neutrophils, basophils, platelets, and endothelial cells.
[24] [25] A peptide of 36–43 amino acids that is the main constituent of amyloid plaques in the brains of Alzheimer's disease patients.
LTA [35] A major constituent of the cell wall of Gram-positive bacteria able to stimulate a specific immune response in animals.
S1P [37] A blood borne lipid mediator and major regulator of vascular and immune systems. In the vascular system, S1P regulates angiogenesis, vascular stability, and permeability. In the immune system it is recognized as a major regulator of trafficking of T-cells and B-cells. Inhibition of S1P receptors has been shown to be critical for immunomodulation.

Anti-microbial resistance

Antimicrobial resistance is a global threat that leads to an estimated 700,000 deaths annually with projections of 10M deaths per year and lost economic potential of $100T by 2050. [93] [94] The United States has released a national action plan to combat antibiotic resistant bacteria. [95]

Recombinant pGSN (rhu-pGSN) supplementation alone shows improved survival and decreased bacteria counts in several mouse models. [91] [96] The bactericidal activity of the antimicrobial peptide LL-37 was shown to be inhibited by F-actin. It formed bundles with F-actin in vitro that were dissolved by pGSN, restoring bactericidal activity. Bacteria growth was reduced when pGSN was added cystic fibrosis sputum, which is known to contain F-actin. [97]

When mice were given a penicillin-resistant strain of pneumococcal pneumonia, penicillin had no effect on mortality or morbidity. rhu-pGSN improved both mortality and morbidity on its own, and the combination of rhu-pGSN and penicillin gave further improvement of both suggesting possible synergism. [96]

Levels of the Protein

Plasma gelsolin is produced and secreted by virtually every cell type with muscle contributing the largest amount. [98] At normal levels of >200 mg/L, it is a highly abundant protein in the circulation. [99]

Decreased levels are often associated with ill health and disease. [85] [100] A growing list of insults showing loss of pGSN includes pneumonia, [101] sepsis, [102] SIRS, [103] traumatic brain injury, [104] autoimmune diseases, [105] chronic kidney disease, [53] [106] HIV-1 disease, [107] tick-borne encephalitis and Lyme, [108] malaria, [109] [110] hepatitis, [111] burns, [112] [113] multiple organ dysfunction syndrome, [112] trauma associated with injury [114] or surgery, [106] bone marrow transplantation, [115] and multiple sclerosis. [116] Severely depleted levels (<150 mg/L) strongly correlate with the onset of systemic inflammatory dysregulation and predict increased morbidity and mortality across a broad spectrum of clinical presentations in the critical care setting. The magnitude of decline in pGSN correlates with the likelihood of mortality in seriously ill patients. [53] [106] [117]

Mediators of inflammation, the body's innate healing mechanism, accumulate at the site of the injury to begin the processes of defense and repair, [78] [79] [80] and the depletion of local plasma gelsolin allows them to do their work. [81] As a result of actin exposure at the local site of injury, the local level of plasma gelsolin around the site of the injury initially becomes depleted as it “debrides” the local involved site (see: Debridement). While local pGSN levels are depressed, the presence of this abundant protein in the circulation ensures that the inflammatory process stays local, [100] and that stores of plasma gelsolin are available to address further injury so that the overall immune response remains intact (see: Binding and inactivation of diverse inflammatory mediators).

Measured levels are higher in serum than plasma due to pGSN's affinity for fibrin. [99]

Animal studies

Human plasma gelsolin has been produced in recombinant form in E. coli (rhu-pGSN), and its efficacy as a therapeutic has been studied in vivo in a number of animal models of inflammatory disease. In models of injury that cause actin release and inflammatory organ damage, pGSN levels consistently drop. In models where gelsolin levels are replenished, adverse outcomes can be prevented. To date, rhu-pGSN has been studied in many independent laboratories providing evidence of efficacy in >20 animal models. Following are descriptions of selected animal studies. All stated results are relative to those of placebo treatments.

Summary of clinical results from selected animal studies
Disease Model Results
influenza mouse Mice given a highly lethal form of influenza show increased survival at 12 day end of study point as well as decreased morbidity and decreased expression of pro-inflammatory genes when rhu-pGSN is administered 3 to 6 days after infection. [118]
pneumococcal pneumonia mouse Mice were given pneumococcal challenge 7 days after being given influenza. Supplementation of endogenous pGSN with rhu-pGSN improved bacterial clearance 15-fold, reduced neutrophilic inflammation, improved recovery of initial weight loss, and showed a dose-dependent improvement on survival. No antibiotics were given, demonstrating pGSN's ability to stimulate the innate immune response. [76]
burn rat Rats receiving 40% body surface area burn showed 90% loss of endogenous pGSN within 12 hours and slowly recovered to almost 50% after 6 days. Intravenous administration of rhu-pGSN partially or totally prevented the burn-associated increase in pulmonary microvascular permeability in a dose-dependent manner. [119] See also [120]
sepsis mouse Mice were intraperitoneally injected with endotoxin (LPS) or subjected to cecal ligation and puncture (CLP) (a small amount of intestinal contents were extracted into the cavity and the wound was sutured). Endogenous pGSN levels dropped to 50% post-challenge. Survival substantially improved with rhu-pGSN treatment in both groups: LPS study, 90% vs 0%; CLP study: 30% vs 0%. [121]
sepsis rat Relative to a previous mouse study [121] a smaller dosage of rhu-pGSN decreased morbidity in a double CLP sepsis model relative to sham treatments. The dosage was effective in intraveneous and subcutaneous injections, but less so with intraperitoneal injection (qualitative but not statistically significant) despite the latter being the site of injury. This evidenced the need for systemic availability of pGSN for recovery. [122]
Acute respiratory distress syndrome mouse Mice were subjected to 95% O2 for 72 hr and treated with rhu-pGSN after 24 and 48 hr. Hyperoxia produced severe diffuse congestion and edema with hemorrhage visible in lung histopathology, 70% reduction in endogenous pGSN, and an influx of neutrophils. Treatment with rhu-pGSN led to a 23% decrease in the authors' histpathological score, 65% decrease in BAL fluid neutrophil count, and a 29% reduction in an overall acute lung injury score. [123]
stroke rat Researchers induced middle cerebral artery occlusion with a direct injection of Endothelin 1, a vasoconstrictor. Animals treated with pGSN at the site of injury exhibited 50% infarction area, >2x use of both forepaws during exploration, and a decrease in whisker-stimulated reaction time (9 s, pGSN treated; 19 s untreated; 1 s healthy rat). [124]
multiple sclerosis mouse Mice with experimental autoimmune encephalomyelitis show decreased levels of pGSN in blood and increased levels in the brain. All rhu-pGSN-treated mice survived whereas 60% of control died within 30 days. Rhu-pGSN mice scored significantly better on clinical scores, smaller brain lesions imaged by MRI, less extra-cellular actin, and decreased myeloperoxidase activity. [125]
Alzheimer's mouse Two models of Alzheimer's were tested. Treatment mice that were tail-injected with a plasmid encoding human pGSN showed reduction in 42 in brain tissue, decreased amyloid, and increased concentration of microglia. [126] See also [127]
radiation mouse Mice irradiated with 137Cs γ-rays show a 50-75% decrease in endogenous levels of pGSN. Bleeding is a common consequence of heavy radiation exposure. Administration of rhu-pGSN improved clotting indices in later, but not middle, phases of recovery. Rhu-pGSN improved GSH and MDA oxidative stress indices. [128]
pain and inflammation mouse Intraperitoneal injection of acetic acid causes a pain response quantified by writhing. [129] Both rhu-pGSN and diclofenac sodium (DS), a standard analgesic drug, caused ~55% reduction in writhing. Similarly, tails placed in hot water caused mice to retract them in an average time of 2.3 s. DS increased time to withdrawal from 5.1 to 7.6 s depending on time of drug administration; rhu-pGSN increased time from 2.9 to 5.5 s. Both DS and rhu-pGSN showed significant reductions in swelling associated with paw injection of an inflammatory agent, γ- carrageenan, as well as decreases in measured cytokines TNF-α and IL-6. [130]
diabetes mouse Endogenous levels of pGSN decrease by ~50% with type 2 diabetes(T2D) in both humans and mice. In an oral glucose tolerance test, rhu-pGSN brought blood sugar levels down to levels comparable to sitagliptin, a T2D drug. Daily dose of rhu-pGSN kept blood sugar levels close to normal for the 7 days of treatment. Daily dose of sitagliptin increased levels of endogenous pGSN. [131]

Human Studies

In 2019 BioAegis Therapeutics conducted a Phase Ib/IIa safety study administering recombinant human pGSN to sick patients with community acquired pneumonia; no safety issues were found. [132] A 2020 Phase IIb placebo-controlled efficacy study has been approved for acute severe pneumonia due to COVID-19. The primary outcome was the proportion of patients surviving on Day 14 without mechanical ventilation, vasopressors, or dialysis. Evaluation of efficacy of rhu-pGSN was confounded by high survival rates of both treatment and placebo cohorts resulting from improvements made to the standard of care for COVID pneumonia. [133]

See also

Cytoplasmic gelsolin

Actin

Vitamin D-binding protein

References

  1. ^ Shi, Yigong (2004-08-01). "Caspase activation, inhibition, and reactivation: A mechanistic view". Protein Science. 13 (8): 1979–1987. doi: 10.1110/ps.04789804. ISSN  0961-8368. PMC  2279816. PMID  15273300.
  2. ^ Klaavuniemi, Tuula; Yamashiro, Sawako; Ono, Shoichiro (2008-09-19). "Caenorhabditis elegans Gelsolin-like Protein 1 Is a Novel Actin Filament-severing Protein with Four Gelsolin-like Repeats". Journal of Biological Chemistry. 283 (38): 26071–26080. doi: 10.1074/jbc.M803618200. ISSN  0021-9258. PMC  2533794. PMID  18640981.
  3. ^ Archer, Stuart K.; Claudianos, Charles; Campbell, Hugh D. (2005). "Evolution of the gelsolin family of actin-binding proteins as novel transcriptional coactivators". BioEssays. 27 (4): 388–396. doi: 10.1002/bies.20200. ISSN  1521-1878. PMID  15770676. S2CID  40585071.
  4. ^ a b c d e f Lee, William M.; Galbraith, Robert M. (1992-05-14). "The Extracellular Actin-Scavenger System and Actin Toxicity". New England Journal of Medicine. 326 (20): 1335–1341. doi: 10.1056/NEJM199205143262006. ISSN  0028-4793. PMID  1314333.
  5. ^ a b Yin, Helen L.; Stossel, Thomas P. (18 October 1979). "Control of cytoplasmic actin gel-sol transformation by gelsolin, a calcium-dependent regulatory protein". Nature. 281 (5732): 583–586. Bibcode: 1979Natur.281..583Y. doi: 10.1038/281583a0. PMID  492320. S2CID  4250013. Retrieved 13 February 2020.
  6. ^ a b Chaponnier, C.; Borgia, R.; Rungger-Brändle, E.; Weil, R.; Gabbiani, G. (1979-08-15). "An actin-destabilizing factor is present in human plasma". Experientia. 35 (8): 1039–1041. doi: 10.1007/bf01949928. ISSN  0014-4754. PMID  477868. S2CID  21319139.
  7. ^ a b Norberg, Renee; Thorstensson, Rigmor; Utter, Goran; Fagraeus, Astrid (1979-10-15). "F-Actin-Depolymerizing Activity of Human Serum". European Journal of Biochemistry. 100 (2): 575–583. doi: 10.1111/j.1432-1033.1979.tb04204.x. ISSN  0014-2956. PMID  389627.
  8. ^ Harris, H.E.; Bamburg, J.R.; Weeds, A.G. (1980-11-17). "Actin filament disassembly in blood plasma". FEBS Letters. 121 (1): 175–177. doi: 10.1016/0014-5793(80)81291-9. ISSN  0014-5793. PMID  6893965. S2CID  30794630.
  9. ^ Harris, H.E.; Gooch, J. (1981-01-12). "An actin depolymerizing protein from pig plasma". FEBS Letters. 123 (1): 49–53. doi: 10.1016/0014-5793(81)80017-8. ISSN  0014-5793. PMID  6894126. S2CID  27405593.
  10. ^ Harris, D. A.; Schwartz, J. H. (1981-11-01). "Characterization of brevin, a serum protein that shortens actin filaments". Proceedings of the National Academy of Sciences. 78 (11): 6798–6802. Bibcode: 1981PNAS...78.6798H. doi: 10.1073/pnas.78.11.6798. ISSN  0027-8424. PMC  349138. PMID  6947253.
  11. ^ a b Kwiatkowski, D. J.; Stossel, T. P.; Orkin, S. H.; Mole, J. E.; Colten, H. R.; Yin, H. L. (1986-10-02). "Plasma and cytoplasmic gelsolins are encoded by a single gene and contain a duplicated actin-binding domain". Nature. 323 (6087): 455–458. Bibcode: 1986Natur.323..455K. doi: 10.1038/323455a0. ISSN  0028-0836. PMID  3020431. S2CID  4356162.
  12. ^ a b Ashish (31 August 2007). "Global Structure Changes Associated with Ca2+ Activation of Full-length Human Plasma Gelsolin" (PDF). J Biol Chem. 282 (35): 25884–25892. doi: 10.1074/jbc.M702446200. PMID  17604278. S2CID  25974945. Retrieved 12 February 2020.
  13. ^ a b Burtnick, Leslie D.; Koepf, Edward K.; Grimes, Jonathan; Jones, E. Yvonne; Stuart, David I.; McLaughlin, Paul J.; Robinson, Robert C. (22 August 1997). "The Crystal Structure of Plasma Gelsolin: Implications for Actin Severing, Capping, and Nucleation". Cell. 90 (4): 661–670. doi: 10.1016/s0092-8674(00)80527-9. PMID  9288746. S2CID  11112433. Retrieved 12 February 2020.
  14. ^ Nag, Shalini; Ma, Qing; Wang, Hui; Chumnarnsilpa, Sakesit; Lee, Wei Lin; Larsson, Mårten; Kannan, Balakrishnan; Hernandez-Valladares, Maria; Burtnick, Leslie D.; Robinson, Robert C. (7 July 2009). "Ca2+ binding by domain 2 plays a critical role in the activation and stabilization of gelsolin" (PDF). PNAS. 106 (33): 13713–13718. Bibcode: 2009PNAS..10613713N. doi: 10.1073/pnas.0812374106. PMC  2720848. PMID  19666512. Retrieved 12 February 2020.
  15. ^ a b c d Nag, Shalini; Larsson, Mårten; Robinson, Robert C.; Burtnick, Leslie D. (2013-06-10). "Gelsolin: The tail of a molecular gymnast: Gelsolin Superfamily Proteins". Cytoskeleton. 70 (7): 360–384. doi: 10.1002/cm.21117. ISSN  1949-3584. PMID  23749648.
  16. ^ Yin, H. L.; Kwiatkowski, D. J.; Mole, J. E.; Cole, F. S. (1984-04-25). "Structure and biosynthesis of cytoplasmic and secreted variants of gelsolin". Journal of Biological Chemistry. 259 (8): 5271–5276. doi: 10.1016/S0021-9258(17)42985-1. ISSN  0021-9258. PMID  6325429. Retrieved 2020-03-02.
  17. ^ Moon, Myeong Hee; Kang, Duk Jin (2013-11-19), Apparatus for protein separation using capillary isoelectric focusing—hollow fiber flow field flow fractionation and method thereof, retrieved 2020-03-02
  18. ^ Vouyiouklis, Demetrius A.; Brophy, Peter J. (2002-11-18). "A Novel Gelsolin Isoform Expressed by Oligodendrocytes in the Central Nervous System". Journal of Neurochemistry. 69 (3): 995–1005. doi: 10.1046/j.1471-4159.1997.69030995.x. ISSN  0022-3042. PMID  9282921. S2CID  44552710.
  19. ^ Zorgati, Habiba; Larsson, Mårten; Ren, Weitong; Sim, Adelene Y. L.; Gettemans, Jan; Grimes, Jonathan M.; Li, Wenfei; Robinson, Robert C. (2019-07-09). "The role of gelsolin domain 3 in familial amyloidosis (Finnish type)". Proceedings of the National Academy of Sciences. 116 (28): 13958–13963. Bibcode: 2019PNAS..11613958Z. doi: 10.1073/pnas.1902189116. ISSN  0027-8424. PMC  6628662. PMID  31243148.
  20. ^ Baig, Ruqia Mehmood; Mahjabeen, Ishrat; Sabir, Maimoona; Masood, Nosheen; Ali, Kashif; Malik, Faraz Arshad; Kayani, Mahmood Akhtar (2013). "Mutational Spectrum of Gelsolin and Its Down Regulation Is Associated with Breast Cancer". Disease Markers. 34 (2): 71–80. doi: 10.1155/2013/795410. ISSN  0278-0240. PMC  3809971. PMID  23324580.
  21. ^ Edgar, Alasdair John (1990-08-01). "Gel electrophoresis of native gelsolin and gelsolin-actin complexes". Journal of Muscle Research and Cell Motility. 11 (4): 323–330. doi: 10.1007/BF01766670. ISSN  0142-4319. PMID  2174905. S2CID  11355042.
  22. ^ Burtnick, Leslie D; Urosev, Dunja; Irobi, Edward; Narayan, Kartik; Robinson, Robert C (2004-07-21). "Structure of the N-terminal half of gelsolin bound to actin: roles in severing, apoptosis and FAF". The EMBO Journal. 23 (14): 2713–2722. doi: 10.1038/sj.emboj.7600280. ISSN  0261-4189. PMC  514944. PMID  15215896.
  23. ^ a b Bohgaki, Miyuki; Matsumoto, Masaki; Atsumi, Tatsuya; Kondo, Takeshi; Yasuda, Shinsuke; Horita, Tetsuya; Nakayama, Keiichi I.; Okumura, Fumihiko; Hatakeyama, Shigetsugu; Koike, Takao (2011-01-24). "Plasma gelsolin facilitates interaction between β2 glycoprotein I and α5β1 integrin". Journal of Cellular and Molecular Medicine. 15 (1): 141–151. doi: 10.1111/j.1582-4934.2009.00940.x. ISSN  1582-1838. PMC  3822501. PMID  19840195.
  24. ^ a b c d Ray, Indrani; Chauhan, Abha; Wegiel, Jerzy; Chauhan, Ved P.S. (2000-01-24). "Gelsolin inhibits the fibrillization of amyloid beta-protein, and also defibrillizes its preformed fibrils". Brain Research. 853 (2): 344–351. doi: 10.1016/S0006-8993(99)02315-X. ISSN  0006-8993. PMID  10640633. S2CID  41363612.
  25. ^ a b c d Chauhan, Ved P S; Ray, Indrani; Chauhan, Abha; Wisniewski, Henryk M (1999). "Binding of Gelsolin, a Secretory Protein, to Amyloid β-Protein". Biochemical and Biophysical Research Communications. 258 (2): 241–6. doi: 10.1006/bbrc.1999.0623. PMID  10329371.
  26. ^ Welander, Hedvig; Bontha, Sai Vineela; Näsström, Thomas; Karlsson, Mikael; Nikolajeff, Fredrik; Danzer, Karin; Kostka, Marcus; Kalimo, Hannu; Lannfelt, Lars; Ingelsson, Martin; Bergström, Joakim (2011-08-19). "Gelsolin co-occurs with Lewy bodies in vivo and accelerates α-synuclein aggregation in vitro". Biochemical and Biophysical Research Communications. 412 (1): 32–38. doi: 10.1016/j.bbrc.2011.07.027. ISSN  0006-291X. PMID  21798243.
  27. ^ Svanström, Andreas; Grantham, Julie (2015-09-12). "The molecular chaperone CCT modulates the activity of the actin filament severing and capping protein gelsolin in vitro". Cell Stress and Chaperones. 21 (1): 55–62. doi: 10.1007/s12192-015-0637-5. ISSN  1355-8145. PMC  4679748. PMID  26364302.
  28. ^ Lind, S. E.; Janmey, P. A. (1984-11-10). "Human plasma gelsolin binds to fibronectin". The Journal of Biological Chemistry. 259 (21): 13262–13266. doi: 10.1016/S0021-9258(18)90687-3. ISSN  0021-9258. PMID  6092370.
  29. ^ Kalwat, Michael A.; Wiseman, Dean A.; Luo, Wei; Wang, Zhanxiang; Thurmond, Debbie C. (2012-01-01). "Gelsolin Associates with the N Terminus of Syntaxin 4 to Regulate Insulin Granule Exocytosis". Molecular Endocrinology. 26 (1): 128–141. doi: 10.1210/me.2011-1112. ISSN  0888-8809. PMC  3248323. PMID  22108804.
  30. ^ Khaitlina, Sofia; Fitz, Helene; Hinssen, Horst (2013-07-11). "The interaction of gelsolin with tropomyosin modulates actin dynamics". The FEBS Journal. 280 (18): 4600–4611. doi: 10.1111/febs.12431. ISSN  1742-4658. PMID  23844991.
  31. ^ a b c d Meerschaert, K. (1998-10-15). "Gelsolin and functionally similar actin-binding proteins are regulated by lysophosphatidic acid". The EMBO Journal. 17 (20): 5923–5932. doi: 10.1093/emboj/17.20.5923. ISSN  1460-2075. PMC  1170920. PMID  9774337.
  32. ^ a b c d Goetzl, Edward J.; Lee, Hsinyu; Azuma, Toshifumi; Stossel, Thomas P.; Turck, Christoph W.; Karliner, Joel S. (2000-05-12). "Gelsolin Binding and Cellular Presentation of Lysophosphatidic Acid". Journal of Biological Chemistry. 275 (19): 14573–14578. doi: 10.1074/jbc.275.19.14573. ISSN  0021-9258. PMID  10799543.
  33. ^ a b c d e f Mintzer, Evan; Sargsyan, Hasmik; Bittman, Robert (2006-01-18). "Lysophosphatidic acid and lipopolysaccharide bind to the PIP2-binding domain of gelsolin". Biochimica et Biophysica Acta (BBA) - Biomembranes. 1758 (1): 85–89. doi: 10.1016/j.bbamem.2005.12.009. ISSN  0005-2736. PMID  16460666.
  34. ^ a b c d Bucki, Robert; Georges, Penelope C.; Espinassous, Quentin; Funaki, Makoto; Pastore, Jennifer J.; Chaby, Richard; Janmey, Paul A. (2005-07-19). "Inactivation of Endotoxin by Human Plasma Gelsolin ". Biochemistry. 44 (28): 9590–9597. doi: 10.1021/bi0503504. ISSN  0006-2960. PMID  16008344.
  35. ^ a b c d e f g Bucki, Robert; Byfield, Fitzroy J.; Kulakowska, Alina; McCormick, Margaret E.; Drozdowski, Wieslaw; Namiot, Zbigniew; Hartung, Thomas; Janmey, Paul A. (2008-10-01). "Extracellular Gelsolin Binds Lipoteichoic Acid and Modulates Cellular Response to Proinflammatory Bacterial Wall Components". The Journal of Immunology. 181 (7): 4936–4944. doi: 10.4049/jimmunol.181.7.4936. ISSN  0022-1767. PMID  18802097.
  36. ^ a b c d e f g Osborn, Teresia M.; Dahlgren, Claes; Hartwig, John H.; Stossel, Thomas P. (2007-04-01). "Modifications of cellular responses to lysophosphatidic acid and platelet-activating factor by plasma gelsolin". American Journal of Physiology. Cell Physiology. 292 (4): –1323–C1330. doi: 10.1152/ajpcell.00510.2006. ISSN  0363-6143. PMID  17135294.
  37. ^ a b c d Bucki, Robert; Kułakowska, Alina; Byfield, Fitzroy J.; Żendzian-Piotrowska, Małgorzata; Baranowski, Marcin; Marzec, Michał; Winer, Jessamine P.; Ciccarelli, Nicholas J.; Górski, Jan; Drozdowski, Wiesław; Bittman, Robert; Janmey, Paul A. (2010-12-01). "Plasma gelsolin modulates cellular response to sphingosine 1-phosphate". American Journal of Physiology. Cell Physiology. 299 (6): –1516–C1523. doi: 10.1152/ajpcell.00051.2010. ISSN  0363-6143. PMC  3006327. PMID  20810916.
  38. ^ Janmey, P. A.; Stossel, T. P. (1987-01-22). "Modulation of gelsolin function by phosphatidylinositol 4,5-bisphosphate". Nature. 325 (6102): 362–364. Bibcode: 1987Natur.325..362J. doi: 10.1038/325362a0. ISSN  0028-0836. PMID  3027569. S2CID  4324043.
  39. ^ a b Lin, Keng-Mean; Wenegieme, Elizabeth; Lu, Pei-Jung; Chen, Ching-Shih; Yin, Helen L. (1997-08-15). "Gelsolin Binding to Phosphatidylinositol 4,5-Bisphosphate Is Modulated by Calcium and pH". Journal of Biological Chemistry. 272 (33): 20443–20450. doi: 10.1074/jbc.272.33.20443. ISSN  0021-9258. PMID  9252353.
  40. ^ a b Sun, Hui-qiao; Lin, Keng-mean; Yin, Helen L. (1997-08-25). "Gelsolin Modulates Phospholipase C Activity In Vivo through Phospholipid Binding". The Journal of Cell Biology. 138 (4): 811–820. doi: 10.1083/jcb.138.4.811. ISSN  0021-9525. PMC  2138049. PMID  9265648.
  41. ^ a b Vartanian, Amalia A (March 2003). "Gelsolin and plasminogen activator inhibitor-1 are Ap3A-binding proteins". The Italian Journal of Biochemistry. 52 (1): 9–16. ISSN  0021-2938. PMID  12833632.
  42. ^ a b Yamamoto, Hideo; Ito, Hiroaki; Nakamura, Hideji; Hayashi, Eijiro; Kishimoto, Susumu; Hashimoto, Tadao; Tagawa, Kunio (1990-10-01). "Human Plasma Gelsolin Binds Adenosine Triphosphate". The Journal of Biochemistry. 108 (4): 505–506. doi: 10.1093/oxfordjournals.jbchem.a123229. ISSN  0021-924X. PMID  1963427. Retrieved 2020-02-28.
  43. ^ a b Urosev, Dunja; Ma, Qing; Tan, Agnes L.C.; Robinson, Robert C.; Burtnick, Leslie D. (2006-03-31). "The Structure of Gelsolin Bound to ATP". Journal of Molecular Biology. 357 (3): 765–772. doi: 10.1016/j.jmb.2006.01.027. ISSN  0022-2836. PMID  16469333.
  44. ^ a b Laham, Lorraine E.; Way, Michael; Yin, Helen L.; Janmey, Paul A. (1995-11-15). "Identification of Two Sites in Gelsolin with Different Sensitivities to Adenine Nucleotides". European Journal of Biochemistry. 234 (1): 1–7. doi: 10.1111/j.1432-1033.1995.001_c.x. ISSN  0014-2956. PMID  8529627.
  45. ^ Szatmári, Dávid; Xue, Bo; Kannan, Balakrishnan; Burtnick, Leslie D.; Bugyi, Beáta; Nyitrai, Miklós; Robinson, Robert C. (2018-08-07). "ATP competes with PIP2 for binding to gelsolin". PLOS ONE. 13 (8). Eugene A. Permyakov (ed.): –0201826. Bibcode: 2018PLoSO..1301826S. doi: 10.1371/journal.pone.0201826. ISSN  1932-6203. PMC  6080781. PMID  30086165.
  46. ^ a b c d Janmey, P. A.; Lind, S. E. (August 1987). "Capacity of human serum to depolymerize actin filaments". Blood. 70 (2): 524–530. doi: 10.1182/blood.V70.2.524.524. ISSN  0006-4971. PMID  3038216.
  47. ^ Janmey, P. A.; Iida, K.; Yin, H. L.; Stossel, T. P. (1987-09-05). "Polyphosphoinositide micelles and polyphosphoinositide-containing vesicles dissociate endogenous gelsolin-actin complexes and promote actin assembly from the fast-growing end of actin filaments blocked by gelsolin". The Journal of Biological Chemistry. 262 (25): 12228–12236. doi: 10.1016/S0021-9258(18)45341-0. ISSN  0021-9258. PMID  3040735.
  48. ^ Erukhimov, Jeffrey A.; Tang, Zi-Lue; Johnson, Bruce A.; Donahoe, Michael P.; Razzack, Jamal A.; Gibson, Kevin F.; Lee, William M.; Wasserloos, Karla J.; Watkins, Simon A.; Pitt, Bruce R. (July 2000). "Actin-Containing Sera From Patients With Adult Respiratory Distress Syndrome Are Toxic to Sheep Pulmonary Endothelial Cells". American Journal of Respiratory and Critical Care Medicine. 162 (1): 288–294. doi: 10.1164/ajrccm.162.1.9806088. ISSN  1073-449X. PMID  10903256. S2CID  23974368.
  49. ^ Martinez-Amat, A (2005-11-01). "Release of α-actin into serum after skeletal muscle damage". British Journal of Sports Medicine. 39 (11): 830–834. doi: 10.1136/bjsm.2004.017566. ISSN  0306-3674. PMC  1725075. PMID  16244192.
  50. ^ a b Haddad, J G; Harper, K D; Guoth, M; Pietra, G G; Sanger, J W (February 1990). "Angiopathic consequences of saturating the plasma scavenger system for actin". Proceedings of the National Academy of Sciences of the United States of America. 87 (4): 1381–1385. Bibcode: 1990PNAS...87.1381H. doi: 10.1073/pnas.87.4.1381. ISSN  0027-8424. PMC  53479. PMID  2154744.
  51. ^ Scarborough, Victoria D.; Bradford, Harvey R.; Ganguly, Pankaj (1981-06-16). "Aggregation of platelets by muscle actin. A multivalent interaction model of platelet aggregation by ADP". Biochemical and Biophysical Research Communications. 100 (3): 1314–1319. doi: 10.1016/0006-291X(81)91967-7. ISSN  0006-291X. PMID  6895029.
  52. ^ a b Vasconcellos, Ca; Lind, Se (1993-12-15). "Coordinated inhibition of actin-induced platelet aggregation by plasma gelsolin and vitamin D-binding protein". Blood. 82 (12): 3648–3657. doi: 10.1182/blood.V82.12.3648.bloodjournal82123648 (inactive 2024-04-27). ISSN  0006-4971. PMID  8260702. Retrieved 2020-02-13.{{ cite journal}}: CS1 maint: DOI inactive as of April 2024 ( link)
  53. ^ a b c Lee, Po-Shun; Sampath, Kartik; Karumanchi, S. Ananth; Tamez, Hector; Bhan, Ishir; Isakova, Tamara; Gutierrez, Orlando M.; Wolf, Myles; Chang, Yuchiao; Stossel, Thomas P.; Thadhani, Ravi (2009-04-23). "Plasma Gelsolin and Circulating Actin Correlate with Hemodialysis Mortality". Journal of the American Society of Nephrology. 20 (5): 1140–1148. doi: 10.1681/ASN.2008091008. ISSN  1046-6673. PMC  2678046. PMID  19389844.
  54. ^ a b c Walker, T. S.; Tomlin, K. L.; Worthen, G. S.; Poch, K. R.; Lieber, J. G.; Saavedra, M. T.; Fessler, M. B.; Malcolm, K. C.; Vasil, M. L.; Nick, J. A. (2005-06-01). "Enhanced Pseudomonas aeruginosa Biofilm Development Mediated by Human Neutrophils". Infection and Immunity. 73 (6): 3693–3701. doi: 10.1128/IAI.73.6.3693-3701.2005. ISSN  0019-9567. PMC  1111839. PMID  15908399.
  55. ^ Kudryashov, Dmitri S.; Reisler, Emil (April 2013). "ATP and ADP Actin States". Biopolymers. 99 (4): 245–256. doi: 10.1002/bip.22155. ISSN  0006-3525. PMC  3670783. PMID  23348672.
  56. ^ Lind, S E; Smith, D B; Janmey, P A; Stossel, T P (1986-09-01). "Role of plasma gelsolin and the vitamin D-binding protein in clearing actin from the circulation". Journal of Clinical Investigation. 78 (3): 736–742. doi: 10.1172/JCI112634. ISSN  0021-9738. PMC  423663. PMID  3018044.
  57. ^ Janmey, Paul A.; Stossel, Thomas P.; Lind, Stuart E. (1986-04-14). "Sequential binding of actin monomers to plasma gelsolin and its inhibition by vitamin D-binding protein". Biochemical and Biophysical Research Communications. 136 (1): 72–79. doi: 10.1016/0006-291X(86)90878-8. ISSN  0006-291X. PMID  3010978.
  58. ^ Herrmannsdoerfer, A. J.; Heeb, G. T.; Feustel, P. J.; Estes, J. E.; Keenan, C. J.; Minnear, F. L.; Selden, L.; Giunta, C.; Flor, J. R.; Blumenstock, F. A. (December 1993). "Vascular clearance and organ uptake of G- and F-actin in the rat". The American Journal of Physiology. 265 (6 Pt 1): –1071–1081. doi: 10.1152/ajpgi.1993.265.6.G1071. ISSN  0002-9513. PMID  8279558.
  59. ^ a b Janmey, P. A.; Lamb, J. A.; Ezzell, R. M.; Hvidt, S.; Lind, S. E. (1992-08-15). "Effects of actin filaments on fibrin clot structure and lysis". Blood. 80 (4): 928–936. doi: 10.1182/blood.V80.4.928.928. ISSN  0006-4971. PMID  1323346.
  60. ^ Orlova, A.; Prochniewicz, E.; Egelman, E. H. (1995-02-03). "Structural dynamics of F-actin: II. Cooperativity in structural transitions". Journal of Molecular Biology. 245 (5): 598–607. doi: 10.1006/jmbi.1994.0049. ISSN  0022-2836. PMID  7844829.
  61. ^ Prochniewicz, Ewa; Zhang, Qingnan; Janmey, Paul A.; Thomas, David D. (August 1996). "Cooperativity in F-Actin: Binding of Gelsolin at the Barbed End Affects Structure and Dynamics of the Whole Filament". Journal of Molecular Biology. 260 (5): 756–766. doi: 10.1006/jmbi.1996.0435. ISSN  0022-2836. PMID  8709153.
  62. ^ Janmey, Paul A.; Chaponnier, Christine; Lind, Stuart E.; Zaner, Ken S.; Stossel, Thomas P.; Yin, Helen L. (July 1985). "Interactions of gelsolin and gelsolin-actin complexes with actin. Effects of calcium on actin nucleation, filament severing, and end blocking". Biochemistry. 24 (14): 3714–3723. doi: 10.1021/bi00335a046. ISSN  0006-2960. PMID  2994715.
  63. ^ Doi, Y.; Frieden, C. (1984-10-10). "Actin polymerization. The effect of brevin on filament size and rate of polymerization". Journal of Biological Chemistry. 259 (19): 11868–11875. doi: 10.1016/S0021-9258(20)71292-5. ISSN  0021-9258. PMID  6480587. Retrieved 2020-02-19.
  64. ^ Brooks, F.J.; Carlsson, A.E. (August 2008). "Actin Polymerization Overshoots and ATP Hydrolysis as Assayed by Pyrene Fluorescence". Biophysical Journal. 95 (3): 1050–1062. Bibcode: 2008BpJ....95.1050B. doi: 10.1529/biophysj.107.123125. ISSN  0006-3495. PMC  2479571. PMID  18390612.
  65. ^ Sept, David; McCammon, J. Andrew (2001-08-01). "Thermodynamics and Kinetics of Actin Filament Nucleation". Biophysical Journal. 81 (2): 667–674. Bibcode: 2001BpJ....81..667S. doi: 10.1016/S0006-3495(01)75731-1. ISSN  0006-3495. PMC  1301543. PMID  11463615.
  66. ^ Qu, Zheng; Silvan, Unai; Jockusch, Brigitte M.; Aebi, Ueli; Schoenenberger, Cora-Ann; Mannherz, Hans Georg (October 2015). "Distinct actin oligomers modulate differently the activity of actin nucleators". FEBS Journal. 282 (19): 3824–3840. doi: 10.1111/febs.13381. ISSN  1742-464X. PMID  26194975.
  67. ^ Edgar, Alasdair John (August 1990). "Gel electrophoresis of native gelsolin and gelsolin-actin complexes". Journal of Muscle Research and Cell Motility. 11 (4): 323–330. doi: 10.1007/BF01766670. ISSN  0142-4319. PMID  2174905. S2CID  11355042.
  68. ^ Lal, A. A.; Korn, E. D.; Brenner, S. L. (1984-07-25). "Rate constants for actin polymerization in ATP determined using cross-linked actin trimers as nuclei". The Journal of Biological Chemistry. 259 (14): 8794–8800. doi: 10.1016/S0021-9258(17)47223-1. ISSN  0021-9258. PMID  6746624.
  69. ^ Janmey, Paul A.; Stossel, Thomas P. (1986-10-01). "Kinetics of actin monomer exchange at the slow growing ends of actin filaments and their relation to the elongation of filaments shortened by gelsolin". Journal of Muscle Research & Cell Motility. 7 (5): 446–454. doi: 10.1007/BF01753587. ISSN  1573-2657. PMID  3025252. S2CID  2644111.
  70. ^ Gregory, Julia L; Prada, Claudia M; Fine, Sara J; Garcia-Alloza, Monica; Betensky, Rebecca A; Arbel-Ornath, Michal; Greenberg, Steven M; Bacskai, Brian J; Frosch, Matthew P (2012). "Reducing Available Soluble A-Amyloid Prevents Progression of Cerebral Amyloid Angiopathy in Transgenic Mice". J Neuropathol Exp Neurol. 71 (11): 1009–17. doi: 10.1097/NEN.0b013e3182729845. PMC  3491571. PMID  23095848.
  71. ^ Leverenz, James B.; Umar, Imran; Wang, Qing; Montine, Thomas J.; McMillan, Pamela J.; Tsuang, Debby W.; Jin, Jinghua; Pan, Catherine; Shin, Jenny; Zhu, David; Zhang, Jing (2007-04-01). "Proteomic identification of novel proteins in cortical lewy bodies". Brain Pathology (Zurich, Switzerland). 17 (2): 139–145. doi: 10.1111/j.1750-3639.2007.00048.x. ISSN  1015-6305. PMC  8095629. PMID  17388944. S2CID  24457175.
  72. ^ Welander, Hedvig; Bontha, Sai Vineela; Näsström, Thomas; Karlsson, Mikael; Nikolajeff, Fredrik; Danzer, Karin; Kostka, Marcus; Kalimo, Hannu; Lannfelt, Lars; Ingelsson, Martin; Bergström, Joakim (2011-07-21). "Gelsolin co-occurs with Lewy bodies in vivo and accelerates α-synuclein aggregation in vitro". Biochemical and Biophysical Research Communications. 412 (1): 32–38. doi: 10.1016/j.bbrc.2011.07.027. ISSN  0006-291X. PMID  21798243.
  73. ^ a b c Ordija, Christine M.; Chiou, Terry Ting-Yu; Yang, Zhiping; Deloid, Glen M.; de Oliveira Valdo, Melina; Wang, Zhi; Bedugnis, Alice; Noah, Terry L.; Jones, Samuel; Koziel, Henry; Kobzik, Lester (2017-06-01). "Free actin impairs macrophage bacterial defenses via scavenger receptor MARCO interaction with reversal by plasma gelsolin". American Journal of Physiology. Lung Cellular and Molecular Physiology. 312 (6): –1018–L1028. doi: 10.1152/ajplung.00067.2017. ISSN  1040-0605. PMC  5495953. PMID  28385809.
  74. ^ Bougaki, Masahiko; Searles, Robert J.; Kida, Kotaro; De Yu, Jia; Buys, Emmanuel S.; Ichinose, Fumito (2010-09-01). "NOS3 protects against systemic inflammation and myocardial dysfunction in murine polymicrobial sepsis". Shock (Augusta, Ga.). 34 (3): 281–290. doi: 10.1097/SHK.0b013e3181cdc327. ISSN  1073-2322. PMC  3774000. PMID  19997049.
  75. ^ Jimenez-Sousa, Ma Angeles; López, Elisabeth; Fernandez-Rodríguez, Amanda; Tamayo, Eduardo; Fernández-Navarro, Pablo; Segura-Roda, Laura; Heredia, María; Gómez-Herreras, José I.; Bustamante, Jesús; García-Gómez, Juan Miguel; Bermejo-Martin, Jesús F.; Resino, Salvador (2012-07-20). "Genetic polymorphisms located in genes related to immune and inflammatory processes are associated with end-stage renal disease: a preliminary study". BMC Medical Genetics. 13 (1): 58. doi: 10.1186/1471-2350-13-58. ISSN  1471-2350. PMC  3412707. PMID  22817530.
  76. ^ a b c d e f g Yang, Zhiping; Chiou, Terry Ting-Yu; Stossel, Thomas P.; Kobzik, Lester (2015-07-01). "Plasma gelsolin improves lung host defense against pneumonia by enhancing macrophage NOS3 function". American Journal of Physiology. Lung Cellular and Molecular Physiology. 309 (1): –11–L16. doi: 10.1152/ajplung.00094.2015. ISSN  1040-0605. PMC  4491512. PMID  25957291.
  77. ^ Chambliss, Ken L.; Shaul, Philip W. (October 2002). "Estrogen Modulation of Endothelial Nitric Oxide Synthase". Endocrine Reviews. 23 (5): 665–686. doi: 10.1210/er.2001-0045. ISSN  0163-769X. PMID  12372846.
  78. ^ a b Zhao, Yutong; Natarajan, Viswanathan (January 2013). "Lysophosphatidic acid (LPA) and its Receptors: Role in Airway Inflammation and Remodeling". Biochimica et Biophysica Acta (BBA) - Molecular and Cell Biology of Lipids. 1831 (1): 86–92. doi: 10.1016/j.bbalip.2012.06.014. ISSN  0006-3002. PMC  3491109. PMID  22809994.
  79. ^ a b Shaw, J. O.; Pinckard, R. N.; Ferrigni, K. S.; McManus, L. M.; Hanahan, D. J. (1981-09-01). "Activation of human neutrophils with 1-O-hexadecyl/octadecyl-2-acetyl-sn-glycerol-3-phosphorylcholine (platelet activating factor)". The Journal of Immunology. 127 (3): 1250–1255. doi: 10.4049/jimmunol.127.3.1250. ISSN  1550-6606. PMID  6267133. S2CID  23647018. Retrieved 2020-03-05.
  80. ^ a b Obinata, Hideru; Hla, Timothy (2012-01-01). "Sphingosine 1-phosphate in coagulation and inflammation". Seminars in Immunopathology. 34 (1): 73–91. doi: 10.1007/s00281-011-0287-3. ISSN  1863-2300. PMC  3237867. PMID  21805322.
  81. ^ a b Piktel, Ewelina; Levental, Ilya; Durnaś, Bonita; Janmey, Paul A.; Bucki, Robert (2018-08-25). "Plasma Gelsolin: Indicator of Inflammation and Its Potential as a Diagnostic Tool and Therapeutic Target". International Journal of Molecular Sciences. 19 (9): 2516. doi: 10.3390/ijms19092516. ISSN  1422-0067. PMC  6164782. PMID  30149613.
  82. ^ Bancos, Simona; Bernard, Matthew P.; Topham, David J.; Phipps, Richard P. (2009). "Ibuprofen and other widely used non-steroidal anti-inflammatory drugs inhibit antibody production in human cells". Cellular Immunology. 258 (1): 18–28. doi: 10.1016/j.cellimm.2009.03.007. ISSN  0008-8749. PMC  2693360. PMID  19345936.
  83. ^ Coutinho, Agnes E.; Chapman, Karen E. (2011-03-15). "The anti-inflammatory and immunosuppressive effects of glucocorticoids, recent developments and mechanistic insights". Molecular and Cellular Endocrinology. 335 (1): 2–13. doi: 10.1016/j.mce.2010.04.005. ISSN  0303-7207. PMC  3047790. PMID  20398732.
  84. ^ Gm, Anstead (1998-10-01). "Steroids, retinoids, and wound healing". Advances in Wound Care: The Journal for Prevention and Healing. 11 (6): 277–285. ISSN  1076-2191. PMID  10326344.
  85. ^ a b Bucki, Robert; Levental, Ilya; Kulakowska, Alina; Janmey, Paul A. (2008-12-01). "Plasma gelsolin: function, prognostic value, and potential therapeutic use". Current Protein & Peptide Science. 9 (6): 541–551. doi: 10.2174/138920308786733912. ISSN  1389-2037. PMID  19075745.
  86. ^ Janmey, P. A.; Lind, S. E. (1987-08-01). "Capacity of human serum to depolymerize actin filaments". Blood. 70 (2): 524–530. doi: 10.1182/blood.V70.2.524.524. ISSN  0006-4971. PMID  3038216.
  87. ^ Stuart E LindS; Smith, Carolyn J (1991-03-15). "Actin Is a Noncompetitive Plasmin Inhibitor" (PDF). Journal of Biological Chemistry. 266 (8): 5273–5278. doi: 10.1016/S0021-9258(19)67783-5. PMID  1848244.
  88. ^ Erukhimov, Jeffrey A.; Tang, Zi-Lue; Johnson, Bruce A.; Donahoe, Michael P.; Razzack, Jamal A.; Gibson, Kevin F.; Lee, William M.; Wasserloos, Karla J.; Watkins, Simon A.; Pitt, Bruce R. (2000-07-01). "Actin-Containing Sera From Patients With Adult Respiratory Distress Syndrome Are Toxic to Sheep Pulmonary Endothelial Cells". American Journal of Respiratory and Critical Care Medicine. 162 (1): 288–294. doi: 10.1164/ajrccm.162.1.9806088. ISSN  1073-449X. PMID  10903256. S2CID  23974368.
  89. ^ Sousa, Caetano Reis e (2017-03-01). "Sensing infection and tissue damage". EMBO Molecular Medicine. 9 (3): 285–288. doi: 10.15252/emmm.201607227. ISSN  1757-4684. PMC  5331196. PMID  28119319.
  90. ^ Parks, Quinn M.; Young, Robert L.; Poch, Katie R.; Malcolm, Kenneth C.; Vasil, Michael L.; Nick, Jerry A. (2009-04-01). "Neutrophil enhancement of Pseudomonas aeruginosa biofilm development: human F-actin and DNA as targets for therapy". Journal of Medical Microbiology. 58 (Pt 4): 492–502. doi: 10.1099/jmm.0.005728-0. ISSN  0022-2615. PMC  2677169. PMID  19273646.
  91. ^ a b Yang, Zhiping; Levinson, Susan; Stossel, Thomas; DiNubile, Mark; Kobzik, Lester (2017-10-04). "Delayed Therapy with Plasma Gelsolin Improves Survival in Murine Pneumococcal Pneumonia". Open Forum Infectious Diseases. 4 (Suppl 1): –474–S475. doi: 10.1093/ofid/ofx163.1215. ISSN  2328-8957. PMC  5630930.
  92. ^ Ji, Lina; Zhao, Xi; Hua, Zichun (2015-01-06). "Potential Therapeutic Implications of Gelsolin in Alzheimer's Disease". Journal of Alzheimer's Disease. 44 (1): 13–25. doi: 10.3233/JAD-141548. ISSN  1875-8908. PMID  25208622.
  93. ^ Hoffman SJ, Outterson K, Røttingen JA, Cars O, Clift C, Rizvi Z, Rotberg F, Tomson G, Zorzet A (February 2015). "An international legal framework to address antimicrobial resistance". Bulletin of the World Health Organization. 93 (2): 66. doi: 10.2471/BLT.15.152710. PMC  4339972. PMID  25883395.
  94. ^ O'Neill, Jim. "Tackling drug-resistant infections globally: final report and recommendations" (PDF). amr-review.org. Review on Antimicrobial Resistance. Retrieved 5 March 2020.
  95. ^ "FACT SHEET: Obama Administration Releases National Action Plan to Combat Antibiotic-Resistant Bacteria". obamawhitehouse.archives.gov. 27 March 2015. Retrieved 5 March 2020.
  96. ^ a b Yang, Zhiping; Bedugnis, Alice; Levinson, Susan; Dinubile, Mark; Stossel, Thomas; Lu, Quan; Kobzik, Lester (2019-09-26). "Delayed Administration of Recombinant Plasma Gelsolin Improves Survival in a Murine Model of Penicillin-Susceptible and Penicillin-Resistant Pneumococcal Pneumonia". The Journal of Infectious Diseases. 220 (9): 1498–1502. doi: 10.1093/infdis/jiz353. ISSN  0022-1899. PMC  6761947. PMID  31287867.
  97. ^ Weiner, Daniel J.; Bucki, Robert; Janmey, Paul A. (June 2003). "The Antimicrobial Activity of the Cathelicidin LL37 Is Inhibited by F-actin Bundles and Restored by Gelsolin". American Journal of Respiratory Cell and Molecular Biology. 28 (6): 738–745. doi: 10.1165/rcmb.2002-0191OC. ISSN  1044-1549. PMID  12600826.
  98. ^ Kwiatkowski, D. J.; Mehl, R.; Izumo, S.; Nadal-Ginard, B.; Yin, H. L. (1988-06-15). "Muscle is the major source of plasma gelsolin". The Journal of Biological Chemistry. 263 (17): 8239–8243. doi: 10.1016/S0021-9258(18)68469-8. ISSN  0021-9258. PMID  2836420.
  99. ^ a b Smith, D. B.; Janmey, P. A.; Herbert, T. J.; Lind, S. E. (August 1987). "Quantitative measurement of plasma gelsolin and its incorporation into fibrin clots". The Journal of Laboratory and Clinical Medicine. 110 (2): 189–195. ISSN  0022-2143. PMID  3036979.
  100. ^ a b Peddada, Nagesh; Sagar, Amin; Ashish; Garg, Renu (February 2012). "Plasma gelsolin: A general prognostic marker of health". Medical Hypotheses. 78 (2): 203–210. doi: 10.1016/j.mehy.2011.10.024. ISSN  0306-9877. PMID  22082609.
  101. ^ Self, Wesley H; Wunderink, Richard G; DiNubile, Mark J; Stossel, Thomas P; Levinson, Susan L; Williams, Derek J; Anderson, Evan J; Bramley, Anna M; Jain, Seema; Edwards, Kathryn M; Grijalva, Carlos G (2019-09-13). "Low Admission Plasma Gelsolin Concentrations Identify Community-acquired Pneumonia Patients at High Risk for Severe Outcomes". Clinical Infectious Diseases. 69 (7): 1218–1225. doi: 10.1093/cid/ciy1049. ISSN  1058-4838. PMC  6743831. PMID  30561561.
  102. ^ Wang, HaiHong; Cheng, BaoLi; Chen, QiXing; Wu, ShuiJing; Lv, Chen; Xie, GuoHao; Jin, Yue; Fang, XiangMing (2008). "Time course of plasma gelsolin concentrations during severe sepsis in critically ill surgical patients". Critical Care. 12 (4): –106. doi: 10.1186/cc6988. ISSN  1364-8535. PMC  2575595. PMID  18706105.
  103. ^ Horváth-Szalai, Zoltán; Kustán, Péter; Mühl, Diána; Ludány, Andrea; Bugyi, Beáta; Kőszegi, Tamás (February 2017). "Antagonistic sepsis markers: Serum gelsolin and actin/gelsolin ratio". Clinical Biochemistry. 50 (3): 127–133. doi: 10.1016/j.clinbiochem.2016.10.018. ISSN  0009-9120. PMID  27823961.
  104. ^ Jin, Yong; Li, Bo-You; Qiu, Ling-Li; Ling, Yuan-Ren; Bai, Zhi-Qiang (2012-10-01). "Decreased plasma gelsolin is associated with 1-year outcome in patients with traumatic brain injury". Journal of Critical Care. 27 (5): 527–1–527.e6. doi: 10.1016/j.jcrc.2012.01.002. ISSN  0883-9441. PMID  22386223. Retrieved 2020-03-04.
  105. ^ Hu, Yl; Li, H.; Li, W. H.; Meng, H. X.; Fan, Y. Z.; Li, W. J.; Ji, Y. T.; Zhao, H.; Zhang, L.; Jin, X. M.; Zhang, F. M. (December 2013). "The value of decreased plasma gelsolin levels in patients with systemic lupus erythematosus and rheumatoid arthritis in diagnosis and disease activity evaluation". Lupus. 22 (14): 1455–1461. doi: 10.1177/0961203313507985. ISSN  1477-0962. PMID  24122723. S2CID  22637925.
  106. ^ a b c Lee, Po-Shun; Drager, Leslie R.; Stossel, Thomas P.; Moore, Francis D.; Rogers, Selwyn O. (March 2006). "Relationship of plasma gelsolin levels to outcomes in critically ill surgical patients". Annals of Surgery. 243 (3): 399–403. doi: 10.1097/01.sla.0000201798.77133.55. ISSN  0003-4932. PMC  1448930. PMID  16495706.
  107. ^ Sinha, Kislay Kumar; Peddada, Nagesh; Jha, Pravin Kumar; Mishra, Anshul; Pandey, Krishna; Das, Vidya Nand Ravi; Ashish; Das, Pradeep (March 2017). "Plasma Gelsolin Level in HIV-1-Infected Patients: An Indicator of Disease Severity". AIDS Research and Human Retroviruses. 33 (3): 254–260. doi: 10.1089/aid.2016.0154. ISSN  0889-2229. PMID  27700141.
  108. ^ Kułakowska, Alina; Zajkowska, Joanna M.; Ciccarelli, Nicholas J.; Mroczko, Barbara; Drozdowski, Wiesław; Bucki, Robert (2011). "Depletion of Plasma Gelsolin in Patients with Tick-Borne Encephalitis and Lyme Neuroborreliosis". Neurodegenerative Diseases. 8 (5): 375–380. doi: 10.1159/000324373. ISSN  1660-2862. PMC  3121545. PMID  21389683.
  109. ^ Huang, S.; Rhoads, S. L.; DiNubile, M. J. (May 1997). "Temporal association between serum gelsolin levels and clinical events in a patient with severe falciparum malaria". Clinical Infectious Diseases. 24 (5): 951–954. doi: 10.1093/clinids/24.5.951. ISSN  1058-4838. PMID  9142799.
  110. ^ Kassa, Fikregabrail Aberra; Shio, Marina Tiemi; Bellemare, Marie-Josée; Faye, Babacar; Ndao, Momar; Olivier, Martin (2011-10-20). "New Inflammation-Related Biomarkers during Malaria Infection". PLOS ONE. 6 (10): e26495. Bibcode: 2011PLoSO...626495K. doi: 10.1371/journal.pone.0026495. ISSN  1932-6203. PMC  3197653. PMID  22028888.
  111. ^ Ito, H.; Kambe, H.; Kimura, Y.; Nakamura, H.; Hayashi, E.; Kishimoto, T.; Kishimoto, S.; Yamamoto, H. (May 1992). "Depression of plasma gelsolin level during acute liver injury". Gastroenterology. 102 (5): 1686–1692. doi: 10.1016/0016-5085(92)91731-i. ISSN  0016-5085. PMID  1314752.
  112. ^ a b Huang, Li-feng; Yao, Yong-ming; Li, Jin-feng; Dong, Ning; Liu, Chen; Yu, Yan; He, Li-xin; Sheng, Zhi-yong (2011-11-01). "Reduction of Plasma Gelsolin Levels Correlates with Development of Multiple Organ Dysfunction Syndrome and Fatal Outcome in Burn Patients". PLOS ONE. 6 (11): –25748. Bibcode: 2011PLoSO...625748H. doi: 10.1371/journal.pone.0025748. ISSN  1932-6203. PMC  3206022. PMID  22069445.
  113. ^ Xianhui, Li; Pinglian, Li; Xiaojuan, Wang; Wei, Chen; Yong, Yang; Feng, Ran; Peng, Sun; Gang, Xue (December 2014). "The association between plasma gelsolin level and prognosis of burn patients". Burns: Journal of the International Society for Burn Injuries. 40 (8): 1552–1555. doi: 10.1016/j.burns.2014.02.020. ISSN  1879-1409. PMID  24690274.
  114. ^ Mounzer, Karam C.; Moncure, Michael; Smith, Yolanda R.; DiNUBILE, Mark J. (November 1999). "Relationship of Admission Plasma Gelsolin Levels to Clinical Outcomes in Patients after Major Trauma". American Journal of Respiratory and Critical Care Medicine. 160 (5): 1673–1681. doi: 10.1164/ajrccm.160.5.9807137. ISSN  1535-4970. PMID  10556139.
  115. ^ DiNubile, Mark J.; Stossel, Thomas P.; Ljunghusen, Olof C.; Ferrara, James L. M.; Antin, Joseph H. (2002-12-15). "Prognostic implications of declining plasma gelsolin levels after allogeneic stem cell transplantation". Blood. 100 (13): 4367–4371. doi: 10.1182/blood-2002-06-1672. ISSN  1528-0020. PMID  12393536. Retrieved 2020-02-12.
  116. ^ Kułakowska, Alina; Ciccarelli, Nicholas J; Wen, Qi; Mroczko, Barbara; Drozdowski, Wiesław; Szmitkowski, Maciej; Janmey, Paul A; Bucki, Robert (December 2010). "Hypogelsolinemia, a disorder of the extracellular actin scavenger system, in patients with multiple sclerosis". BMC Neurology. 10 (1): 107. doi: 10.1186/1471-2377-10-107. ISSN  1471-2377. PMC  2989318. PMID  21040581.
  117. ^ Lee, Po-Shun; Patel, Sanjay R.; Christiani, David C.; Bajwa, Ednan; Stossel, Thomas P.; Waxman, Aaron B. (2008-11-12). "Plasma Gelsolin Depletion and Circulating Actin in Sepsis—A Pilot Study". PLOS ONE. 3 (11): –3712. Bibcode: 2008PLoSO...3.3712L. doi: 10.1371/journal.pone.0003712. ISSN  1932-6203. PMC  2577888. PMID  19002257.
  118. ^ Yang, Zhiping; Bedugnis, Alice; Levinson, Susan; DiNubile, Mark; Stossel, Thomas; Lu, Quan; Kobzik, Lester (2020-02-21). "Delayed administration of recombinant plasma gelsolin improves survival in a murine model of severe influenza". F1000Research. 8: 1860. doi: 10.12688/f1000research.21082.2. ISSN  2046-1402. PMC  6894358. PMID  31824672.
  119. ^ Rothenbach, Patricia A.; Dahl, Benny; Schwartz, Jason J.; O'Keefe, Grant E.; Yamamoto, Masaya; Lee, William M.; Horton, Jureta W.; Yin, Helen L.; Turnage, Richard H. (January 2004). "Recombinant plasma gelsolin infusion attenuates burn-induced pulmonary microvascular dysfunction". Journal of Applied Physiology. 96 (1): 25–31. doi: 10.1152/japplphysiol.01074.2002. ISSN  8750-7587. PMID  12730154.
  120. ^ Zhang, Qing-Hong; Chen, Qi; Kang, Jia-Rui; Liu, Chen; Dong, Ning; Zhu, Xiao-Mei; Sheng, Zhi-Yong; Yao, Yong-Ming (2011-09-21). "Treatment with gelsolin reduces brain inflammation and apoptotic signaling in mice following thermal injury". Journal of Neuroinflammation. 8 (1): 118. doi: 10.1186/1742-2094-8-118. ISSN  1742-2094. PMC  3191361. PMID  21936896.
  121. ^ a b Lee, Po-Shun; Waxman, Aaron B.; Cotich, Kara L.; Chung, Su Wol; Perrella, Mark A.; Stossel, Thomas P. (March 2007). "Plasma gelsolin is a marker and therapeutic agent in animal sepsis*". Critical Care Medicine. 35 (3): 849–855. doi: 10.1097/01.CCM.0000253815.26311.24. ISSN  0090-3493. PMID  17205019. S2CID  21641666. Retrieved 2020-02-12.
  122. ^ Cohen, Taylor S.; Bucki, Robert; Byfield, Fitzroy J.; Ciccarelli, Nicholas J.; Rosenberg, Brenna; DiNubile, Mark J.; Janmey, Paul A.; Margulies, Susan S. (June 2011). "Therapeutic potential of plasma gelsolin administration in a rat model of sepsis". Cytokine. 54 (3): 235–238. doi: 10.1016/j.cyto.2011.02.006. ISSN  1043-4666. PMC  3083472. PMID  21420877.
  123. ^ Christofidou-Solomidou, Melpo; Scherpereel, Arnaud; Solomides, Charalambos C.; Christie, Jason D.; Stossel, Thomas P.; Goelz, Susan; DiNubile, Mark J. (2002-01-01). "Recombinant Plasma Gelsolin Diminishes the Acute Inflammatory Response to Hyperoxia in Mice". Journal of Investigative Medicine. 50 (1): 54–60. doi: 10.2310/6650.2002.33518. ISSN  1081-5589. PMID  11813829. S2CID  1981768. Retrieved 2020-02-24.
  124. ^ Le, Huong T; Hirko, Aaron C; Thinschmidt, Jeffrey S; Grant, Maria; Li, Zhimin; Peris, Joanna; King, Michael A; Hughes, Jeffrey A; Song, Sihong (2011). "The protective effects of plasma gelsolin on stroke outcome in rats". Experimental & Translational Stroke Medicine. 3 (1): 13. doi: 10.1186/2040-7378-3-13. ISSN  2040-7378. PMC  3224589. PMID  22047744.
  125. ^ Kevin Li-Chun, Hsieh; Schob, Stefan; Zeller, Matthias W.G.; Pulli, Benjamin; Ali, Muhammad; Wang, Cuihua; Chiou, Terry Ting-Yu; Tsang, Yuk-Ming; Lee, Po-Shun; Stossel, Thomas P.; Chen, John W. (October 2015). "Gelsolin decreases actin toxicity and inflammation in murine multiple sclerosis". Journal of Neuroimmunology. 287: 36–42. doi: 10.1016/j.jneuroim.2015.08.006. ISSN  0165-5728. PMC  4595933. PMID  26439960.
  126. ^ Hirko, Aaron C; Meyer, Edwin M; King, Michael A; Hughes, Jeffery A (September 2007). "Peripheral Transgene Expression of Plasma Gelsolin Reduces Amyloid in Transgenic Mouse Models of Alzheimer's Disease". Molecular Therapy. 15 (9): 1623–1629. doi: 10.1038/sj.mt.6300253. ISSN  1525-0016. PMID  17609655.
  127. ^ Matsuoka, Yasuji; Saito, Mitsuo; LaFrancois, John; Saito, Mariko; Gaynor, Kate; Olm, Vicki; Wang, Lili; Casey, Evelyn; Lu, Yifan; Shiratori, Chiharu; Lemere, Cynthia; Duff, Karen (2003-01-01). "Novel therapeutic approach for the treatment of Alzheimer's disease by peripheral administration of agents with an affinity to beta-amyloid". The Journal of Neuroscience. 23 (1): 29–33. doi: 10.1523/JNEUROSCI.23-01-00029.2003. ISSN  1529-2401. PMC  6742136. PMID  12514198.
  128. ^ Li, Mingjuan; Cui, Fengmei; Cheng, Ying; Han, Ling; Wang, Jia; Sun, Ding; Liu, Yu-long; Zhou, Ping-kun; Min, Rui (2014-08-28). "Gelsolin: role of a functional protein in mitigating radiation injury". Cell Biochemistry and Biophysics. 71 (1): 389–396. doi: 10.1007/s12013-014-0210-3. ISSN  1559-0283. PMID  25164111. S2CID  942471. Retrieved 2020-02-25.
  129. ^ Gawade, Shivaji P. (2012). "Acetic acid induced painful endogenous infliction in writhing test on mice". Journal of Pharmacology & Pharmacotherapeutics. 3 (4): 348. doi: 10.4103/0976-500X.103699 (inactive 2024-04-27). ISSN  0976-500X. PMC  3543562. PMID  23326113.{{ cite journal}}: CS1 maint: DOI inactive as of April 2024 ( link)
  130. ^ Gupta, Ashok Kumar; Parasar, Devraj; Sagar, Amin; Choudhary, Vikas; Chopra, Bhupinder Singh; Garg, Renu; Ashish; Khatri, Neeraj (2015-08-14). "Analgesic and Anti-Inflammatory Properties of Gelsolin in Acetic Acid Induced Writhing, Tail Immersion and Carrageenan Induced Paw Edema in Mice". PLOS ONE. 10 (8). Prasun K Datta (ed.): –0135558. Bibcode: 2015PLoSO..1035558G. doi: 10.1371/journal.pone.0135558. ISSN  1932-6203. PMC  4537109. PMID  26426535.
  131. ^ Khatri, Neeraj; Sagar, Amin; Peddada, Nagesh; Choudhary, Vikas; Chopra, Bhupinder Singh; Garg, Veena; Garg, Renu; Ashish (2014). "Plasma Gelsolin Levels Decrease in Diabetic State and Increase upon Treatment with F-Actin Depolymerizing Versions of Gelsolin". Journal of Diabetes Research. 2014: 152075. doi: 10.1155/2014/152075. ISSN  2314-6745. PMC  4247973. PMID  25478578.
  132. ^ BioAegis Therapeutics (14 January 2020). "A Phase 1b/2a Study of the Safety and Pharmacokinetics of Rhu-plasma Gelsolin in Hospitalized Subjects With CAP". ClinicalTrials.gov. U.S. National Library of Medicine. Retrieved 24 February 2020.
  133. ^ BioAegis Therapeutics. "Rhu-pGSN for Severe Covid-19 Pneumonia". ClinicalTrials.gov. U.S. National Library of Medicine. Retrieved 16 July 2020.
From Wikipedia, the free encyclopedia
Plasma Gelsolin
Crystal structure of the cytoplasmic form of human Gelsolin ( 3FFN​)
Identifiers
SymbolPlasma Gelsolin
Pfam PF00626
Pfam clan CL0092
InterPro IPR007123
SCOP2 1vil / SCOPe / SUPFAM
Available protein structures:
Pfam   structures / ECOD  
PDB RCSB PDB; PDBe; PDBj
PDBsum structure summary
PDB 2FGH​, 1H1V​, 3CIP​, 6QW3​, 1D0N​, 1KCQ​, 1P8X​, 3C15​, 3FFN​, 5DD2​, 5UBO​, 5ZZ0​, 6LJK​, 1C0F​, 1EQY​, 1ESV​, 1RGI​, 1YAG​, 2FH1​, 2FH2​, 2FH3​, 2FH4​, 2LLF​, 3FFK​, 3TU5​, 5FAE​, 5FAF​, 5H3M​, 5H3N​, 5O2Z​, 6JCO​, 6JEG​, 6JEH​, 6LJE​, 6Q9R​, 6Q9Z​, 6QBF​, 1SVY​, 4CBX​, 1D4X​, 1MDU​, 1NLV​, 1NM1​, 1NMD​, 1NPH​, 1P8Z​, 1SOL​, 1YVN​, 2FF3​, 2FF6​, 3A5L​, 3A5M​, 3A5N​, 3A5O​, 3CJB​, 3CJC​, 5MVV​, 1C0G

Plasma gelsolin (pGSN) is an 83 kDa abundant protein constituent of normal plasma and an important component of the innate immune system. The identification of pGSN in Drosophila melanogaster [1] and C. elegans [2] points to an ancient origin early in evolution. [3] Its extraordinary structural conservation reflects its critical regulatory role in multiple essential functions. [4] Its roles include the breakdown of filamentous actin released from dead cells, activation of macrophages, and localization of the inflammatory response. Substantial decreases in plasma levels are observed in acute and chronic infection and injury in both animal models and in humans. Supplementation therapies with recombinant human pGSN have been shown effective in more than 20 animal models.

pGSN has a cytoplasmic isoform (cGSN) known to be an actin-binding protein controlling cytoskeletal dynamics. cGSN is expressed from the same gene, and is identical to pGSN except for its lack of a 24 amino acid N-terminal extension.

History

The cellular isoform of Gelsolin was discovered in 1979 in the lab of Thomas P. Stossel. Its name comes from observed calcium-dependent reversible gel-sol transitions of macrophage cytoplasmic extract. [5] Around the same time a similarly sized plasma protein was discovered and shown to depolymerize actin; it was named Brevin, due to its ability to shorten actin filaments. [6] [7] [8] [9] [10] In 1986 it was demonstrated that Brevin was identical to cellular Gelsolin except for a 24 AA N-terminal extension, and was renamed Plasma Gelsolin. [11]

Structure

A solution phase representation of pGSN in the presence of Ca2+ adapted from 3FFN​ and low-resolution SAXS information. [12] The 24 AA N-terminal extension unique to the plasma isoform was manually added (left, light blue); no structural information for it is known nor represented. Colors represent the six domains of Gelsolin. [13] [14]

Plasma Gelsolin is a 755 AA, 83 kDa plasma protein made up of six "gelsolin domains," each composed of a 5-6 strand β-sheet between one long and one short α-helix. [15] It exhibits a weak homology between domains S1 and S4, S2 and S5, and S3 and S6, and is identical to the cytoplasmic form of the protein except for the addition of a 24 AA N-terminal extension. Additionally a 27 AA N-terminal signal peptide is cleaved prior to pGSN's secretion from the cell. Both forms of the protein are encoded by highly conserved genes on chromosome 9 in humans, but are under the control of different promoters. [11] There is a single disulfide bond formed on the second domain of the plasma protein, [15] there are no documented natural post-translational modifications, and the pI ≈ 6. [16] [17]

Isoforms and mutations

Aside from the cellular form, the only other known isoform is Gelsolin-3, an identical non-secreted protein containing an 11 AA, rather than 24 AA, N-terminal extension. It has been found in brain, testes, and lung oligodendrocytes, and is reportedly involved in myelin remodeling during spiralization around the axon. [18]

Plasma Gelsolin is highly conserved, [4] and its only known mutations are single point mutations. One of several such mutations leads to Finnish Familial Amyloidosis, a disorder in which pGSN becomes more conformationally flexible and susceptible to enzymatic cleavage resulting in accumulation of peptide fragments into amyloid fibrils. D187N/Y is the most common mutation with additional reports of G167R, N184K, P432R, A551P, and Ala7fs in the medical literature. [19] In addition to this several mutations as well as down-regulation of the protein are associated with breast cancer. [20]

Ca2+

At moderate pH in the absence of Ca2+ pGSN is compact and globular. Low pH or the presence of >nM Ca2+ is associated with an elongated structure with greater backbone flexibility. [12] This flexibility exposes the actin binding sites. [13] Since physiological levels of Ca2+ are ~2 mM, pGSN is natively elongated and able to bind to leaked actin from cellular damage.

Functions

Binding

Plasma Gelsolin is a sticky protein known to bind to a number of peptides and proteins: Actin (see: Relationships with actin), [5] [21] [22] Apo-H, [23] , [24] [25] α-Synuclein, [26] Integrin, [23] Tcp-1, [27] Fibronectin, [28] Syntaxin-4, [29] Tropomyosin, [30] fatty acids and phospholipids (see: Binding and inactivation of diverse inflammatory mediators): LPA, [31] [32] [33] LPS (endotoxin), [33] [34] [35] LTA, [35] PAF, [36] S1P, [37] polyphosphoinositides including PIP2; [38] [39] [40] and nucleic acids: Ap3A, [41] ATP, [42] [43] ADP. [44] PIP2, a phospholipid component of cell membranes, competes with ATP and actin for pGSN binding, [45] and will dissociate F-Actin-capped pGSN. [46] [47]

Relationships with actin

Actin toxicity and removal

Actin is the most abundant cellular protein, and its release into extracellular fluid and circulation following cellular injury from disease [4] [48] or injury [49] leads to increased blood viscosity, [4] hindered microcirculation, [50] and activation of platelets. [51] [52] Hemodialysis patients with low levels of pGSN and high levels of actin in blood had markedly higher mortality. [53] Actin is a major component of biofilms that accumulate at local sites of injury and infection, impeding access of host immune components and therapeutics such as antibiotics. Biofilms are particularly pathogenic in the setting of foreign bodies like indwelling catheters and tissue implants. [54]

Actin exchanges between monomeric (G) and filamentous (F) forms according to the concentrations of it, ATP, and cations. [55] pGSN along with Vitamin D-binding protein (DBP) bind and clear monomeric actin. [46] DBP binds with greater affinity to G-actin, leaving pGSN available to sever F-actin. [56] Furthermore, DBP is capable of removing one actin from a 2:1 actin-pGSN complex, restoring its ability to sever F-actin. [57] F-actin, severed and capped by pGSN, is removed by sinusoidal endothelial cells of the liver. [58] pGSN removes 60% of actin trapped in fibrin clots in vitro leading to an increased rate of clot lysis. [59]

Severing, capping, nucleation, and polymerization

Although pGSN is capable of initiating the polymerization of actin through nucleation, its primary relationship with it in blood is depolymerization through filament severing. [4] Actin severing occurs rapidly in the presence of pGSN and Ca2+. [46] pGSN wraps around filaments, non-enzymatically cleaving them. [15] It remains attached, "capping" the barbed/plus end of the severed filament and inducing a torsional twist that is cooperative through its length. [60] [61] Capping has a reported binding affinity <250 pM in the presence of Ca2+ that is substantially weakened in its absence. Capping also blocks further polymerization at the fast growing, barbed end. [62]

While no evidence exists for nucleating/polymerizing of G-actin by pGSN in vivo, the ability of it to do so in vitro is well documented. [63] [64] Actin polymerization is initiated by the production of an actin trimer nucleus. [65] Formation of nuclei is energetically disfavored, but dimers and/or trimers can be catalyzed/stabilized by a number of cellular proteins. [66] In excess of a 2:1 actin:gelsolin stoichiometry and in the presence of Ca2+, gelsolin will bind three actin monomers. [67] A monomer adds to the trimer creating a tetramer that undergoes an internal conversion to an active tetramer witnessed by a concentration-independent lag phase. Subsequent fibrilization proceeds by monomer addition. [68] Gelsolin remains attached to the fast-growing (barbed/plus) end of actin, producing short, slow-growing fibrils. [69]

These actions are similar to those of cytoplasmic form of pGSN, cGSN, which contributes to structural changes of cells through both nucleating/polymerizing and severing/capping. [15]

Amyloid prevention and clearance

pGSN may play an important role in the prevention and management of amyloidosis in several diseases. It is found in complex with in plasma [25] and reported to both inhibit amyloid formation and defibrillize preformed fibrils in vitro. [24] Mice with an Alzheimer's disease model given pGSN showed a 5-fold decrease in progression of Cerebral Amyloid Angiopathy. [70] pGSN has also been found in Lewy Bodies, amyloid containing protein aggregates associated with Parkinson's disease and Dementia with Lewy bodies. [71] [72]

Role in inflammation

Macrophage stimulation

MARCO receptor

Macrophage receptor MARCO is responsible for pathogen recognition and phagocytosis. Macrophages incubated with actin at concentrations consistent with lung injury showed decreased uptake of bacteria. Uptake was restored when actin was administered in the presence of pGSN. [73]

NOS3

NOS3 is an enzyme that is protective against systemic inflammation and myocardial dysfunction. [74] [75] pGSN activates phosphorylation of Ser1177 in NOS3 and Ser473 in Akt. [76] NOS3 is known to be activated by phosphorylation of Akt. [77] Mouse macrophage uptake and killing of bacteria in vitro was enhanced by pGSN, and no significant enhancement was found for NOS3-/- macrophages. In vivo, mice showed 15-fold improvement in bacterial clearance when given pGSN, and no significant enhancement was found for NOS3-/- mice. [76]

Inflammatory mediators

pGSN has been shown to bind to the fatty acid inflammatory mediators LPA, [31] [32] [33] LPS (endotoxin), [33] [34] [35] LTA, [35] PAF, [36] S1P, [37] and polyphosphoinositides including PIP2. [46] [39] [40] Mediators of inflammation, the body's innate healing mechanism, accumulate at the site of the injury to begin the processes of defense and repair, [78] [79] [80] and the depletion of local pGSN allows them to do their work. [81]

See Binding and inactivation of diverse inflammatory mediators

Therapeutic potential

The broad therapeutic potential of pGSN supplementation resides in the fact that the molecule embodies a multifunctional system contributing importantly to innate immunity rather than a pharmacologic intervention with selective and specific activities.

Plasma gelsolin's primary function is to keep inflammation local and enhance the function of the innate immune system. It functions through a pleiotropic mechanism of action; severing toxic filamentous actin ( F-actin), binding inflammatory mediators, and enhancing pathogen clearance. These mechanisms are quite distinct from other anti-inflammatory agents that function as antagonists of individual mediators or inhibitors of specific enzymes, and work to ablate inflammation. Most systemic anti-inflammatory agents also suppress the immune system [82] [83] and often require caution in administration because they increase the risk of infection. [84] Plasma gelsolin is unique in that it has also been demonstrated to enhance the antimicrobial action of macrophages, [73] which engulf and digest cellular debris and pathogens, boosting immunity against both gram positive and gram negative bacterial infections. [76]

Mechanisms of action

Plasma gelsolin plays a central role in the body's innate immune system and is responsible for localizing inflammation—a mechanism so central to species survival that it has been highly conserved by evolution. [4] Experimental and epidemiology data suggest that pGSN performs the role of a buffer or shield that modulates the inflammatory response to injury or infection. [85] The system accomplishes this goal in three key ways described below:

Debridement

Plasma gelsolin binds and severs filamentous actin exposed from cells damaged by injury, [6] [7] [86] including both infectious and sterile injury. Actin has been reported to activate platelets, [52] interfere with fibrinolysis, [59] [87] damage endothelial cells, [88] and to function as a danger signal ( DAMP). [89] Administration of large quantities of filamentous actin to rats resulted in lethal pulmonary hemorrhage and thrombosis. [50]

Another key “toxicity” of exposed actin is the fact that it is a major component of biofilms that accumulate at local sites of injury and infection, and that it impedes the access of host immune components and therapeutics such as antibiotics. [54] [90] Biofilms are particularly pathogenic in the setting of foreign bodies like indwelling catheters and tissue implants. [54] As a result of actin exposure at the local site of injury, the local level of plasma gelsolin around the site of the injury initially becomes depleted as it “debrides” the local involved site. [36] Mediators of inflammation, the body's innate healing mechanism, accumulate at the site of the injury to begin the processes of defense and repair, and the depletion of local plasma gelsolin allows them to do their work. [36] While local pGSN levels are depressed, the presence of this abundant protein in the circulation ensures that the inflammatory process stays local, and that stores of plasma gelsolin are available to address further injury so that the overall immune response remains intact.

Augmentation of macrophage antimicrobial activity

pGSN has antimicrobial activity in vitro and in vivo. Administration of pGSN subcutaneously or by inhalation to mice challenged with lethal inocula of S. pneumoniae or even more lethal combinations of influenza virus and bacteria markedly diminished the number of viable bacteria in the animals’ airways and significantly reduced mortality. The number of inflammation-inducing neutrophils was also considerably reduced, presumably as a result of enhanced bacterial clearance. This is true for contemporaneous or delayed administration of recombinant pGSN. [76] [91]

A basis of pGSN's antimicrobial action is that it enhances the ability of cultivated lung macrophages to ingest gram positive and gram negative bacteria. This has been demonstrated in vitro. [76] Improved phagocytosis is the product of pGSN debriding actin bound to macrophage scavenger receptors preventing their function. [73] pGSN also increases the ability of macrophages to kill ingested microorganisms by inducing macrophage nitric oxide synthase activity. [76]

Binding and inactivation of diverse inflammatory mediators

pGSN binds to a number of inflammatory mediators and signaling agents. Binding to LPA occurs at the same site on the molecule that ligates actin and interacts with polyphosphoinositides. [31] Subsequent studies showed that gelsolin alters the effector function of LPA's receptor binding. [32] [36] Binding to inflammatory mediators, and in some cases inhibition of their effector function, has been shown for platelet-activating factor, [36] lipopolysaccharide endotoxin, [34] sphingosine-1-phosphate, [35] and lipoteichoic acid [37] and small molecule purinergic agonists including ATP and ADP. [41] [43] [44] [42] The binding of pGSN to Alzheimer peptide has also been well documented. [24] [25] [92]

Role of mediators which bind to plasma gelsolin
Mediator Role
LPA [31] [32] [33] A phospholipid derivative that can act as a signaling molecule and activates G protein coupled receptors. It has been associated with cell proliferation.
LPS/endotoxin [33] [34] [35] Found in the outer membrane of Gram-negative bacteria, it elicits a strong immune response in animals.
PAF [36] A potent phospholipid activator and mediator of many leukocyte functions, including platelet aggregation, inflammation, and anaphylaxis. It is produced in response to specific stimuli by a variety of cell types, including neutrophils, basophils, platelets, and endothelial cells.
[24] [25] A peptide of 36–43 amino acids that is the main constituent of amyloid plaques in the brains of Alzheimer's disease patients.
LTA [35] A major constituent of the cell wall of Gram-positive bacteria able to stimulate a specific immune response in animals.
S1P [37] A blood borne lipid mediator and major regulator of vascular and immune systems. In the vascular system, S1P regulates angiogenesis, vascular stability, and permeability. In the immune system it is recognized as a major regulator of trafficking of T-cells and B-cells. Inhibition of S1P receptors has been shown to be critical for immunomodulation.

Anti-microbial resistance

Antimicrobial resistance is a global threat that leads to an estimated 700,000 deaths annually with projections of 10M deaths per year and lost economic potential of $100T by 2050. [93] [94] The United States has released a national action plan to combat antibiotic resistant bacteria. [95]

Recombinant pGSN (rhu-pGSN) supplementation alone shows improved survival and decreased bacteria counts in several mouse models. [91] [96] The bactericidal activity of the antimicrobial peptide LL-37 was shown to be inhibited by F-actin. It formed bundles with F-actin in vitro that were dissolved by pGSN, restoring bactericidal activity. Bacteria growth was reduced when pGSN was added cystic fibrosis sputum, which is known to contain F-actin. [97]

When mice were given a penicillin-resistant strain of pneumococcal pneumonia, penicillin had no effect on mortality or morbidity. rhu-pGSN improved both mortality and morbidity on its own, and the combination of rhu-pGSN and penicillin gave further improvement of both suggesting possible synergism. [96]

Levels of the Protein

Plasma gelsolin is produced and secreted by virtually every cell type with muscle contributing the largest amount. [98] At normal levels of >200 mg/L, it is a highly abundant protein in the circulation. [99]

Decreased levels are often associated with ill health and disease. [85] [100] A growing list of insults showing loss of pGSN includes pneumonia, [101] sepsis, [102] SIRS, [103] traumatic brain injury, [104] autoimmune diseases, [105] chronic kidney disease, [53] [106] HIV-1 disease, [107] tick-borne encephalitis and Lyme, [108] malaria, [109] [110] hepatitis, [111] burns, [112] [113] multiple organ dysfunction syndrome, [112] trauma associated with injury [114] or surgery, [106] bone marrow transplantation, [115] and multiple sclerosis. [116] Severely depleted levels (<150 mg/L) strongly correlate with the onset of systemic inflammatory dysregulation and predict increased morbidity and mortality across a broad spectrum of clinical presentations in the critical care setting. The magnitude of decline in pGSN correlates with the likelihood of mortality in seriously ill patients. [53] [106] [117]

Mediators of inflammation, the body's innate healing mechanism, accumulate at the site of the injury to begin the processes of defense and repair, [78] [79] [80] and the depletion of local plasma gelsolin allows them to do their work. [81] As a result of actin exposure at the local site of injury, the local level of plasma gelsolin around the site of the injury initially becomes depleted as it “debrides” the local involved site (see: Debridement). While local pGSN levels are depressed, the presence of this abundant protein in the circulation ensures that the inflammatory process stays local, [100] and that stores of plasma gelsolin are available to address further injury so that the overall immune response remains intact (see: Binding and inactivation of diverse inflammatory mediators).

Measured levels are higher in serum than plasma due to pGSN's affinity for fibrin. [99]

Animal studies

Human plasma gelsolin has been produced in recombinant form in E. coli (rhu-pGSN), and its efficacy as a therapeutic has been studied in vivo in a number of animal models of inflammatory disease. In models of injury that cause actin release and inflammatory organ damage, pGSN levels consistently drop. In models where gelsolin levels are replenished, adverse outcomes can be prevented. To date, rhu-pGSN has been studied in many independent laboratories providing evidence of efficacy in >20 animal models. Following are descriptions of selected animal studies. All stated results are relative to those of placebo treatments.

Summary of clinical results from selected animal studies
Disease Model Results
influenza mouse Mice given a highly lethal form of influenza show increased survival at 12 day end of study point as well as decreased morbidity and decreased expression of pro-inflammatory genes when rhu-pGSN is administered 3 to 6 days after infection. [118]
pneumococcal pneumonia mouse Mice were given pneumococcal challenge 7 days after being given influenza. Supplementation of endogenous pGSN with rhu-pGSN improved bacterial clearance 15-fold, reduced neutrophilic inflammation, improved recovery of initial weight loss, and showed a dose-dependent improvement on survival. No antibiotics were given, demonstrating pGSN's ability to stimulate the innate immune response. [76]
burn rat Rats receiving 40% body surface area burn showed 90% loss of endogenous pGSN within 12 hours and slowly recovered to almost 50% after 6 days. Intravenous administration of rhu-pGSN partially or totally prevented the burn-associated increase in pulmonary microvascular permeability in a dose-dependent manner. [119] See also [120]
sepsis mouse Mice were intraperitoneally injected with endotoxin (LPS) or subjected to cecal ligation and puncture (CLP) (a small amount of intestinal contents were extracted into the cavity and the wound was sutured). Endogenous pGSN levels dropped to 50% post-challenge. Survival substantially improved with rhu-pGSN treatment in both groups: LPS study, 90% vs 0%; CLP study: 30% vs 0%. [121]
sepsis rat Relative to a previous mouse study [121] a smaller dosage of rhu-pGSN decreased morbidity in a double CLP sepsis model relative to sham treatments. The dosage was effective in intraveneous and subcutaneous injections, but less so with intraperitoneal injection (qualitative but not statistically significant) despite the latter being the site of injury. This evidenced the need for systemic availability of pGSN for recovery. [122]
Acute respiratory distress syndrome mouse Mice were subjected to 95% O2 for 72 hr and treated with rhu-pGSN after 24 and 48 hr. Hyperoxia produced severe diffuse congestion and edema with hemorrhage visible in lung histopathology, 70% reduction in endogenous pGSN, and an influx of neutrophils. Treatment with rhu-pGSN led to a 23% decrease in the authors' histpathological score, 65% decrease in BAL fluid neutrophil count, and a 29% reduction in an overall acute lung injury score. [123]
stroke rat Researchers induced middle cerebral artery occlusion with a direct injection of Endothelin 1, a vasoconstrictor. Animals treated with pGSN at the site of injury exhibited 50% infarction area, >2x use of both forepaws during exploration, and a decrease in whisker-stimulated reaction time (9 s, pGSN treated; 19 s untreated; 1 s healthy rat). [124]
multiple sclerosis mouse Mice with experimental autoimmune encephalomyelitis show decreased levels of pGSN in blood and increased levels in the brain. All rhu-pGSN-treated mice survived whereas 60% of control died within 30 days. Rhu-pGSN mice scored significantly better on clinical scores, smaller brain lesions imaged by MRI, less extra-cellular actin, and decreased myeloperoxidase activity. [125]
Alzheimer's mouse Two models of Alzheimer's were tested. Treatment mice that were tail-injected with a plasmid encoding human pGSN showed reduction in 42 in brain tissue, decreased amyloid, and increased concentration of microglia. [126] See also [127]
radiation mouse Mice irradiated with 137Cs γ-rays show a 50-75% decrease in endogenous levels of pGSN. Bleeding is a common consequence of heavy radiation exposure. Administration of rhu-pGSN improved clotting indices in later, but not middle, phases of recovery. Rhu-pGSN improved GSH and MDA oxidative stress indices. [128]
pain and inflammation mouse Intraperitoneal injection of acetic acid causes a pain response quantified by writhing. [129] Both rhu-pGSN and diclofenac sodium (DS), a standard analgesic drug, caused ~55% reduction in writhing. Similarly, tails placed in hot water caused mice to retract them in an average time of 2.3 s. DS increased time to withdrawal from 5.1 to 7.6 s depending on time of drug administration; rhu-pGSN increased time from 2.9 to 5.5 s. Both DS and rhu-pGSN showed significant reductions in swelling associated with paw injection of an inflammatory agent, γ- carrageenan, as well as decreases in measured cytokines TNF-α and IL-6. [130]
diabetes mouse Endogenous levels of pGSN decrease by ~50% with type 2 diabetes(T2D) in both humans and mice. In an oral glucose tolerance test, rhu-pGSN brought blood sugar levels down to levels comparable to sitagliptin, a T2D drug. Daily dose of rhu-pGSN kept blood sugar levels close to normal for the 7 days of treatment. Daily dose of sitagliptin increased levels of endogenous pGSN. [131]

Human Studies

In 2019 BioAegis Therapeutics conducted a Phase Ib/IIa safety study administering recombinant human pGSN to sick patients with community acquired pneumonia; no safety issues were found. [132] A 2020 Phase IIb placebo-controlled efficacy study has been approved for acute severe pneumonia due to COVID-19. The primary outcome was the proportion of patients surviving on Day 14 without mechanical ventilation, vasopressors, or dialysis. Evaluation of efficacy of rhu-pGSN was confounded by high survival rates of both treatment and placebo cohorts resulting from improvements made to the standard of care for COVID pneumonia. [133]

See also

Cytoplasmic gelsolin

Actin

Vitamin D-binding protein

References

  1. ^ Shi, Yigong (2004-08-01). "Caspase activation, inhibition, and reactivation: A mechanistic view". Protein Science. 13 (8): 1979–1987. doi: 10.1110/ps.04789804. ISSN  0961-8368. PMC  2279816. PMID  15273300.
  2. ^ Klaavuniemi, Tuula; Yamashiro, Sawako; Ono, Shoichiro (2008-09-19). "Caenorhabditis elegans Gelsolin-like Protein 1 Is a Novel Actin Filament-severing Protein with Four Gelsolin-like Repeats". Journal of Biological Chemistry. 283 (38): 26071–26080. doi: 10.1074/jbc.M803618200. ISSN  0021-9258. PMC  2533794. PMID  18640981.
  3. ^ Archer, Stuart K.; Claudianos, Charles; Campbell, Hugh D. (2005). "Evolution of the gelsolin family of actin-binding proteins as novel transcriptional coactivators". BioEssays. 27 (4): 388–396. doi: 10.1002/bies.20200. ISSN  1521-1878. PMID  15770676. S2CID  40585071.
  4. ^ a b c d e f Lee, William M.; Galbraith, Robert M. (1992-05-14). "The Extracellular Actin-Scavenger System and Actin Toxicity". New England Journal of Medicine. 326 (20): 1335–1341. doi: 10.1056/NEJM199205143262006. ISSN  0028-4793. PMID  1314333.
  5. ^ a b Yin, Helen L.; Stossel, Thomas P. (18 October 1979). "Control of cytoplasmic actin gel-sol transformation by gelsolin, a calcium-dependent regulatory protein". Nature. 281 (5732): 583–586. Bibcode: 1979Natur.281..583Y. doi: 10.1038/281583a0. PMID  492320. S2CID  4250013. Retrieved 13 February 2020.
  6. ^ a b Chaponnier, C.; Borgia, R.; Rungger-Brändle, E.; Weil, R.; Gabbiani, G. (1979-08-15). "An actin-destabilizing factor is present in human plasma". Experientia. 35 (8): 1039–1041. doi: 10.1007/bf01949928. ISSN  0014-4754. PMID  477868. S2CID  21319139.
  7. ^ a b Norberg, Renee; Thorstensson, Rigmor; Utter, Goran; Fagraeus, Astrid (1979-10-15). "F-Actin-Depolymerizing Activity of Human Serum". European Journal of Biochemistry. 100 (2): 575–583. doi: 10.1111/j.1432-1033.1979.tb04204.x. ISSN  0014-2956. PMID  389627.
  8. ^ Harris, H.E.; Bamburg, J.R.; Weeds, A.G. (1980-11-17). "Actin filament disassembly in blood plasma". FEBS Letters. 121 (1): 175–177. doi: 10.1016/0014-5793(80)81291-9. ISSN  0014-5793. PMID  6893965. S2CID  30794630.
  9. ^ Harris, H.E.; Gooch, J. (1981-01-12). "An actin depolymerizing protein from pig plasma". FEBS Letters. 123 (1): 49–53. doi: 10.1016/0014-5793(81)80017-8. ISSN  0014-5793. PMID  6894126. S2CID  27405593.
  10. ^ Harris, D. A.; Schwartz, J. H. (1981-11-01). "Characterization of brevin, a serum protein that shortens actin filaments". Proceedings of the National Academy of Sciences. 78 (11): 6798–6802. Bibcode: 1981PNAS...78.6798H. doi: 10.1073/pnas.78.11.6798. ISSN  0027-8424. PMC  349138. PMID  6947253.
  11. ^ a b Kwiatkowski, D. J.; Stossel, T. P.; Orkin, S. H.; Mole, J. E.; Colten, H. R.; Yin, H. L. (1986-10-02). "Plasma and cytoplasmic gelsolins are encoded by a single gene and contain a duplicated actin-binding domain". Nature. 323 (6087): 455–458. Bibcode: 1986Natur.323..455K. doi: 10.1038/323455a0. ISSN  0028-0836. PMID  3020431. S2CID  4356162.
  12. ^ a b Ashish (31 August 2007). "Global Structure Changes Associated with Ca2+ Activation of Full-length Human Plasma Gelsolin" (PDF). J Biol Chem. 282 (35): 25884–25892. doi: 10.1074/jbc.M702446200. PMID  17604278. S2CID  25974945. Retrieved 12 February 2020.
  13. ^ a b Burtnick, Leslie D.; Koepf, Edward K.; Grimes, Jonathan; Jones, E. Yvonne; Stuart, David I.; McLaughlin, Paul J.; Robinson, Robert C. (22 August 1997). "The Crystal Structure of Plasma Gelsolin: Implications for Actin Severing, Capping, and Nucleation". Cell. 90 (4): 661–670. doi: 10.1016/s0092-8674(00)80527-9. PMID  9288746. S2CID  11112433. Retrieved 12 February 2020.
  14. ^ Nag, Shalini; Ma, Qing; Wang, Hui; Chumnarnsilpa, Sakesit; Lee, Wei Lin; Larsson, Mårten; Kannan, Balakrishnan; Hernandez-Valladares, Maria; Burtnick, Leslie D.; Robinson, Robert C. (7 July 2009). "Ca2+ binding by domain 2 plays a critical role in the activation and stabilization of gelsolin" (PDF). PNAS. 106 (33): 13713–13718. Bibcode: 2009PNAS..10613713N. doi: 10.1073/pnas.0812374106. PMC  2720848. PMID  19666512. Retrieved 12 February 2020.
  15. ^ a b c d Nag, Shalini; Larsson, Mårten; Robinson, Robert C.; Burtnick, Leslie D. (2013-06-10). "Gelsolin: The tail of a molecular gymnast: Gelsolin Superfamily Proteins". Cytoskeleton. 70 (7): 360–384. doi: 10.1002/cm.21117. ISSN  1949-3584. PMID  23749648.
  16. ^ Yin, H. L.; Kwiatkowski, D. J.; Mole, J. E.; Cole, F. S. (1984-04-25). "Structure and biosynthesis of cytoplasmic and secreted variants of gelsolin". Journal of Biological Chemistry. 259 (8): 5271–5276. doi: 10.1016/S0021-9258(17)42985-1. ISSN  0021-9258. PMID  6325429. Retrieved 2020-03-02.
  17. ^ Moon, Myeong Hee; Kang, Duk Jin (2013-11-19), Apparatus for protein separation using capillary isoelectric focusing—hollow fiber flow field flow fractionation and method thereof, retrieved 2020-03-02
  18. ^ Vouyiouklis, Demetrius A.; Brophy, Peter J. (2002-11-18). "A Novel Gelsolin Isoform Expressed by Oligodendrocytes in the Central Nervous System". Journal of Neurochemistry. 69 (3): 995–1005. doi: 10.1046/j.1471-4159.1997.69030995.x. ISSN  0022-3042. PMID  9282921. S2CID  44552710.
  19. ^ Zorgati, Habiba; Larsson, Mårten; Ren, Weitong; Sim, Adelene Y. L.; Gettemans, Jan; Grimes, Jonathan M.; Li, Wenfei; Robinson, Robert C. (2019-07-09). "The role of gelsolin domain 3 in familial amyloidosis (Finnish type)". Proceedings of the National Academy of Sciences. 116 (28): 13958–13963. Bibcode: 2019PNAS..11613958Z. doi: 10.1073/pnas.1902189116. ISSN  0027-8424. PMC  6628662. PMID  31243148.
  20. ^ Baig, Ruqia Mehmood; Mahjabeen, Ishrat; Sabir, Maimoona; Masood, Nosheen; Ali, Kashif; Malik, Faraz Arshad; Kayani, Mahmood Akhtar (2013). "Mutational Spectrum of Gelsolin and Its Down Regulation Is Associated with Breast Cancer". Disease Markers. 34 (2): 71–80. doi: 10.1155/2013/795410. ISSN  0278-0240. PMC  3809971. PMID  23324580.
  21. ^ Edgar, Alasdair John (1990-08-01). "Gel electrophoresis of native gelsolin and gelsolin-actin complexes". Journal of Muscle Research and Cell Motility. 11 (4): 323–330. doi: 10.1007/BF01766670. ISSN  0142-4319. PMID  2174905. S2CID  11355042.
  22. ^ Burtnick, Leslie D; Urosev, Dunja; Irobi, Edward; Narayan, Kartik; Robinson, Robert C (2004-07-21). "Structure of the N-terminal half of gelsolin bound to actin: roles in severing, apoptosis and FAF". The EMBO Journal. 23 (14): 2713–2722. doi: 10.1038/sj.emboj.7600280. ISSN  0261-4189. PMC  514944. PMID  15215896.
  23. ^ a b Bohgaki, Miyuki; Matsumoto, Masaki; Atsumi, Tatsuya; Kondo, Takeshi; Yasuda, Shinsuke; Horita, Tetsuya; Nakayama, Keiichi I.; Okumura, Fumihiko; Hatakeyama, Shigetsugu; Koike, Takao (2011-01-24). "Plasma gelsolin facilitates interaction between β2 glycoprotein I and α5β1 integrin". Journal of Cellular and Molecular Medicine. 15 (1): 141–151. doi: 10.1111/j.1582-4934.2009.00940.x. ISSN  1582-1838. PMC  3822501. PMID  19840195.
  24. ^ a b c d Ray, Indrani; Chauhan, Abha; Wegiel, Jerzy; Chauhan, Ved P.S. (2000-01-24). "Gelsolin inhibits the fibrillization of amyloid beta-protein, and also defibrillizes its preformed fibrils". Brain Research. 853 (2): 344–351. doi: 10.1016/S0006-8993(99)02315-X. ISSN  0006-8993. PMID  10640633. S2CID  41363612.
  25. ^ a b c d Chauhan, Ved P S; Ray, Indrani; Chauhan, Abha; Wisniewski, Henryk M (1999). "Binding of Gelsolin, a Secretory Protein, to Amyloid β-Protein". Biochemical and Biophysical Research Communications. 258 (2): 241–6. doi: 10.1006/bbrc.1999.0623. PMID  10329371.
  26. ^ Welander, Hedvig; Bontha, Sai Vineela; Näsström, Thomas; Karlsson, Mikael; Nikolajeff, Fredrik; Danzer, Karin; Kostka, Marcus; Kalimo, Hannu; Lannfelt, Lars; Ingelsson, Martin; Bergström, Joakim (2011-08-19). "Gelsolin co-occurs with Lewy bodies in vivo and accelerates α-synuclein aggregation in vitro". Biochemical and Biophysical Research Communications. 412 (1): 32–38. doi: 10.1016/j.bbrc.2011.07.027. ISSN  0006-291X. PMID  21798243.
  27. ^ Svanström, Andreas; Grantham, Julie (2015-09-12). "The molecular chaperone CCT modulates the activity of the actin filament severing and capping protein gelsolin in vitro". Cell Stress and Chaperones. 21 (1): 55–62. doi: 10.1007/s12192-015-0637-5. ISSN  1355-8145. PMC  4679748. PMID  26364302.
  28. ^ Lind, S. E.; Janmey, P. A. (1984-11-10). "Human plasma gelsolin binds to fibronectin". The Journal of Biological Chemistry. 259 (21): 13262–13266. doi: 10.1016/S0021-9258(18)90687-3. ISSN  0021-9258. PMID  6092370.
  29. ^ Kalwat, Michael A.; Wiseman, Dean A.; Luo, Wei; Wang, Zhanxiang; Thurmond, Debbie C. (2012-01-01). "Gelsolin Associates with the N Terminus of Syntaxin 4 to Regulate Insulin Granule Exocytosis". Molecular Endocrinology. 26 (1): 128–141. doi: 10.1210/me.2011-1112. ISSN  0888-8809. PMC  3248323. PMID  22108804.
  30. ^ Khaitlina, Sofia; Fitz, Helene; Hinssen, Horst (2013-07-11). "The interaction of gelsolin with tropomyosin modulates actin dynamics". The FEBS Journal. 280 (18): 4600–4611. doi: 10.1111/febs.12431. ISSN  1742-4658. PMID  23844991.
  31. ^ a b c d Meerschaert, K. (1998-10-15). "Gelsolin and functionally similar actin-binding proteins are regulated by lysophosphatidic acid". The EMBO Journal. 17 (20): 5923–5932. doi: 10.1093/emboj/17.20.5923. ISSN  1460-2075. PMC  1170920. PMID  9774337.
  32. ^ a b c d Goetzl, Edward J.; Lee, Hsinyu; Azuma, Toshifumi; Stossel, Thomas P.; Turck, Christoph W.; Karliner, Joel S. (2000-05-12). "Gelsolin Binding and Cellular Presentation of Lysophosphatidic Acid". Journal of Biological Chemistry. 275 (19): 14573–14578. doi: 10.1074/jbc.275.19.14573. ISSN  0021-9258. PMID  10799543.
  33. ^ a b c d e f Mintzer, Evan; Sargsyan, Hasmik; Bittman, Robert (2006-01-18). "Lysophosphatidic acid and lipopolysaccharide bind to the PIP2-binding domain of gelsolin". Biochimica et Biophysica Acta (BBA) - Biomembranes. 1758 (1): 85–89. doi: 10.1016/j.bbamem.2005.12.009. ISSN  0005-2736. PMID  16460666.
  34. ^ a b c d Bucki, Robert; Georges, Penelope C.; Espinassous, Quentin; Funaki, Makoto; Pastore, Jennifer J.; Chaby, Richard; Janmey, Paul A. (2005-07-19). "Inactivation of Endotoxin by Human Plasma Gelsolin ". Biochemistry. 44 (28): 9590–9597. doi: 10.1021/bi0503504. ISSN  0006-2960. PMID  16008344.
  35. ^ a b c d e f g Bucki, Robert; Byfield, Fitzroy J.; Kulakowska, Alina; McCormick, Margaret E.; Drozdowski, Wieslaw; Namiot, Zbigniew; Hartung, Thomas; Janmey, Paul A. (2008-10-01). "Extracellular Gelsolin Binds Lipoteichoic Acid and Modulates Cellular Response to Proinflammatory Bacterial Wall Components". The Journal of Immunology. 181 (7): 4936–4944. doi: 10.4049/jimmunol.181.7.4936. ISSN  0022-1767. PMID  18802097.
  36. ^ a b c d e f g Osborn, Teresia M.; Dahlgren, Claes; Hartwig, John H.; Stossel, Thomas P. (2007-04-01). "Modifications of cellular responses to lysophosphatidic acid and platelet-activating factor by plasma gelsolin". American Journal of Physiology. Cell Physiology. 292 (4): –1323–C1330. doi: 10.1152/ajpcell.00510.2006. ISSN  0363-6143. PMID  17135294.
  37. ^ a b c d Bucki, Robert; Kułakowska, Alina; Byfield, Fitzroy J.; Żendzian-Piotrowska, Małgorzata; Baranowski, Marcin; Marzec, Michał; Winer, Jessamine P.; Ciccarelli, Nicholas J.; Górski, Jan; Drozdowski, Wiesław; Bittman, Robert; Janmey, Paul A. (2010-12-01). "Plasma gelsolin modulates cellular response to sphingosine 1-phosphate". American Journal of Physiology. Cell Physiology. 299 (6): –1516–C1523. doi: 10.1152/ajpcell.00051.2010. ISSN  0363-6143. PMC  3006327. PMID  20810916.
  38. ^ Janmey, P. A.; Stossel, T. P. (1987-01-22). "Modulation of gelsolin function by phosphatidylinositol 4,5-bisphosphate". Nature. 325 (6102): 362–364. Bibcode: 1987Natur.325..362J. doi: 10.1038/325362a0. ISSN  0028-0836. PMID  3027569. S2CID  4324043.
  39. ^ a b Lin, Keng-Mean; Wenegieme, Elizabeth; Lu, Pei-Jung; Chen, Ching-Shih; Yin, Helen L. (1997-08-15). "Gelsolin Binding to Phosphatidylinositol 4,5-Bisphosphate Is Modulated by Calcium and pH". Journal of Biological Chemistry. 272 (33): 20443–20450. doi: 10.1074/jbc.272.33.20443. ISSN  0021-9258. PMID  9252353.
  40. ^ a b Sun, Hui-qiao; Lin, Keng-mean; Yin, Helen L. (1997-08-25). "Gelsolin Modulates Phospholipase C Activity In Vivo through Phospholipid Binding". The Journal of Cell Biology. 138 (4): 811–820. doi: 10.1083/jcb.138.4.811. ISSN  0021-9525. PMC  2138049. PMID  9265648.
  41. ^ a b Vartanian, Amalia A (March 2003). "Gelsolin and plasminogen activator inhibitor-1 are Ap3A-binding proteins". The Italian Journal of Biochemistry. 52 (1): 9–16. ISSN  0021-2938. PMID  12833632.
  42. ^ a b Yamamoto, Hideo; Ito, Hiroaki; Nakamura, Hideji; Hayashi, Eijiro; Kishimoto, Susumu; Hashimoto, Tadao; Tagawa, Kunio (1990-10-01). "Human Plasma Gelsolin Binds Adenosine Triphosphate". The Journal of Biochemistry. 108 (4): 505–506. doi: 10.1093/oxfordjournals.jbchem.a123229. ISSN  0021-924X. PMID  1963427. Retrieved 2020-02-28.
  43. ^ a b Urosev, Dunja; Ma, Qing; Tan, Agnes L.C.; Robinson, Robert C.; Burtnick, Leslie D. (2006-03-31). "The Structure of Gelsolin Bound to ATP". Journal of Molecular Biology. 357 (3): 765–772. doi: 10.1016/j.jmb.2006.01.027. ISSN  0022-2836. PMID  16469333.
  44. ^ a b Laham, Lorraine E.; Way, Michael; Yin, Helen L.; Janmey, Paul A. (1995-11-15). "Identification of Two Sites in Gelsolin with Different Sensitivities to Adenine Nucleotides". European Journal of Biochemistry. 234 (1): 1–7. doi: 10.1111/j.1432-1033.1995.001_c.x. ISSN  0014-2956. PMID  8529627.
  45. ^ Szatmári, Dávid; Xue, Bo; Kannan, Balakrishnan; Burtnick, Leslie D.; Bugyi, Beáta; Nyitrai, Miklós; Robinson, Robert C. (2018-08-07). "ATP competes with PIP2 for binding to gelsolin". PLOS ONE. 13 (8). Eugene A. Permyakov (ed.): –0201826. Bibcode: 2018PLoSO..1301826S. doi: 10.1371/journal.pone.0201826. ISSN  1932-6203. PMC  6080781. PMID  30086165.
  46. ^ a b c d Janmey, P. A.; Lind, S. E. (August 1987). "Capacity of human serum to depolymerize actin filaments". Blood. 70 (2): 524–530. doi: 10.1182/blood.V70.2.524.524. ISSN  0006-4971. PMID  3038216.
  47. ^ Janmey, P. A.; Iida, K.; Yin, H. L.; Stossel, T. P. (1987-09-05). "Polyphosphoinositide micelles and polyphosphoinositide-containing vesicles dissociate endogenous gelsolin-actin complexes and promote actin assembly from the fast-growing end of actin filaments blocked by gelsolin". The Journal of Biological Chemistry. 262 (25): 12228–12236. doi: 10.1016/S0021-9258(18)45341-0. ISSN  0021-9258. PMID  3040735.
  48. ^ Erukhimov, Jeffrey A.; Tang, Zi-Lue; Johnson, Bruce A.; Donahoe, Michael P.; Razzack, Jamal A.; Gibson, Kevin F.; Lee, William M.; Wasserloos, Karla J.; Watkins, Simon A.; Pitt, Bruce R. (July 2000). "Actin-Containing Sera From Patients With Adult Respiratory Distress Syndrome Are Toxic to Sheep Pulmonary Endothelial Cells". American Journal of Respiratory and Critical Care Medicine. 162 (1): 288–294. doi: 10.1164/ajrccm.162.1.9806088. ISSN  1073-449X. PMID  10903256. S2CID  23974368.
  49. ^ Martinez-Amat, A (2005-11-01). "Release of α-actin into serum after skeletal muscle damage". British Journal of Sports Medicine. 39 (11): 830–834. doi: 10.1136/bjsm.2004.017566. ISSN  0306-3674. PMC  1725075. PMID  16244192.
  50. ^ a b Haddad, J G; Harper, K D; Guoth, M; Pietra, G G; Sanger, J W (February 1990). "Angiopathic consequences of saturating the plasma scavenger system for actin". Proceedings of the National Academy of Sciences of the United States of America. 87 (4): 1381–1385. Bibcode: 1990PNAS...87.1381H. doi: 10.1073/pnas.87.4.1381. ISSN  0027-8424. PMC  53479. PMID  2154744.
  51. ^ Scarborough, Victoria D.; Bradford, Harvey R.; Ganguly, Pankaj (1981-06-16). "Aggregation of platelets by muscle actin. A multivalent interaction model of platelet aggregation by ADP". Biochemical and Biophysical Research Communications. 100 (3): 1314–1319. doi: 10.1016/0006-291X(81)91967-7. ISSN  0006-291X. PMID  6895029.
  52. ^ a b Vasconcellos, Ca; Lind, Se (1993-12-15). "Coordinated inhibition of actin-induced platelet aggregation by plasma gelsolin and vitamin D-binding protein". Blood. 82 (12): 3648–3657. doi: 10.1182/blood.V82.12.3648.bloodjournal82123648 (inactive 2024-04-27). ISSN  0006-4971. PMID  8260702. Retrieved 2020-02-13.{{ cite journal}}: CS1 maint: DOI inactive as of April 2024 ( link)
  53. ^ a b c Lee, Po-Shun; Sampath, Kartik; Karumanchi, S. Ananth; Tamez, Hector; Bhan, Ishir; Isakova, Tamara; Gutierrez, Orlando M.; Wolf, Myles; Chang, Yuchiao; Stossel, Thomas P.; Thadhani, Ravi (2009-04-23). "Plasma Gelsolin and Circulating Actin Correlate with Hemodialysis Mortality". Journal of the American Society of Nephrology. 20 (5): 1140–1148. doi: 10.1681/ASN.2008091008. ISSN  1046-6673. PMC  2678046. PMID  19389844.
  54. ^ a b c Walker, T. S.; Tomlin, K. L.; Worthen, G. S.; Poch, K. R.; Lieber, J. G.; Saavedra, M. T.; Fessler, M. B.; Malcolm, K. C.; Vasil, M. L.; Nick, J. A. (2005-06-01). "Enhanced Pseudomonas aeruginosa Biofilm Development Mediated by Human Neutrophils". Infection and Immunity. 73 (6): 3693–3701. doi: 10.1128/IAI.73.6.3693-3701.2005. ISSN  0019-9567. PMC  1111839. PMID  15908399.
  55. ^ Kudryashov, Dmitri S.; Reisler, Emil (April 2013). "ATP and ADP Actin States". Biopolymers. 99 (4): 245–256. doi: 10.1002/bip.22155. ISSN  0006-3525. PMC  3670783. PMID  23348672.
  56. ^ Lind, S E; Smith, D B; Janmey, P A; Stossel, T P (1986-09-01). "Role of plasma gelsolin and the vitamin D-binding protein in clearing actin from the circulation". Journal of Clinical Investigation. 78 (3): 736–742. doi: 10.1172/JCI112634. ISSN  0021-9738. PMC  423663. PMID  3018044.
  57. ^ Janmey, Paul A.; Stossel, Thomas P.; Lind, Stuart E. (1986-04-14). "Sequential binding of actin monomers to plasma gelsolin and its inhibition by vitamin D-binding protein". Biochemical and Biophysical Research Communications. 136 (1): 72–79. doi: 10.1016/0006-291X(86)90878-8. ISSN  0006-291X. PMID  3010978.
  58. ^ Herrmannsdoerfer, A. J.; Heeb, G. T.; Feustel, P. J.; Estes, J. E.; Keenan, C. J.; Minnear, F. L.; Selden, L.; Giunta, C.; Flor, J. R.; Blumenstock, F. A. (December 1993). "Vascular clearance and organ uptake of G- and F-actin in the rat". The American Journal of Physiology. 265 (6 Pt 1): –1071–1081. doi: 10.1152/ajpgi.1993.265.6.G1071. ISSN  0002-9513. PMID  8279558.
  59. ^ a b Janmey, P. A.; Lamb, J. A.; Ezzell, R. M.; Hvidt, S.; Lind, S. E. (1992-08-15). "Effects of actin filaments on fibrin clot structure and lysis". Blood. 80 (4): 928–936. doi: 10.1182/blood.V80.4.928.928. ISSN  0006-4971. PMID  1323346.
  60. ^ Orlova, A.; Prochniewicz, E.; Egelman, E. H. (1995-02-03). "Structural dynamics of F-actin: II. Cooperativity in structural transitions". Journal of Molecular Biology. 245 (5): 598–607. doi: 10.1006/jmbi.1994.0049. ISSN  0022-2836. PMID  7844829.
  61. ^ Prochniewicz, Ewa; Zhang, Qingnan; Janmey, Paul A.; Thomas, David D. (August 1996). "Cooperativity in F-Actin: Binding of Gelsolin at the Barbed End Affects Structure and Dynamics of the Whole Filament". Journal of Molecular Biology. 260 (5): 756–766. doi: 10.1006/jmbi.1996.0435. ISSN  0022-2836. PMID  8709153.
  62. ^ Janmey, Paul A.; Chaponnier, Christine; Lind, Stuart E.; Zaner, Ken S.; Stossel, Thomas P.; Yin, Helen L. (July 1985). "Interactions of gelsolin and gelsolin-actin complexes with actin. Effects of calcium on actin nucleation, filament severing, and end blocking". Biochemistry. 24 (14): 3714–3723. doi: 10.1021/bi00335a046. ISSN  0006-2960. PMID  2994715.
  63. ^ Doi, Y.; Frieden, C. (1984-10-10). "Actin polymerization. The effect of brevin on filament size and rate of polymerization". Journal of Biological Chemistry. 259 (19): 11868–11875. doi: 10.1016/S0021-9258(20)71292-5. ISSN  0021-9258. PMID  6480587. Retrieved 2020-02-19.
  64. ^ Brooks, F.J.; Carlsson, A.E. (August 2008). "Actin Polymerization Overshoots and ATP Hydrolysis as Assayed by Pyrene Fluorescence". Biophysical Journal. 95 (3): 1050–1062. Bibcode: 2008BpJ....95.1050B. doi: 10.1529/biophysj.107.123125. ISSN  0006-3495. PMC  2479571. PMID  18390612.
  65. ^ Sept, David; McCammon, J. Andrew (2001-08-01). "Thermodynamics and Kinetics of Actin Filament Nucleation". Biophysical Journal. 81 (2): 667–674. Bibcode: 2001BpJ....81..667S. doi: 10.1016/S0006-3495(01)75731-1. ISSN  0006-3495. PMC  1301543. PMID  11463615.
  66. ^ Qu, Zheng; Silvan, Unai; Jockusch, Brigitte M.; Aebi, Ueli; Schoenenberger, Cora-Ann; Mannherz, Hans Georg (October 2015). "Distinct actin oligomers modulate differently the activity of actin nucleators". FEBS Journal. 282 (19): 3824–3840. doi: 10.1111/febs.13381. ISSN  1742-464X. PMID  26194975.
  67. ^ Edgar, Alasdair John (August 1990). "Gel electrophoresis of native gelsolin and gelsolin-actin complexes". Journal of Muscle Research and Cell Motility. 11 (4): 323–330. doi: 10.1007/BF01766670. ISSN  0142-4319. PMID  2174905. S2CID  11355042.
  68. ^ Lal, A. A.; Korn, E. D.; Brenner, S. L. (1984-07-25). "Rate constants for actin polymerization in ATP determined using cross-linked actin trimers as nuclei". The Journal of Biological Chemistry. 259 (14): 8794–8800. doi: 10.1016/S0021-9258(17)47223-1. ISSN  0021-9258. PMID  6746624.
  69. ^ Janmey, Paul A.; Stossel, Thomas P. (1986-10-01). "Kinetics of actin monomer exchange at the slow growing ends of actin filaments and their relation to the elongation of filaments shortened by gelsolin". Journal of Muscle Research & Cell Motility. 7 (5): 446–454. doi: 10.1007/BF01753587. ISSN  1573-2657. PMID  3025252. S2CID  2644111.
  70. ^ Gregory, Julia L; Prada, Claudia M; Fine, Sara J; Garcia-Alloza, Monica; Betensky, Rebecca A; Arbel-Ornath, Michal; Greenberg, Steven M; Bacskai, Brian J; Frosch, Matthew P (2012). "Reducing Available Soluble A-Amyloid Prevents Progression of Cerebral Amyloid Angiopathy in Transgenic Mice". J Neuropathol Exp Neurol. 71 (11): 1009–17. doi: 10.1097/NEN.0b013e3182729845. PMC  3491571. PMID  23095848.
  71. ^ Leverenz, James B.; Umar, Imran; Wang, Qing; Montine, Thomas J.; McMillan, Pamela J.; Tsuang, Debby W.; Jin, Jinghua; Pan, Catherine; Shin, Jenny; Zhu, David; Zhang, Jing (2007-04-01). "Proteomic identification of novel proteins in cortical lewy bodies". Brain Pathology (Zurich, Switzerland). 17 (2): 139–145. doi: 10.1111/j.1750-3639.2007.00048.x. ISSN  1015-6305. PMC  8095629. PMID  17388944. S2CID  24457175.
  72. ^ Welander, Hedvig; Bontha, Sai Vineela; Näsström, Thomas; Karlsson, Mikael; Nikolajeff, Fredrik; Danzer, Karin; Kostka, Marcus; Kalimo, Hannu; Lannfelt, Lars; Ingelsson, Martin; Bergström, Joakim (2011-07-21). "Gelsolin co-occurs with Lewy bodies in vivo and accelerates α-synuclein aggregation in vitro". Biochemical and Biophysical Research Communications. 412 (1): 32–38. doi: 10.1016/j.bbrc.2011.07.027. ISSN  0006-291X. PMID  21798243.
  73. ^ a b c Ordija, Christine M.; Chiou, Terry Ting-Yu; Yang, Zhiping; Deloid, Glen M.; de Oliveira Valdo, Melina; Wang, Zhi; Bedugnis, Alice; Noah, Terry L.; Jones, Samuel; Koziel, Henry; Kobzik, Lester (2017-06-01). "Free actin impairs macrophage bacterial defenses via scavenger receptor MARCO interaction with reversal by plasma gelsolin". American Journal of Physiology. Lung Cellular and Molecular Physiology. 312 (6): –1018–L1028. doi: 10.1152/ajplung.00067.2017. ISSN  1040-0605. PMC  5495953. PMID  28385809.
  74. ^ Bougaki, Masahiko; Searles, Robert J.; Kida, Kotaro; De Yu, Jia; Buys, Emmanuel S.; Ichinose, Fumito (2010-09-01). "NOS3 protects against systemic inflammation and myocardial dysfunction in murine polymicrobial sepsis". Shock (Augusta, Ga.). 34 (3): 281–290. doi: 10.1097/SHK.0b013e3181cdc327. ISSN  1073-2322. PMC  3774000. PMID  19997049.
  75. ^ Jimenez-Sousa, Ma Angeles; López, Elisabeth; Fernandez-Rodríguez, Amanda; Tamayo, Eduardo; Fernández-Navarro, Pablo; Segura-Roda, Laura; Heredia, María; Gómez-Herreras, José I.; Bustamante, Jesús; García-Gómez, Juan Miguel; Bermejo-Martin, Jesús F.; Resino, Salvador (2012-07-20). "Genetic polymorphisms located in genes related to immune and inflammatory processes are associated with end-stage renal disease: a preliminary study". BMC Medical Genetics. 13 (1): 58. doi: 10.1186/1471-2350-13-58. ISSN  1471-2350. PMC  3412707. PMID  22817530.
  76. ^ a b c d e f g Yang, Zhiping; Chiou, Terry Ting-Yu; Stossel, Thomas P.; Kobzik, Lester (2015-07-01). "Plasma gelsolin improves lung host defense against pneumonia by enhancing macrophage NOS3 function". American Journal of Physiology. Lung Cellular and Molecular Physiology. 309 (1): –11–L16. doi: 10.1152/ajplung.00094.2015. ISSN  1040-0605. PMC  4491512. PMID  25957291.
  77. ^ Chambliss, Ken L.; Shaul, Philip W. (October 2002). "Estrogen Modulation of Endothelial Nitric Oxide Synthase". Endocrine Reviews. 23 (5): 665–686. doi: 10.1210/er.2001-0045. ISSN  0163-769X. PMID  12372846.
  78. ^ a b Zhao, Yutong; Natarajan, Viswanathan (January 2013). "Lysophosphatidic acid (LPA) and its Receptors: Role in Airway Inflammation and Remodeling". Biochimica et Biophysica Acta (BBA) - Molecular and Cell Biology of Lipids. 1831 (1): 86–92. doi: 10.1016/j.bbalip.2012.06.014. ISSN  0006-3002. PMC  3491109. PMID  22809994.
  79. ^ a b Shaw, J. O.; Pinckard, R. N.; Ferrigni, K. S.; McManus, L. M.; Hanahan, D. J. (1981-09-01). "Activation of human neutrophils with 1-O-hexadecyl/octadecyl-2-acetyl-sn-glycerol-3-phosphorylcholine (platelet activating factor)". The Journal of Immunology. 127 (3): 1250–1255. doi: 10.4049/jimmunol.127.3.1250. ISSN  1550-6606. PMID  6267133. S2CID  23647018. Retrieved 2020-03-05.
  80. ^ a b Obinata, Hideru; Hla, Timothy (2012-01-01). "Sphingosine 1-phosphate in coagulation and inflammation". Seminars in Immunopathology. 34 (1): 73–91. doi: 10.1007/s00281-011-0287-3. ISSN  1863-2300. PMC  3237867. PMID  21805322.
  81. ^ a b Piktel, Ewelina; Levental, Ilya; Durnaś, Bonita; Janmey, Paul A.; Bucki, Robert (2018-08-25). "Plasma Gelsolin: Indicator of Inflammation and Its Potential as a Diagnostic Tool and Therapeutic Target". International Journal of Molecular Sciences. 19 (9): 2516. doi: 10.3390/ijms19092516. ISSN  1422-0067. PMC  6164782. PMID  30149613.
  82. ^ Bancos, Simona; Bernard, Matthew P.; Topham, David J.; Phipps, Richard P. (2009). "Ibuprofen and other widely used non-steroidal anti-inflammatory drugs inhibit antibody production in human cells". Cellular Immunology. 258 (1): 18–28. doi: 10.1016/j.cellimm.2009.03.007. ISSN  0008-8749. PMC  2693360. PMID  19345936.
  83. ^ Coutinho, Agnes E.; Chapman, Karen E. (2011-03-15). "The anti-inflammatory and immunosuppressive effects of glucocorticoids, recent developments and mechanistic insights". Molecular and Cellular Endocrinology. 335 (1): 2–13. doi: 10.1016/j.mce.2010.04.005. ISSN  0303-7207. PMC  3047790. PMID  20398732.
  84. ^ Gm, Anstead (1998-10-01). "Steroids, retinoids, and wound healing". Advances in Wound Care: The Journal for Prevention and Healing. 11 (6): 277–285. ISSN  1076-2191. PMID  10326344.
  85. ^ a b Bucki, Robert; Levental, Ilya; Kulakowska, Alina; Janmey, Paul A. (2008-12-01). "Plasma gelsolin: function, prognostic value, and potential therapeutic use". Current Protein & Peptide Science. 9 (6): 541–551. doi: 10.2174/138920308786733912. ISSN  1389-2037. PMID  19075745.
  86. ^ Janmey, P. A.; Lind, S. E. (1987-08-01). "Capacity of human serum to depolymerize actin filaments". Blood. 70 (2): 524–530. doi: 10.1182/blood.V70.2.524.524. ISSN  0006-4971. PMID  3038216.
  87. ^ Stuart E LindS; Smith, Carolyn J (1991-03-15). "Actin Is a Noncompetitive Plasmin Inhibitor" (PDF). Journal of Biological Chemistry. 266 (8): 5273–5278. doi: 10.1016/S0021-9258(19)67783-5. PMID  1848244.
  88. ^ Erukhimov, Jeffrey A.; Tang, Zi-Lue; Johnson, Bruce A.; Donahoe, Michael P.; Razzack, Jamal A.; Gibson, Kevin F.; Lee, William M.; Wasserloos, Karla J.; Watkins, Simon A.; Pitt, Bruce R. (2000-07-01). "Actin-Containing Sera From Patients With Adult Respiratory Distress Syndrome Are Toxic to Sheep Pulmonary Endothelial Cells". American Journal of Respiratory and Critical Care Medicine. 162 (1): 288–294. doi: 10.1164/ajrccm.162.1.9806088. ISSN  1073-449X. PMID  10903256. S2CID  23974368.
  89. ^ Sousa, Caetano Reis e (2017-03-01). "Sensing infection and tissue damage". EMBO Molecular Medicine. 9 (3): 285–288. doi: 10.15252/emmm.201607227. ISSN  1757-4684. PMC  5331196. PMID  28119319.
  90. ^ Parks, Quinn M.; Young, Robert L.; Poch, Katie R.; Malcolm, Kenneth C.; Vasil, Michael L.; Nick, Jerry A. (2009-04-01). "Neutrophil enhancement of Pseudomonas aeruginosa biofilm development: human F-actin and DNA as targets for therapy". Journal of Medical Microbiology. 58 (Pt 4): 492–502. doi: 10.1099/jmm.0.005728-0. ISSN  0022-2615. PMC  2677169. PMID  19273646.
  91. ^ a b Yang, Zhiping; Levinson, Susan; Stossel, Thomas; DiNubile, Mark; Kobzik, Lester (2017-10-04). "Delayed Therapy with Plasma Gelsolin Improves Survival in Murine Pneumococcal Pneumonia". Open Forum Infectious Diseases. 4 (Suppl 1): –474–S475. doi: 10.1093/ofid/ofx163.1215. ISSN  2328-8957. PMC  5630930.
  92. ^ Ji, Lina; Zhao, Xi; Hua, Zichun (2015-01-06). "Potential Therapeutic Implications of Gelsolin in Alzheimer's Disease". Journal of Alzheimer's Disease. 44 (1): 13–25. doi: 10.3233/JAD-141548. ISSN  1875-8908. PMID  25208622.
  93. ^ Hoffman SJ, Outterson K, Røttingen JA, Cars O, Clift C, Rizvi Z, Rotberg F, Tomson G, Zorzet A (February 2015). "An international legal framework to address antimicrobial resistance". Bulletin of the World Health Organization. 93 (2): 66. doi: 10.2471/BLT.15.152710. PMC  4339972. PMID  25883395.
  94. ^ O'Neill, Jim. "Tackling drug-resistant infections globally: final report and recommendations" (PDF). amr-review.org. Review on Antimicrobial Resistance. Retrieved 5 March 2020.
  95. ^ "FACT SHEET: Obama Administration Releases National Action Plan to Combat Antibiotic-Resistant Bacteria". obamawhitehouse.archives.gov. 27 March 2015. Retrieved 5 March 2020.
  96. ^ a b Yang, Zhiping; Bedugnis, Alice; Levinson, Susan; Dinubile, Mark; Stossel, Thomas; Lu, Quan; Kobzik, Lester (2019-09-26). "Delayed Administration of Recombinant Plasma Gelsolin Improves Survival in a Murine Model of Penicillin-Susceptible and Penicillin-Resistant Pneumococcal Pneumonia". The Journal of Infectious Diseases. 220 (9): 1498–1502. doi: 10.1093/infdis/jiz353. ISSN  0022-1899. PMC  6761947. PMID  31287867.
  97. ^ Weiner, Daniel J.; Bucki, Robert; Janmey, Paul A. (June 2003). "The Antimicrobial Activity of the Cathelicidin LL37 Is Inhibited by F-actin Bundles and Restored by Gelsolin". American Journal of Respiratory Cell and Molecular Biology. 28 (6): 738–745. doi: 10.1165/rcmb.2002-0191OC. ISSN  1044-1549. PMID  12600826.
  98. ^ Kwiatkowski, D. J.; Mehl, R.; Izumo, S.; Nadal-Ginard, B.; Yin, H. L. (1988-06-15). "Muscle is the major source of plasma gelsolin". The Journal of Biological Chemistry. 263 (17): 8239–8243. doi: 10.1016/S0021-9258(18)68469-8. ISSN  0021-9258. PMID  2836420.
  99. ^ a b Smith, D. B.; Janmey, P. A.; Herbert, T. J.; Lind, S. E. (August 1987). "Quantitative measurement of plasma gelsolin and its incorporation into fibrin clots". The Journal of Laboratory and Clinical Medicine. 110 (2): 189–195. ISSN  0022-2143. PMID  3036979.
  100. ^ a b Peddada, Nagesh; Sagar, Amin; Ashish; Garg, Renu (February 2012). "Plasma gelsolin: A general prognostic marker of health". Medical Hypotheses. 78 (2): 203–210. doi: 10.1016/j.mehy.2011.10.024. ISSN  0306-9877. PMID  22082609.
  101. ^ Self, Wesley H; Wunderink, Richard G; DiNubile, Mark J; Stossel, Thomas P; Levinson, Susan L; Williams, Derek J; Anderson, Evan J; Bramley, Anna M; Jain, Seema; Edwards, Kathryn M; Grijalva, Carlos G (2019-09-13). "Low Admission Plasma Gelsolin Concentrations Identify Community-acquired Pneumonia Patients at High Risk for Severe Outcomes". Clinical Infectious Diseases. 69 (7): 1218–1225. doi: 10.1093/cid/ciy1049. ISSN  1058-4838. PMC  6743831. PMID  30561561.
  102. ^ Wang, HaiHong; Cheng, BaoLi; Chen, QiXing; Wu, ShuiJing; Lv, Chen; Xie, GuoHao; Jin, Yue; Fang, XiangMing (2008). "Time course of plasma gelsolin concentrations during severe sepsis in critically ill surgical patients". Critical Care. 12 (4): –106. doi: 10.1186/cc6988. ISSN  1364-8535. PMC  2575595. PMID  18706105.
  103. ^ Horváth-Szalai, Zoltán; Kustán, Péter; Mühl, Diána; Ludány, Andrea; Bugyi, Beáta; Kőszegi, Tamás (February 2017). "Antagonistic sepsis markers: Serum gelsolin and actin/gelsolin ratio". Clinical Biochemistry. 50 (3): 127–133. doi: 10.1016/j.clinbiochem.2016.10.018. ISSN  0009-9120. PMID  27823961.
  104. ^ Jin, Yong; Li, Bo-You; Qiu, Ling-Li; Ling, Yuan-Ren; Bai, Zhi-Qiang (2012-10-01). "Decreased plasma gelsolin is associated with 1-year outcome in patients with traumatic brain injury". Journal of Critical Care. 27 (5): 527–1–527.e6. doi: 10.1016/j.jcrc.2012.01.002. ISSN  0883-9441. PMID  22386223. Retrieved 2020-03-04.
  105. ^ Hu, Yl; Li, H.; Li, W. H.; Meng, H. X.; Fan, Y. Z.; Li, W. J.; Ji, Y. T.; Zhao, H.; Zhang, L.; Jin, X. M.; Zhang, F. M. (December 2013). "The value of decreased plasma gelsolin levels in patients with systemic lupus erythematosus and rheumatoid arthritis in diagnosis and disease activity evaluation". Lupus. 22 (14): 1455–1461. doi: 10.1177/0961203313507985. ISSN  1477-0962. PMID  24122723. S2CID  22637925.
  106. ^ a b c Lee, Po-Shun; Drager, Leslie R.; Stossel, Thomas P.; Moore, Francis D.; Rogers, Selwyn O. (March 2006). "Relationship of plasma gelsolin levels to outcomes in critically ill surgical patients". Annals of Surgery. 243 (3): 399–403. doi: 10.1097/01.sla.0000201798.77133.55. ISSN  0003-4932. PMC  1448930. PMID  16495706.
  107. ^ Sinha, Kislay Kumar; Peddada, Nagesh; Jha, Pravin Kumar; Mishra, Anshul; Pandey, Krishna; Das, Vidya Nand Ravi; Ashish; Das, Pradeep (March 2017). "Plasma Gelsolin Level in HIV-1-Infected Patients: An Indicator of Disease Severity". AIDS Research and Human Retroviruses. 33 (3): 254–260. doi: 10.1089/aid.2016.0154. ISSN  0889-2229. PMID  27700141.
  108. ^ Kułakowska, Alina; Zajkowska, Joanna M.; Ciccarelli, Nicholas J.; Mroczko, Barbara; Drozdowski, Wiesław; Bucki, Robert (2011). "Depletion of Plasma Gelsolin in Patients with Tick-Borne Encephalitis and Lyme Neuroborreliosis". Neurodegenerative Diseases. 8 (5): 375–380. doi: 10.1159/000324373. ISSN  1660-2862. PMC  3121545. PMID  21389683.
  109. ^ Huang, S.; Rhoads, S. L.; DiNubile, M. J. (May 1997). "Temporal association between serum gelsolin levels and clinical events in a patient with severe falciparum malaria". Clinical Infectious Diseases. 24 (5): 951–954. doi: 10.1093/clinids/24.5.951. ISSN  1058-4838. PMID  9142799.
  110. ^ Kassa, Fikregabrail Aberra; Shio, Marina Tiemi; Bellemare, Marie-Josée; Faye, Babacar; Ndao, Momar; Olivier, Martin (2011-10-20). "New Inflammation-Related Biomarkers during Malaria Infection". PLOS ONE. 6 (10): e26495. Bibcode: 2011PLoSO...626495K. doi: 10.1371/journal.pone.0026495. ISSN  1932-6203. PMC  3197653. PMID  22028888.
  111. ^ Ito, H.; Kambe, H.; Kimura, Y.; Nakamura, H.; Hayashi, E.; Kishimoto, T.; Kishimoto, S.; Yamamoto, H. (May 1992). "Depression of plasma gelsolin level during acute liver injury". Gastroenterology. 102 (5): 1686–1692. doi: 10.1016/0016-5085(92)91731-i. ISSN  0016-5085. PMID  1314752.
  112. ^ a b Huang, Li-feng; Yao, Yong-ming; Li, Jin-feng; Dong, Ning; Liu, Chen; Yu, Yan; He, Li-xin; Sheng, Zhi-yong (2011-11-01). "Reduction of Plasma Gelsolin Levels Correlates with Development of Multiple Organ Dysfunction Syndrome and Fatal Outcome in Burn Patients". PLOS ONE. 6 (11): –25748. Bibcode: 2011PLoSO...625748H. doi: 10.1371/journal.pone.0025748. ISSN  1932-6203. PMC  3206022. PMID  22069445.
  113. ^ Xianhui, Li; Pinglian, Li; Xiaojuan, Wang; Wei, Chen; Yong, Yang; Feng, Ran; Peng, Sun; Gang, Xue (December 2014). "The association between plasma gelsolin level and prognosis of burn patients". Burns: Journal of the International Society for Burn Injuries. 40 (8): 1552–1555. doi: 10.1016/j.burns.2014.02.020. ISSN  1879-1409. PMID  24690274.
  114. ^ Mounzer, Karam C.; Moncure, Michael; Smith, Yolanda R.; DiNUBILE, Mark J. (November 1999). "Relationship of Admission Plasma Gelsolin Levels to Clinical Outcomes in Patients after Major Trauma". American Journal of Respiratory and Critical Care Medicine. 160 (5): 1673–1681. doi: 10.1164/ajrccm.160.5.9807137. ISSN  1535-4970. PMID  10556139.
  115. ^ DiNubile, Mark J.; Stossel, Thomas P.; Ljunghusen, Olof C.; Ferrara, James L. M.; Antin, Joseph H. (2002-12-15). "Prognostic implications of declining plasma gelsolin levels after allogeneic stem cell transplantation". Blood. 100 (13): 4367–4371. doi: 10.1182/blood-2002-06-1672. ISSN  1528-0020. PMID  12393536. Retrieved 2020-02-12.
  116. ^ Kułakowska, Alina; Ciccarelli, Nicholas J; Wen, Qi; Mroczko, Barbara; Drozdowski, Wiesław; Szmitkowski, Maciej; Janmey, Paul A; Bucki, Robert (December 2010). "Hypogelsolinemia, a disorder of the extracellular actin scavenger system, in patients with multiple sclerosis". BMC Neurology. 10 (1): 107. doi: 10.1186/1471-2377-10-107. ISSN  1471-2377. PMC  2989318. PMID  21040581.
  117. ^ Lee, Po-Shun; Patel, Sanjay R.; Christiani, David C.; Bajwa, Ednan; Stossel, Thomas P.; Waxman, Aaron B. (2008-11-12). "Plasma Gelsolin Depletion and Circulating Actin in Sepsis—A Pilot Study". PLOS ONE. 3 (11): –3712. Bibcode: 2008PLoSO...3.3712L. doi: 10.1371/journal.pone.0003712. ISSN  1932-6203. PMC  2577888. PMID  19002257.
  118. ^ Yang, Zhiping; Bedugnis, Alice; Levinson, Susan; DiNubile, Mark; Stossel, Thomas; Lu, Quan; Kobzik, Lester (2020-02-21). "Delayed administration of recombinant plasma gelsolin improves survival in a murine model of severe influenza". F1000Research. 8: 1860. doi: 10.12688/f1000research.21082.2. ISSN  2046-1402. PMC  6894358. PMID  31824672.
  119. ^ Rothenbach, Patricia A.; Dahl, Benny; Schwartz, Jason J.; O'Keefe, Grant E.; Yamamoto, Masaya; Lee, William M.; Horton, Jureta W.; Yin, Helen L.; Turnage, Richard H. (January 2004). "Recombinant plasma gelsolin infusion attenuates burn-induced pulmonary microvascular dysfunction". Journal of Applied Physiology. 96 (1): 25–31. doi: 10.1152/japplphysiol.01074.2002. ISSN  8750-7587. PMID  12730154.
  120. ^ Zhang, Qing-Hong; Chen, Qi; Kang, Jia-Rui; Liu, Chen; Dong, Ning; Zhu, Xiao-Mei; Sheng, Zhi-Yong; Yao, Yong-Ming (2011-09-21). "Treatment with gelsolin reduces brain inflammation and apoptotic signaling in mice following thermal injury". Journal of Neuroinflammation. 8 (1): 118. doi: 10.1186/1742-2094-8-118. ISSN  1742-2094. PMC  3191361. PMID  21936896.
  121. ^ a b Lee, Po-Shun; Waxman, Aaron B.; Cotich, Kara L.; Chung, Su Wol; Perrella, Mark A.; Stossel, Thomas P. (March 2007). "Plasma gelsolin is a marker and therapeutic agent in animal sepsis*". Critical Care Medicine. 35 (3): 849–855. doi: 10.1097/01.CCM.0000253815.26311.24. ISSN  0090-3493. PMID  17205019. S2CID  21641666. Retrieved 2020-02-12.
  122. ^ Cohen, Taylor S.; Bucki, Robert; Byfield, Fitzroy J.; Ciccarelli, Nicholas J.; Rosenberg, Brenna; DiNubile, Mark J.; Janmey, Paul A.; Margulies, Susan S. (June 2011). "Therapeutic potential of plasma gelsolin administration in a rat model of sepsis". Cytokine. 54 (3): 235–238. doi: 10.1016/j.cyto.2011.02.006. ISSN  1043-4666. PMC  3083472. PMID  21420877.
  123. ^ Christofidou-Solomidou, Melpo; Scherpereel, Arnaud; Solomides, Charalambos C.; Christie, Jason D.; Stossel, Thomas P.; Goelz, Susan; DiNubile, Mark J. (2002-01-01). "Recombinant Plasma Gelsolin Diminishes the Acute Inflammatory Response to Hyperoxia in Mice". Journal of Investigative Medicine. 50 (1): 54–60. doi: 10.2310/6650.2002.33518. ISSN  1081-5589. PMID  11813829. S2CID  1981768. Retrieved 2020-02-24.
  124. ^ Le, Huong T; Hirko, Aaron C; Thinschmidt, Jeffrey S; Grant, Maria; Li, Zhimin; Peris, Joanna; King, Michael A; Hughes, Jeffrey A; Song, Sihong (2011). "The protective effects of plasma gelsolin on stroke outcome in rats". Experimental & Translational Stroke Medicine. 3 (1): 13. doi: 10.1186/2040-7378-3-13. ISSN  2040-7378. PMC  3224589. PMID  22047744.
  125. ^ Kevin Li-Chun, Hsieh; Schob, Stefan; Zeller, Matthias W.G.; Pulli, Benjamin; Ali, Muhammad; Wang, Cuihua; Chiou, Terry Ting-Yu; Tsang, Yuk-Ming; Lee, Po-Shun; Stossel, Thomas P.; Chen, John W. (October 2015). "Gelsolin decreases actin toxicity and inflammation in murine multiple sclerosis". Journal of Neuroimmunology. 287: 36–42. doi: 10.1016/j.jneuroim.2015.08.006. ISSN  0165-5728. PMC  4595933. PMID  26439960.
  126. ^ Hirko, Aaron C; Meyer, Edwin M; King, Michael A; Hughes, Jeffery A (September 2007). "Peripheral Transgene Expression of Plasma Gelsolin Reduces Amyloid in Transgenic Mouse Models of Alzheimer's Disease". Molecular Therapy. 15 (9): 1623–1629. doi: 10.1038/sj.mt.6300253. ISSN  1525-0016. PMID  17609655.
  127. ^ Matsuoka, Yasuji; Saito, Mitsuo; LaFrancois, John; Saito, Mariko; Gaynor, Kate; Olm, Vicki; Wang, Lili; Casey, Evelyn; Lu, Yifan; Shiratori, Chiharu; Lemere, Cynthia; Duff, Karen (2003-01-01). "Novel therapeutic approach for the treatment of Alzheimer's disease by peripheral administration of agents with an affinity to beta-amyloid". The Journal of Neuroscience. 23 (1): 29–33. doi: 10.1523/JNEUROSCI.23-01-00029.2003. ISSN  1529-2401. PMC  6742136. PMID  12514198.
  128. ^ Li, Mingjuan; Cui, Fengmei; Cheng, Ying; Han, Ling; Wang, Jia; Sun, Ding; Liu, Yu-long; Zhou, Ping-kun; Min, Rui (2014-08-28). "Gelsolin: role of a functional protein in mitigating radiation injury". Cell Biochemistry and Biophysics. 71 (1): 389–396. doi: 10.1007/s12013-014-0210-3. ISSN  1559-0283. PMID  25164111. S2CID  942471. Retrieved 2020-02-25.
  129. ^ Gawade, Shivaji P. (2012). "Acetic acid induced painful endogenous infliction in writhing test on mice". Journal of Pharmacology & Pharmacotherapeutics. 3 (4): 348. doi: 10.4103/0976-500X.103699 (inactive 2024-04-27). ISSN  0976-500X. PMC  3543562. PMID  23326113.{{ cite journal}}: CS1 maint: DOI inactive as of April 2024 ( link)
  130. ^ Gupta, Ashok Kumar; Parasar, Devraj; Sagar, Amin; Choudhary, Vikas; Chopra, Bhupinder Singh; Garg, Renu; Ashish; Khatri, Neeraj (2015-08-14). "Analgesic and Anti-Inflammatory Properties of Gelsolin in Acetic Acid Induced Writhing, Tail Immersion and Carrageenan Induced Paw Edema in Mice". PLOS ONE. 10 (8). Prasun K Datta (ed.): –0135558. Bibcode: 2015PLoSO..1035558G. doi: 10.1371/journal.pone.0135558. ISSN  1932-6203. PMC  4537109. PMID  26426535.
  131. ^ Khatri, Neeraj; Sagar, Amin; Peddada, Nagesh; Choudhary, Vikas; Chopra, Bhupinder Singh; Garg, Veena; Garg, Renu; Ashish (2014). "Plasma Gelsolin Levels Decrease in Diabetic State and Increase upon Treatment with F-Actin Depolymerizing Versions of Gelsolin". Journal of Diabetes Research. 2014: 152075. doi: 10.1155/2014/152075. ISSN  2314-6745. PMC  4247973. PMID  25478578.
  132. ^ BioAegis Therapeutics (14 January 2020). "A Phase 1b/2a Study of the Safety and Pharmacokinetics of Rhu-plasma Gelsolin in Hospitalized Subjects With CAP". ClinicalTrials.gov. U.S. National Library of Medicine. Retrieved 24 February 2020.
  133. ^ BioAegis Therapeutics. "Rhu-pGSN for Severe Covid-19 Pneumonia". ClinicalTrials.gov. U.S. National Library of Medicine. Retrieved 16 July 2020.

Videos

Youtube | Vimeo | Bing

Websites

Google | Yahoo | Bing

Encyclopedia

Google | Yahoo | Bing

Facebook