Your browser doesn't support javascript.
loading
Show: 20 | 50 | 100
Results 1 - 20 de 41
Filter
Add more filters










Publication year range
1.
Transfusion ; 59(8): 2643-2651, 2019 08.
Article in English | MEDLINE | ID: mdl-31135973

ABSTRACT

BACKGROUND: American Association of Blood Banks (AABB) guidelines suggest that packed red blood cells (PRBCs) be administered through a dedicated intravenous (IV) catheter. Literature supporting this broad-scope declaration are scarce. Obtaining additional IV access is painful, costly, and an infectious risk. We evaluated the effect of co-incubating PRBCs with crystalloids and medications on PRBC hemolysis, membrane deformability, and aggregation, as well as medication concentration. METHODS: PRBCs were co-incubated 5 minutes with plasma, normal saline (NS), 5% dextrose in water (D5W), Plasmalyte, epinephrine (epi), norepinephrine (norepi), dopamine (dopa), or Propofol (prop). Samples were then assessed for hemolysis (free hemoglobin, serum potassium), membrane deformability (elongation index [EI]), aggregation (smear, critical shear stress [mPa]) and drug concentration (High Performance Liquid Chromatography/Tandem Mass Spectrometry [LCMS-MS]). Significance (p ≤ 0.05) was determined by Wilcoxon-paired comparisons or Wilcoxon/Kruskall Willis with post-hoc Dunn's test. RESULTS: Compared to co-incubation with plasma: 1) co-incubation resulted in significantly increased hemolysis only when D5W as used (free hemoglobin, increased potassium); 2) EI trended lower when co-incubated with D5W and trended toward higher when co-incubated with prop; 3) aggregation was significantly lower when PRBCs co-incubated with NS, D5W, or Plasmalyte, and trended lower when co-incubated with epi, norepi, or dopa. Medication concentrations were between those predicted by distribution only in plasma and distribution through the entire intra- and extracellular space. CONCLUSION: Our data suggest that 5 minutes of PRBC incubation with isotonic crystalloids or catecholamines does not deleteriously alter PRBC hemolysis, membrane deformability, or aggregation. Co-incubation with D5W likely increases hemolysis. Propofol may promote hemolysis.


Subject(s)
Blood Banks , Crystalloid Solutions/pharmacology , Erythrocytes/metabolism , Hemolysis/drug effects , Crystalloid Solutions/chemistry , Erythrocyte Transfusion , Humans , Time Factors
2.
Front Physiol ; 9: 690, 2018.
Article in English | MEDLINE | ID: mdl-29930515

ABSTRACT

The primary role for erythrocytes is oxygen transport that requires the reversible binding of oxygen to hemoglobin. There are, however, secondary reactions whereby the erythrocyte can generate reactive oxygen species (ROS) and nitric oxide (NO). ROS such as superoxide anion and hydrogen peroxide are generated by the autoxidation of hemoglobin. NO can be generated when nitrite reacts with hemoglobin forming an HbNO+ intermediate. Both of these reactions are dramatically enhanced under hypoxic conditions. Within the erythrocyte, interactions of NO with hemoglobin and enzymatic reactions that neutralize ROS are expected to prevent the release of any generated NO or ROS. We have, however, demonstrated that partially oxygenated hemoglobin has a distinct conformation that enhances hemoglobin-membrane interactions involving Band 3 protein. Autoxidation of the membrane bound partially oxygenated hemoglobin facilitates the release of ROS from the erythrocyte. NO release is made possible when HbNO+, the hemoglobin nitrite-reduced intermediate, which is not neutralized by hemoglobin, is bound to the membrane and releases NO. Some of the released ROS has been shown to be transferred to the vasculature suggesting that some of the released NO may also be transferred to the vasculature. NO is known to have a major effect on the vasculature regulating vascular dilatation. Erythrocyte generated NO may be important when NO production by the vasculature is impaired. Furthermore, the erythrocyte NO released, may play an important role in regulating vascular function under hypoxic conditions when endothelial eNOS is less active. ROS can react with NO and, can thereby modulate the vascular effects of NO. We have also demonstrated an inflammatory response due to erythrocyte ROS. This reflects the ability of ROS to react with various cellular components affecting cellular function.

3.
Free Radic Biol Med ; 101: 296-304, 2016 12.
Article in English | MEDLINE | ID: mdl-27693379

ABSTRACT

S-nitrosothiols (SNO) perform many important functions in biological systems, but the mechanism by which they are generated in vivo remains a contentious issue. Nitric oxide (NO) reacts with thiols to form SNO only in the presence of a molecule that will accept an electron from either NO or the thiol. In this study, we present evidence that ferriheme accepts an electron from NO or glutathione (GSH) to generate S-nitrosoglutathione (GSNO) in vitro under anaerobic or hypoxic (2% O2) conditions. Ferriheme formed charge transfer-stable complexes with NO to form ferriheme-NO (heme-Fe(II)-NO+) and with GSH to form ferriheme-GS (heme-Fe(II)-GS•) under anaerobic conditions. The reaction between GSH and the heme-Fe(II)-NO+ complex or between NO and the heme-Fe(II)-GS• complex resulted in simultaneous reductive ferriheme nitrosylation (heme-Fe(II)NO) and the generation of GSNO. Thus, ferriheme is readily reduced to ferroheme in the presence of NO and GSH together, but not with either individually. The reaction between NO and the heme-Fe(II)-GS• complex to generate GSNO occurred more rapidly than NO was consumed by endothelial cells, but not red blood cells. In addition, pretreatment of endothelial cells with ferriheme or the ferriheme-GS complex generated SNO upon addition of NO under hypoxic conditions. The results of this study raise the possibility that in vivo, ferriheme can complex with GSH to form ferriheme-GS complex (heme-Fe(II)-GS•), which rapidly reacts with NO to generate GSNO under intracellular oxygen levels. The GSNO formation by this mechanism is more efficient than any other in vitro mechanism(s) reported so far.


Subject(s)
Endothelial Cells/drug effects , Glutathione/chemistry , Hemin/chemistry , Nitric Oxide/chemistry , S-Nitrosoglutathione/chemistry , Biocatalysis , Cell Hypoxia , Cells, Cultured , Endothelial Cells/cytology , Endothelial Cells/metabolism , Erythrocytes/cytology , Erythrocytes/drug effects , Erythrocytes/metabolism , Glutathione/metabolism , Hemin/pharmacology , Humans , Kinetics , Nitric Oxide/metabolism , Organ Specificity , Oxidation-Reduction , S-Nitrosoglutathione/metabolism
4.
Anesth Analg ; 123(2): 274-82, 2016 08.
Article in English | MEDLINE | ID: mdl-27308950

ABSTRACT

BACKGROUND: Cell-free hemoglobin (Hb) forms in stored red blood cells (RBCs) as a result of hemolysis. Studies suggest that this cell-free Hb may decrease nitric oxide (NO) bioavailability, potentially leading to endothelial dysfunction, vascular injury, and multiorgan dysfunction after transfusion. We tested the hypothesis that moderate doses of stored RBC transfusions increase cell-free Hb and decrease NO availability in postoperative surgical patients. METHODS: Twenty-six patients undergoing multilevel spine fusion surgery were studied. We compared those who received no stored RBCs (n = 9) with those who received moderate amounts (6.1 ± 3.0 units) of stored RBCs over 3 perioperative days (n = 17). Percent hemolysis (cell-free Hb), RBC-NO (heme-NO), and plasma nitrite and nitrate were measured in samples from the stored RBC bags and from patients' blood, before and after surgery. RESULTS: Posttransfusion hemolysis was increased approximately 3.5-fold over preoperative levels (P = 0.0002) in blood samples collected immediately after surgery but not on postoperative days 1 to 3. Decreases in both heme-NO (by approximately 50%) and plasma nitrite (by approximately 40%) occurred postoperatively, both in nontransfused patients (P = 0.036 and P = 0.026, respectively) and transfused patients (P = 0.0068 and P = 0.003, respectively) and returned to preoperative baseline levels by postoperative day 2 or 3. Postoperative plasma nitrite and nitrate were decreased significantly in both groups, and this change was slower to return to baseline in the transfused patients, suggesting that blood loss and hemodilution from crystalloid administration contribute to this finding. CONCLUSIONS: The decrease in NO metabolites occurred irrespective of stored RBC transfusions, suggesting this decrease may be related to blood loss during surgery and hemodilution rather than to scavenging of NO or inhibition of NO synthesis by stored RBC transfusions.


Subject(s)
Blood Loss, Surgical/prevention & control , Erythrocyte Transfusion/adverse effects , Erythrocytes/metabolism , Hemoglobins/metabolism , Hemolysis , Nitric Oxide/blood , Spinal Fusion/adverse effects , Adult , Aged , Aged, 80 and over , Biomarkers/blood , Down-Regulation , Female , Hemodilution/adverse effects , Humans , Male , Middle Aged , Nitrates/blood , Nitrites/blood , Time Factors , Treatment Outcome
5.
Transfusion ; 56(7): 1690-8, 2016 07.
Article in English | MEDLINE | ID: mdl-27062463

ABSTRACT

BACKGROUND: Clinical trials have shown that longer red blood cell (RBC) storage duration does not worsen outcomes; however, these studies included few RBCs near the end of the 42-day storage limit. We tested the hypothesis that these "oldest" RBCs are associated with adverse outcomes. STUDY DESIGN AND METHODS: In a retrospective study, 28,247 transfused patients given 129,483 RBC units were assessed. Morbidity, mortality, and length of stay (LOS) were compared in patients transfused exclusively with RBCs stored not more than 21 days versus patients transfused exclusively with RBCs stored 28 days or more and patients transfused exclusively with RBCs stored 35 days or more. RESULTS: After risk adjustment, ≥35-day RBCs were associated with increased morbidity (adjusted odds ratio [adjOR], 1.19; 95% confidence interval [CI], 1.07-1.32; p = 0.002), but ≥28-day RBCs were not (adjOR, 1.06; 95% CI, 0.97-1.15; p = 0.2). Neither ≥35-day nor ≥28-day RBCs were associated with increased mortality. In critically ill patients, ≥35-day RBCs were associated with increased morbidity (adjOR, 1.25; 95% CI, 1.08-1.44; p = 0.002) and mortality (adjOR, 1.38; 95% CI, 1.08-1.74; p = 0.009), but ≥28-day RBCs were associated with neither. In older patients, ≥35-day RBCs were associated with increased morbidity (adjOR, 1.22; 95% CI, 1.04-1.42; p = 0.01), but not mortality (adjOR, 1.28; 95% CI, 0.96-1.71; p = 0.1), and ≥28-day RBCs were associated with neither. LOS was increased for both ≥28- and ≥35-day RBCs for all patients and the critically ill and older subgroups. CONCLUSIONS: RBCs transfused in the last 7 days of their 42-day storage limit may be associated with adverse clinical outcomes in high-risk patients.


Subject(s)
Blood Preservation/standards , Erythrocyte Transfusion/adverse effects , Erythrocytes , Adult , Aged , Critical Illness , Female , Humans , Male , Middle Aged , Retrospective Studies , Risk , Time Factors
6.
Anesth Analg ; 122(3): 616-623, 2016 Mar.
Article in English | MEDLINE | ID: mdl-26891388

ABSTRACT

BACKGROUND: Stored red blood cells (RBCs) are deficient in 2,3-diphosphoglycerate (2,3-DPG), but it is unclear how autologous salvaged blood (ASB) compares with stored blood and how rapidly 2,3-DPG levels return to normal after transfusion. Therefore, we compared levels of 2,3-DPG in stored versus ASB RBCs and in patients' blood after transfusion. METHODS: Twenty-four patients undergoing multilevel spine fusion surgery were enrolled. We measured 2,3-DPG and the oxyhemoglobin dissociation curve (P50) in samples taken from the ASB and stored blood bags before transfusion and in blood samples drawn from patients before and after transfusion. RESULTS: The mean storage duration for stored RBCs was 24 ± 8 days. Compared with fresh RBCs, stored RBCs had decreased 2,3-DPG levels (by approximately 90%; P < 0.0001) and a decreased P50 (by approximately 30%; P < 0.0001). However, ASB RBCs did not exhibit these changes. The mean 2,3-DPG concentration decreased by approximately 20% (P < 0.05) in postoperative blood sampled from patients who received 1 to 3 stored RBC units and by approximately 30% (P < 0.01) in those who received ≥4 stored RBC units. 2,3-DPG was unchanged in patients who received no stored blood or ASB alone. After surgery, 2,3-DPG levels recovered gradually over 3 postoperative days in patients who received stored RBCs. CONCLUSIONS: Stored RBCs, but not ASB RBCs, have decreased levels of 2,3-DPG and a left-shift in the oxyhemoglobin dissociation curve. Postoperatively, 2,3-DPG levels remain below preoperative baseline levels for up to 3 postoperative days in patients who receive stored RBCs but are unchanged in those who receive only ASB RBCs.


Subject(s)
2,3-Diphosphoglycerate/blood , Blood Transfusion, Autologous , Erythrocytes/chemistry , Operative Blood Salvage , Adult , Aged , Blood Preservation , Erythrocyte Transfusion , Female , Humans , Male , Middle Aged , Oxyhemoglobins/analysis , Spinal Fusion
7.
Transfusion ; 56(5): 1101-11, 2016 05.
Article in English | MEDLINE | ID: mdl-26825863

ABSTRACT

BACKGROUND: The loss of structural and functional integrity of red blood cells (RBCs) during storage, collectively referred to as "storage lesion," has been implicated in reduced oxygen delivery after transfusion. RBCs are highly susceptible to oxidative damage from generation of reactive oxygen species by autoxidation of hemoglobin. Therefore, we examined whether increased oxidative stress (OS) in stored RBCs is associated with impaired cell membrane deformability before or after transfusion. STUDY DESIGN AND METHODS: Thirty-four patients undergoing multilevel spine fusion surgery were enrolled. OS in RBCs was assessed by the presence of fluorescent heme degradation products and methemoglobin, which were measured with fluorimetric and spectrophotometric methods, respectively. Deformability and aggregation were determined by ektacytometry in stored RBCs, autologous salvaged RBCs, and posttransfusion blood samples. RESULTS: OS in stored RBCs was significantly increased with longer storage (R = 0.54, p = 0.032) and significantly higher than that in fresh RBCs (9.1 ± 1.3 fluorescent arbitrary units vs. 7.7 ± 0.9 fluorescent arbitrary units, p < 0.001). Deformability decreased (R = -0.60, p = 0.009) with increasing storage duration. OS was elevated (p < 0.05) and deformability was decreased (p < 0.05) in postoperative blood from patients who had undergone moderate (≥4 RBC units) but not minimal or no transfusion. Neither the decrease in deformability of RBCs nor the aggregation changes were correlated with OS. CONCLUSIONS: Although stored RBCs show signs of increased OS and loss of cell membrane deformability, these changes were not directly correlated and were only evident after moderate but not lower dose transfusion in postoperative surgical patients. These findings suggest that factors other than OS may contribute to impaired rheology with stored RBCs in the clinical setting.


Subject(s)
Erythrocyte Deformability , Erythrocyte Transfusion , Erythrocytes/cytology , Hemorheology , Oxidative Stress , Adult , Aged , Aged, 80 and over , Blood Preservation/methods , Female , Humans , Male , Middle Aged , Postoperative Care
8.
Biochemistry ; 54(44): 6712-23, 2015 Nov 10.
Article in English | MEDLINE | ID: mdl-26478948

ABSTRACT

The reduction of nitrite by deoxyhemoglobin to nitric oxide (NO) has been proposed as a mechanism for the transfer of NO bioactivity from the red blood cell (RBC) to the vasculature. This transfer can increase vascular dilatation. The major challenge to this hypothesis is the very efficient scavenging of NO by hemoglobin, which prevents the release of NO from RBCs. Previous studies indicate that the reaction of nitrite with deoxyhemoglobin produces two metastable intermediates involving nitrite bound to deoxyhemoglobin and a hybrid intermediate [Hb(II)NO(+) ↔ Hb(III)NO] where the nitrite is reduced, but unavailable to react with hemoglobin. We have now shown how unique properties of these intermediates provide a pathway for the release of NO bioactivity from RBCs. The high membrane affinity of these intermediates (>100-fold greater than that of deoxyhemoglobin) places these intermediates on the membrane. Furthermore, membrane-induced conformational changes of the nitrite-reacted intermediates facilitate the release of NO from the hybrid intermediate and nitrite from the nitrite-bound intermediate. Increased membrane affinity, coupled with facilitated dissociation of NO and nitrite from the membrane-bound intermediates, provides the first realistic mechanism for the potential release of NO and nitrite from the RBC and their potential transfer to the vasculature.


Subject(s)
Erythrocyte Membrane/metabolism , Hemoglobins/metabolism , Nitric Oxide/metabolism , Nitrites/metabolism , Humans , Oxidation-Reduction
9.
Nitric Oxide ; 40: 117-22, 2014 Aug 31.
Article in English | MEDLINE | ID: mdl-24973574

ABSTRACT

BACKGROUND: Intermittent pneumatic compression (IPC) of legs exerts beneficial local vascular effects, possibly through local release of nitric oxide (NO). However, studies demonstrating systemic transport of nitrogen oxide species and release of NO prompt the question of whether IPC could also exert nonlocal effects. We tested whether IPC (1) affects systemic levels of nitrite, S-nitrosothiols and red blood cell (RBC) NO, and (2) exerts vasoactive effects in the brachial artery (BA), although this hypothesis-generating pilot study did not investigate cause and effect relationship between (1) and (2). METHODS: In 10 healthy subjects, ages 24-39 years, we measured plasma nitrite, plasma S-nitrosothiols and RBC-NO from venous blood samples drawn before and after IPC treatment. We also measured BA responses to 5 min of upper arm occlusion at rest and during 1 h of leg IPC. RESULTS: There was a significant decrease in plasma nitrite (112±26 nM to 90±15 nM, p=0.0008) and RBC-NO (129±72 nM to 102±41 nM, p=0.02). Plasma S-nitrosothiols were unchanged (5.79±4.81 nM to 6.27±5.79 nM, p=0.3). BA occlusion-mediated constriction (OMC) was significantly attenuated with IPC treatment (-43±13% to -33±12%, p=0.003). High-flow mediated BA dilation was unchanged (13.3±9.4% to 11.5±7.2%, p=0.2). CONCLUSION: Plasma nitrite, RBC-NO, and BA OMC decreased with leg IPC. We hypothesize that this decrease in circulatory pool of plasma nitrite and RBC-NO may result from the transfer of their NO-bioactivity from blood to the hypoxic arm tissue, to be stored and released under hypoxic stress and oppose OMC. Future studies should investigate whether IPC-induced decreases in brachial OMC are caused by the changes in systemic NO activity, and whether leg IPC could benefit distant arterial function in systemic cardiovascular disease.


Subject(s)
Arm/blood supply , Intermittent Pneumatic Compression Devices , Leg/blood supply , Nitric Oxide/blood , Adult , Female , Humans , Male , Nitric Oxide/metabolism , Pilot Projects , Young Adult
10.
Front Physiol ; 5: 84, 2014.
Article in English | MEDLINE | ID: mdl-24616707

ABSTRACT

Red Blood Cells (RBCs) need to deform and squeeze through narrow capillaries. Decreased deformability of RBCs is, therefore, one of the factors that can contribute to the elimination of aged or damaged RBCs from the circulation. This process can also cause impaired oxygen delivery, which contributes to the pathology of a number of diseases. Studies from our laboratory have shown that oxidative stress plays a significant role in damaging the RBC membrane and impairing its deformability. RBCs are continuously exposed to both endogenous and exogenous sources of reactive oxygen species (ROS) like superoxide and hydrogen peroxide (H2O2). The bulk of the ROS are neutralized by the RBC antioxidant system consisting of both non-enzymatic and enzymatic antioxidants including catalase, glutathione peroxidase and peroxiredoxin-2. However, the autoxidation of hemoglobin (Hb) bound to the membrane is relatively inaccessible to the predominantly cytosolic RBC antioxidant system. This inaccessibility becomes more pronounced under hypoxic conditions when Hb is partially oxygenated, resulting in an increased rate of autoxidation and increased affinity for the RBC membrane. We have shown that a fraction of peroxyredoxin-2 present on the RBC membrane may play a major role in neutralizing these ROS. H2O2 that is not neutralized by the RBC antioxidant system can react with the heme producing fluorescent heme degradation products (HDPs). We have used the level of these HDP as a measure of RBC oxidative Stress. Increased levels of HDP are detected during cellular aging and various diseases. The negative correlation (p < 0.0001) between the level of HDP and RBC deformability establishes a contribution of RBC oxidative stress to impaired deformability and cellular stiffness. While decreased deformability contributes to the removal of RBCs from the circulation, oxidative stress also contributes to the uptake of RBCs by macrophages, which plays a major role in the removal of RBCs from circulation. The contribution of oxidative stress to the removal of RBCs by macrophages involves caspase-3 activation, which requires oxidative stress. RBC oxidative stress, therefore, plays a significant role in inducing RBC aging.

11.
Blood Cells Mol Dis ; 52(4): 230-5, 2014 Apr.
Article in English | MEDLINE | ID: mdl-24246527

ABSTRACT

Sickle cell disease (SCD) is associated with increase in oxidative stress and irreversible membrane changes that originates from the instability and polymerization of deoxygenated hemoglobin S (HbS). The relationship between erythrocyte membrane changes as assessed by a decrease in deformability and oxidative stress as assessed by an increase in heme degradation was investigated. The erythrocyte deformability and heme degradation for 27 subjects with SCD and 7 with sickle trait were compared with normal healthy adults. Changes in both deformability and heme degradation increased in the order of control to trait to non-crisis SCD to crisis SCD resulting in a very significantly negative correlation between deformability and heme degradation. However, a quantitative analysis of the changes in deformability and heme degradation for these different groups of subjects indicated that sickle trait had a much smaller effect on deformability than on heme degradation, while crisis affects deformability to a greater extent than heme degradation. These findings provide insights into the relative contributions of erythrocyte oxidative stress and membrane damage during the progression of SCD providing a better understanding of the pathophysiology of SCD.


Subject(s)
Anemia, Sickle Cell/metabolism , Anemia, Sickle Cell/pathology , Erythrocyte Deformability , Erythrocytes, Abnormal/metabolism , Erythrocytes/metabolism , Erythrocytes/pathology , Oxidative Stress , Adolescent , Adult , Anemia, Sickle Cell/diagnosis , Child , Female , Fetal Hemoglobin/metabolism , Heme/metabolism , Hemoglobin, Sickle/metabolism , Hemoglobins/metabolism , Humans , Male , Proteolysis , Sickle Cell Trait/diagnosis , Sickle Cell Trait/metabolism , Sickle Cell Trait/pathology , Young Adult
12.
Front Physiol ; 5: 500, 2014.
Article in English | MEDLINE | ID: mdl-25642190

ABSTRACT

Hemoglobin (Hb) continuously undergoes autoxidation producing superoxide which dismutates into hydrogen peroxide (H2O2) and is a potential source for subsequent oxidative reactions. Autoxidation is most pronounced under hypoxic conditions in the microcirculation and for unstable dimers formed at reduced Hb concentrations. In the red blood cell (RBC), oxidative reactions are inhibited by an extensive antioxidant system. For extracellular Hb, whether from hemolysis of RBCs and/or the infusion of Hb-based blood substitutes, the oxidative reactions are not completely neutralized by the available antioxidant system. Un-neutralized H2O2 oxidizes ferrous and ferric Hbs to Fe(IV)-ferrylHb and OxyferrylHb, respectively. FerrylHb further reacts with H2O2 producing heme degradation products and free iron. OxyferrylHb, in addition to Fe(IV) contains a free radical that can undergo additional oxidative reactions. Fe(III)Hb produced during Hb autoxidation also readily releases heme, an additional source for oxidative stress. These oxidation products are a potential source for oxidative reactions in the plasma, but to a greater extent when the lower molecular weight Hb dimers are taken up into cells and tissues. Heme and oxyferryl have been shown to have a proinflammatory effect further increasing their potential for oxidative stress. These oxidative reactions contribute to a number of pathological situations including atherosclerosis, kidney malfunction, sickle cell disease, and malaria. The toxic effects of extracellular Hb are of particular concern with hemolytic anemia where there is an increase in hemolysis. Hemolysis is further exacerbated in various diseases and their treatments. Blood transfusions are required whenever there is an appreciable decrease in RBCs due to hemolysis or blood loss. It is, therefore, essential that the transfused blood, whether stored RBCs or the blood obtained by an Autologous Blood Recovery System from the patient, do not further increase extracellular Hb.

13.
Cell Biochem Biophys ; 67(2): 385-98, 2013 Nov.
Article in English | MEDLINE | ID: mdl-22161622

ABSTRACT

S-nitrosothiols (RSNO) are involved in post-translational modifications of many proteins analogous to protein phosphorylation. In addition, RSNO have many physiological roles similar to nitric oxide ((•)NO), which are presumably involving the release of (•)NO from the RSNO. However, the much longer life span in biological systems for RSNO than (•)NO suggests a dominant role for RSNO in mediating (•)NO bioactivity. RSNO are detected in plasma in low nanomolar levels in healthy human subjects. These RSNO are believed to be redirecting the (•)NO to the vasculature. However, the mechanism for the formation of RSNO in vivo has not been established. We have reviewed the reactions of (•)NO with oxygen, metalloproteins, and free radicals that can lead to the formation of RSNO and have evaluated the potential for each mechanism to provide a source for RSNO in vivo.


Subject(s)
Blood/metabolism , S-Nitrosothiols/blood , Blood Cells/metabolism , Humans
14.
Exp Hematol ; 41(3): 316-21, 2013 Mar.
Article in English | MEDLINE | ID: mdl-23142655

ABSTRACT

Among the three types of super oxide dismutases (SODs) known, SOD2 deficiency is lethal in neonatal mice owing to cardiomyopathy caused by severe oxidative damage. SOD2 is found in red blood cell (RBC) precursors, but not in mature RBCs. To investigate the potential damage to mature RBCs resulting from SOD2 deficiency in precursor cells, we studied RBCs from mice in which fetal liver stem cells deficient in SOD2 were capable of efficiently rescuing lethally irradiated host animals. These transplanted animals lack SOD2 only in hematopoietically generated cells and live longer than SOD2 knockouts. In these mice, approximately 2.8% of their total RBCs in circulation are iron-laden reticulocytes, with numerous siderocytic granules and increased protein oxidation similar to that seen in sideroblastic anemia. We have studied the RBC deformability and oxidative stress in these animals and the control group by measuring them with a microfluidic ektacytometer and assaying fluorescent heme degradation products with a fluorimeter, respectively. In addition, the rate of hemoglobin oxidation in RBCs from these mice and the control group were measured spectrophotometrically. The results show that RBCs from these SOD2-deficient mice have reduced deformability, increased heme degradation products, and an increased rate of hemoglobin oxidation compared with control animals, indicative of increased RBC oxidative stress.


Subject(s)
Erythrocyte Deformability/physiology , Hematopoietic Stem Cells/physiology , Heme/metabolism , Superoxide Dismutase/deficiency , Animals , Erythrocyte Deformability/genetics , Erythrocytes/enzymology , Erythrocytes/metabolism , Erythrocytes/physiology , Erythroid Precursor Cells/enzymology , Erythroid Precursor Cells/metabolism , Erythroid Precursor Cells/physiology , Hematopoietic Stem Cells/enzymology , Hematopoietic Stem Cells/metabolism , Hemoglobins/metabolism , Mice , Mice, Knockout , Mice, Transgenic , Oxidation-Reduction , Oxidative Stress/physiology , Peroxiredoxins/deficiency , Peroxiredoxins/genetics , Spectrometry, Fluorescence , Superoxide Dismutase/genetics
15.
Antioxid Redox Signal ; 18(17): 2274-83, 2013 Jun 10.
Article in English | MEDLINE | ID: mdl-23025272

ABSTRACT

SIGNIFICANCE: The physiological mechanism(s) for recognition and removal of red blood cells (RBCs) from circulation after 120 days of its lifespan is not fully understood. Many of the processes thought to be associated with the removal of RBCs involve oxidative stress. We have focused on hemoglobin (Hb) redox reactions, which is the major source of RBC oxidative stress. RECENT ADVANCES: The importance of Hb redox reactions have been shown to originate in large parts from the continuous slow autoxidation of Hb producing superoxide and its dramatic increase under hypoxic conditions. In addition, oxidative stress has been shown to be associated with redox reactions that originate from Hb reactions with nitrite and nitric oxide (NO) and the resultant formation of highly toxic peroxynitrite when NO reacts with superoxide released during Hb autoxidation. CRITICAL ISSUES: The interaction of Hb, particularly under hypoxic conditions with band 3 of the RBC membrane is critical for the generating the RBC membrane changes that trigger the removal of cells from circulation. These changes include exposure of antigenic sites, increased calcium leakage into the RBC, and the resultant leakage of potassium out of the RBC causing cell shrinkage and impaired deformability. FUTURE DIRECTIONS: The need to understand the oxidative damage to specific membrane proteins that result from redox reactions occurring when Hb is bound to the membrane. Proteomic studies that can pinpoint the specific proteins damaged under different conditions will help elucidate the cellular aging processes that result in cells being removed from circulation.


Subject(s)
Cellular Senescence/physiology , Erythrocytes/metabolism , Hemoglobins/metabolism , Animals , Antioxidants/metabolism , Erythrocyte Membrane/metabolism , Humans , Oxidation-Reduction , Oxidative Stress , Protein Binding , Reactive Nitrogen Species/metabolism , Reactive Oxygen Species/metabolism
16.
Nitric Oxide ; 24(2): 102-9, 2011 Mar 15.
Article in English | MEDLINE | ID: mdl-21236353

ABSTRACT

Deoxyhemoglobin reduces nitrite to nitric oxide (NO). In order to study the effect of the hemoglobin quaternary conformation on the nitrite reaction, we compared T-state deoxyhemoglobin with R-state deoxyhemoglobin produced by reacting hemoglobin with carboxypeptidase-A prior to deoxygenation. The nitrite reaction with deoxyhemoglobin was followed by chemiluminescence, electron paramagnetic resonance and visible spectroscopy. The initial steps in this reaction involve the binding of nitrite to deoxyhemoglobin followed by the formation of an electron delocalized metastable intermediate that retains potential NO bioactivity. This reaction is shown by visible spectroscopy to occur 5.6 times faster in the R-state than in the T-state. However, the dissociation of NO from the delocalized intermediate is shown to be facilitated by the T-quaternary conformation with a 9.6 fold increase in the rate constant. The preferred NO-release in the T-state, which has a higher affinity for the membrane, can result in the NO diffusing out of the RBC and being released to the vasculature at low partial pressures of oxygen.


Subject(s)
Hemoglobins/chemistry , Nitric Oxide/metabolism , Nitrites/metabolism , Animals , Hemoglobins/pharmacology , Humans , Nitric Oxide/chemistry , Nitrites/chemistry , Oxidation-Reduction , Protein Structure, Quaternary
17.
Methods Mol Biol ; 704: 27-37, 2011.
Article in English | MEDLINE | ID: mdl-21161627

ABSTRACT

S-nitrosothiols present in nanomolar concentrations in cells and body fluids play an important role in vasodilation, in preventing platelet aggregation, leukocyte adhesion, and for cellular signaling. However, because of the low levels of s-nitrosothiols and interference with other nitric oxide species, reliable assays that measure both high molecular weight and low molecular weight s-nitrosothiols in plasma and red blood cells red blood cells have been difficult to develop. We have previously developed a sensitive method using Cu(II)-ascorbic acid Cu(II)-ascorbic acid at a neutral pH, which was specific for s-nitrosothiols without interference of nitrite or other NOx species. However, due to neutral pH foaming, this method was not suitable for determinations in plasma or red blood cells with high protein content. This method has now been modified by using copper (II) chloride (CuCl(2)) and ascorbic acid in glacial acetic acid. The low pH solves the foaming problem. However, protonation of nitrite under acidic conditions facilitates the formation of s-nitrosothiols. For this method to specifically measure s-nitrosothiols in the sample, the unreacted thiols are blocked by reacting with N-ethylmaleimide and nitrite is blocked by reacting with acidified sulfanilamide before being analyzed by chemiluminescence. Using this method, s-nitrosothiols have been determined in the range of 2 nM to 26 nM (mean ± SE = 10.18±2.1) in plasma and up to 88.1 nM (mean ± SE = 51.27 ± 10.5) in red blood cells.


Subject(s)
Body Fluids/chemistry , Luminescent Measurements/methods , S-Nitrosothiols/blood , Humans , Luminescence , S-Nitrosothiols/analysis
18.
Life Sci ; 86(3-4): 133-8, 2010 Jan 16.
Article in English | MEDLINE | ID: mdl-19958781

ABSTRACT

AIMS: Red blood cells (RBCs) have an extensive antioxidant system designed to eliminate the formation of reactive oxygen species (ROS). Nevertheless, RBC oxidant stress has been demonstrated by the formation of a fluorescent heme degradation product (excitation (ex) 321 nm, emission (em) 465 nm) both in vitro and in vivo. We investigated the possibility that the observed heme degradation results from ROS generated on the membrane surface that are relatively inaccessible to the cellular antioxidants. MAIN METHODS: Membrane and cytosol were separated by centrifugation and the fluorescence intensity and emission maximum were measured. The effect on the maximum emission of adding oxidized and reduced hemoglobin to the fluorescent product formed when hemin is degraded by hydrogen peroxide (H(2)O(2)) was studied. KEY FINDINGS: 90% of the fluorescent heme degradation products in hemolysates are found on the membrane. Furthermore, these products are not transferred from the cytosol to the membrane and must, therefore, be formed on the membrane. We also showed that the elevated level of heme degradation in HbCC cells that is attributed to increased oxidative stress was found on the membrane. SIGNIFICANCE: These results suggest that, although ROS generated in the cytosol are neutralized by antioxidant enzymes, H(2)O(2) generated by the membrane bound hemoglobin is not accessible to the cytosolic antioxidants and reacts to generate fluorescent heme degradation products. The formation of H(2)O(2) on the membrane surface can explain the release of ROS from the RBC to other tissues and ROS damage to the membrane that can alter red cell function and lead to the removal of RBCs from circulation by macrophages.


Subject(s)
Cell Membrane/metabolism , Erythrocytes/metabolism , Heme/metabolism , Hemoglobins/metabolism , Oxidative Stress , Cells, Cultured , Cytosol/metabolism , Erythrocyte Aging , Erythrocytes/cytology , Fluorescence , Humans , Hydrogen Peroxide/pharmacology , Oxidation-Reduction
19.
Methods Mol Biol ; 610: 41-9, 2010.
Article in English | MEDLINE | ID: mdl-20013171

ABSTRACT

Studies have demonstrated that plasma nitrite (N(O-)(2)) reflects endothelial nitric oxide (NO) production. In addition, N(O-)(2) has been shown to have biological activities associated with its reduction to NO in blood and tissues. Therefore, determination of plasma N(O-)(2) has been proposed as a prognostic marker for cardiovascular diseases. Typical concentrations of N(O-)(2) in the plasma are in the nanomolar range and determination of this N(O-)(2) poses a challenge in terms of both sensitivity and specificity. Thus, a highly sensitive, chemiluminescence method that is based on the reduction of N(O-)(2) by potassium iodide and iodine is being used to determine the nitrite in biological fluids. This method has the sensitivity, but also measures other nitric oxide species such as S-nitrosothiols and N-nitrosamines. We, therefore, developed an alternative method based on the reduction of N(O-)(2) by ascorbic acid in strongly acidic media. As part of the methodology, glacial acetic acid and ascorbic acid are introduced into the purge vessel of the NO analyzer. Samples containing N(O-)(2) are injected into the purge vessel and the chemiluminescence signals generated as a result of the formation of NO are then measured. We find that under these conditions N(O-)(2) is stoichiometrically reduced to NO. Other traditional NO-generating species, such as S-nitrosothiols, N-nitrosamines, nitrated lipids, and nitrated proteins, did not interfere in the determination of plasma N(O-)(2). Using the present method, plasma N(O-)(2) in fasting human subjects has been determined to be in the range of 56-210 nM (mean +/- SD = 110 +/- 36 nM; n = 8).


Subject(s)
Luminescent Measurements/methods , Nitrites/blood , Area Under Curve , Calibration , Humans , Luminescent Measurements/instrumentation , Luminescent Measurements/standards , Nitric Oxide/blood , Sensitivity and Specificity
20.
Methods Mol Biol ; 610: 165-80, 2010.
Article in English | MEDLINE | ID: mdl-20013178

ABSTRACT

Reactive oxygen species are implicated in many human diseases and aging process. Much of the evidence is based on experimental data indicating increasing rates of lipid peroxidation in disease states and the ameliorating effects of antioxidants. It is becoming increasingly evident that the natural antioxidants, which have phenolic structure, play an important role in protecting the tissues against free radical damage. Eugenol (4-allyl-2 methoxyphenol) is one such naturally occurring phenolic compound. The antioxidant activity of eugenol was evaluated by the extent of protection offered against free radical-mediated lipid peroxidation using both in vitro and in vivo studies. The in vitro lipid peroxidation was induced in mitochondria by (Fe(II)-ascorbate) or (Fe(II) + H(2)O(2)). The lipid peroxidation was assessed colorimetrically by measuring the formation of thiobarbituric acid reactive substances (TBARS) following the reaction of oxidized lipids with TBA. Eugenol completely inhibited both iron and Fenton reagent-mediated lipid peroxidation. The inhibitory activity of eugenol was about fivefold higher than that observed for alpha-tocopherol and about tenfold less than that observed for BHT. The in vivo lipid peroxidation-mediated liver damage was induced by administration of CCl(4) to rats. Eugenol significantly inhibited the rise in SGOT activity and cell necrosis without protecting the endoplasmic reticulum (ER) damage as assessed by its failure to prevent a decrease in cytochrome p450 and G-6-phosphatase activity. The protective action of eugenol has been found to be due to interception of secondary radicals derived from ER lipids rather than interfering with primary radicals of CCl(4) (CCl(3)/CCl(3)OO).


Subject(s)
Antioxidants/metabolism , Antioxidants/pharmacology , Eugenol/metabolism , Eugenol/pharmacology , Lipid Peroxidation/drug effects , Animals , Antioxidants/chemistry , Eugenol/chemistry , Free Radical Scavengers/metabolism , Glucose-6-Phosphatase/metabolism , Humans , Liver/drug effects , Liver/metabolism , Liver/pathology , Malondialdehyde/metabolism , Molecular Structure , Plant Extracts/chemistry , Rats , Rats, Wistar , Thiobarbituric Acid Reactive Substances/metabolism
SELECTION OF CITATIONS
SEARCH DETAIL
...