Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2020 Apr 1.
Published in final edited form as: Free Radic Biol Med. 2019 Feb 13;134:581–597. doi: 10.1016/j.freeradbiomed.2019.02.003

A thermodynamically-constrained mathematical model for the kinetics and regulation of NADPH oxidase 2 complex-mediated electron transfer and superoxide production

Namrata Tomar a, Shima Sadri a, Allen W Cowley Jr b, Chun Yang b, Nabeel Quryshi a, Venkat R Pannala a, Said H Audi c,a, Ranjan K Dash a,b,c,*
PMCID: PMC6588456  NIHMSID: NIHMS1013519  PMID: 30769160

Abstract

Reactive oxygen species (ROS) play an important role in cell signaling, growth, and immunity. However, when produced in excess, they are toxic to the cell and lead to premature aging and a myriad of pathologies, including cardiovascular and renal diseases. A major source of ROS in many cells is the family of NADPH oxidase (NOX), comprising of membrane and cytosolic components. NOX2 is among the most widely expressed and well-studied NOX isoform. Although details on the NOX2 structure, its assembly and activation, and ROS production are well elucidated experimentally, there is a lack of a quantitative and integrative understanding of the kinetics of NOX2 complex, and the various factors such as pH, inhibitory drugs, and temperature that regulate the activity of this oxidase. To this end, we have developed here a thermodynamically-constrained mathematical model for the kinetics and regulation of NOX2 complex based on diverse published experimental data on the NOX2 complex function in cell-free and cell-based assay systems. The model incorporates (i) thermodynamics of electron transfer from NADPH to O2 through different redox centers of the NOX2 complex, (ii) dependence of the NOX2 complex activity upon pH and temperature variations, and (iii) distinct inhibitory effects of different drugs on the NOX2 complex activity. The model provides the first quantitative and integrated understanding of the kinetics and regulation of NOX2 complex, enabling simulation of diverse experimental data. The model also provides several novel insights into the NOX2 complex function, including alkaline pH-dependent inhibition of the NOX2 complex activity by its reaction product NADP+. The model provides a mechanistic framework for investigating the critical role of NOX2 complex in ROS production and its regulation of diverse cellular functions in health and disease. Specifically, the model enables examining the effects of specific targeting of various enzymatic sources of pathological ROS which could overcome the limitations of pharmacological efforts aimed at scavenging ROS which has resulted in poor outcomes of antioxidant therapies in clinical studies.

Keywords: NOX2 complex, Phagocyte oxidase, Competitive and uncompetitive inhibition, Oxidation-reduction reaction, Reactive oxygen species, Oxidative stress

1. Introduction

Reactive oxygen species (ROS; e.g. superoxide: O2, hydrogen peroxide: H2O2) are known to play a critical role in cell signaling, which regulates gene expression, cell proliferation and division, cell immunity, and cell death [14]. They are generated under both physiological and pathophysiological conditions, albeit in excess under pathophysiological conditions, in almost all tissues and organs of multicellular organisms. In addition, when produced in excess, they are known to have injurious effects on biomolecules (e.g. DNA, proteins, and lipids), resulting in organelles, cells, and tissues/organs dysfunction and diseases [58].

There are several cellular sources of ROS, but the family of NADPH oxidase (NOX) represents the major non-mitochondrial source [912]. There is ample evidence that ROS generated by NOX play a key role in the pathogenesis of many chronic diseases, including atherosclerosis, heart disease, hypertension, diabetic nephropathy, chronic kidney disease, lung fibrosis, cancer, and Alzheimer’s disease, among others [1218]. Their existence in neutrophils and other phagocytes was first identified by Baehner et al. [19]. It is now well recognized that, in phagocytic cells, the NOX-generated ROS (primarily O2) play an important role in the cellular innate immune responses [20,21]. The NOX enzyme expressed in phagocytic cells (NOX2) upon stimulation generates a huge amount of O2, so called “respiratory burst”, to kill pathogens [22]. Relevant to the kinetics of this system, which is explored in the present study, the NOX2 enzyme when stimulated in human neutrophils can generate ∼10 nmol O2/min/106 cells, which requires, remarkably, the transfer of ∼108 electrons/s/cell or a current of ∼16 pA [2325]. In non-phagocytic cells, the NOX enzymes are of enormous importance with ROS affecting nearly every cellular process yet studied, including transcription factor activation, protein synthesis, cell division and proliferation, nearly all metabolic processes, and apoptosis [13,915].

Seven NOX isoforms have been identified, each containing at least six trans-membrane helices, two iron-heme groups, a FAD molecule, and a NADPH-binding domain [10,12]. Each of these NOX isoforms differs in composition, modes of activation, and its enzymatic reaction products. NOX2 is the most studied of NOXs, and is found in a variety of cell types of mesodermal origin. It consists of a group of plasma membrane proteins (e.g. gp91phox, p22phox) and several cytosolic proteins (e.g. p47phox, p40phox, p67phox, Rac-GTP) (Fig. 1A) [2632]; note that phox refers to a phagocytic oxidase. The membrane-bound NOX2 components remain dormant in unstimulated cells, but become activated when complexed with the cytosolic components in response to stimulation by various neuroendocrine, paracrine, microorganisms, or inflammatory mediators. The assembled and activated NOX2 complex mediates the transfer of electrons from substrate NADPH to molecular O2 through the action of an integral membrane-bound protein, namely flavocytochrome b245 or b558 (FlavoCytB), within the phagocytic oxidase subunits gp91phox and p22phox (Fig. 1B) [3335]. Flavo-CytB contains all the catalytic machinery for oxidation of NADPH to NADP+ and reduction of O2 to O2. However, it is only facilitated when the membrane subunits are complexed with the cytosolic subunits, leading to NOX2 assembly and activation [2632].

Fig. 1. Proposed kinetic mechanisms for the NOX2 complex-mediated electron flow and superoxide production.

Fig. 1

(A) Schematics of unassembled and inactivated NOX2 enzyme showing different cytosolic (p67phox, p47phox, p40phox, and Rac-GTP) and membrane-bound (gp91phox and p22phox) subunits. (B) Schematics of an assembled and activated NOX2 complex showing different cytosolic subunits complexed with the membrane-bound subunits, facilitating electron transfer from substrate NADPH to molecular oxygen (O2) resulting in superoxide (O2) production. (C) Schematics shows the five elementary electron transfer reactions and the associated midpoint redox potentials from NADPH through different redox centers of the NOX2 complex, and ultimately to O2 for O2 production. The NOX2 subunit gp91phox encompass an NADPH binding site and two redox centers, viz. a FAD-containing flavoprotein and a heme-containing heterodimeric cytochrome b245 (CytB), together termed as flavocytochrome b245 (FlavoCytB). In the first step, two electrons are transferred from NADPH to FAD to reduce FAD to FADH2. Subsequently, FADH2 provides two electrons to two heme sites of two CytBox in two different steps via the semiradical intermediate FADH to produce two CytBred. Finally, each CytBred provides an electron to O2 to form O2 via sequential one electron reduction of O2 [38] to complete the catalytic cycle. (D) The five-state catalytic scheme corresponding to the electron transfer reactions schematized in Panel C. The five different states (En.X, n = 1, …, 5) of the NOX2 complex (E.X) are based on the redox status of FAD and 2CytB. (E) Lumped electron transfer reaction schematics in which the two elementary electron transfer reactions from FADH2 to 2CytBox via FADH and from 2CytBred to 2O2 are lumped. (F) The three-state catalytic scheme corresponding to the electron transfer reactions schematized in Panel E. The three different states (En.X, n = 1, …, 3) of the NOX2 complex (E.X) are based on the redox status of FAD and 2CytB; FADH.CytBox.CytBred and FAD.CytBred. CytBox states are neglected from the 5-state catalytic scheme of Panel D.

Several early studies have evaluated the kinetics of NOX2 enzyme and NADPH-dependent O2 production in cell-free and cell-based assay systems, and have reported the Km of NADPH to vary widely (high Km of 40–300 μM in cell-free systems vs. as low as Km of 5–15 μM in whole-cell systems) [36,37]. Nisimoto et al. [38] recently examined the O2 dependency of NOX2 complex-mediated O2 generation in intact human neutrophils and cell-free assay systems and found the Km of O2 to be 3.1 and 2.3% O2 (22.1 and 16.4 mmHg), respectively. These values are within the range of Km values reported for enzymes contributing to cellular housekeeping functions. Also relevant to the current study, it is recognized that the NOX2 enzyme, as most others, are sensitive to pH and temperature, working optimally at neutral pH and physiological temperature. Few early studies have demonstrated high NOX2 enzyme activity in cell-free systems at pH in the vicinity of 7.0, with NADPH as a substrate [39,40]. Subsequently, Morgan et al. reported that the activated NOX2 enzyme-mediated electron current (proportional to O2 production) across the plasma membrane is maximal at around pH = 7.2, decreasing at higher or lower pH values [41]. Other related studies have also measured electron current generated by the activated NOX2 enzyme across the plasma membrane at fixed temperature [2325] and at varying temperatures [42]. Also relevant to the present study, few NOX2 inhibitors have been developed and their modes of actions have been characterized in efforts to assess their potential use as therapeutic agents [4345].

Although details on the NOX2 enzyme structure, its assembly and activation, electron transfer, and O2 production have been extensively studied, there is a lack of a quantitative and integrated understanding of the kinetic mechanisms of NOX2 complex-mediated electron transfer and O2 generation, and their regulations by pH, inhibitory drugs, and temperature. To this end, we have developed here a thermo-dynamically-constrained mathematical model to explore the catalytic mechanisms and regulations of NOX2 complex based on a diverse set of published experimental data obtained from cell-free and cell-based assay systems [3742,45]. The model development and parameterization use strict thermodynamics of elementary electron transfer reactions based on midpoint redox potentials of associated redox centers to constrain some of the kinetic parameters, while fixing some parameters at their expected values and estimating the remaining parameters in a modular and multi-step fashion based on the available data. The proposed model is the first thermodynamically rigorous kinetic model of the complex enzymatic function of NOX2 that describes well all the existing data, and provides critical quantitative understanding of the roles of pH, temperature, and inhibitory mechanisms of various drugs on the NOX2 enzymatic function under physiological and pathophysiological conditions.

2. Materials and methods

Proposed kinetic mechanisms for NOX2 complex-mediated electron transfer, superoxide generation, and regulations by pH and inhibitory drugs:

Upon assembly and activation (Fig. 1B), the NOX2 complex catalyzes the following electron transfer reaction to-wards O2 production [46] (see Appendix for further details and thermodynamics):

NADPH+2O2NADP++2O2+H+ (1)

The five well-described elementary electron transfer reactions resulting O2 generation are schematized in Fig. 1C, D (see Appendix for further details and thermodynamics). As rationalized in the Appendix, the NOX2 kinetic model based on the catalytic scheme of Fig. 1C, D would be complex containing five unknown parameters for estimation for the base model. In order to reduce the complexity of the model and the number of unknown parameters, we lumped some of the elementary electron transfer reactions, as schematized in Fig. 1E, F, resulting in three unknown parameters for the base model. The corresponding regulation mechanisms by pH and inhibitory drugs are schematized in Fig. 2(AC).

Fig. 2. Proposed kinetic mechanisms for the regulations of NOX2 complex-mediated electron flow and superoxide production by pH and inhibitory drugs.

Fig. 2

(A) Schematics shows our hypothesized pH regulation mechanism of the electron transfer reactions. We assume rapid equilibrium binding of protons (H+) at the states En.X (inactive) to form the states En.X.H (active and promotes electron transfer reactions) and the states En.X.H2 (inactive), which collectively termed here as (En.X)T, n = 1, 2 and 3. This H+ binding modifies the rate constants knf and knr, by the pH-dependent factors fnf and fnr, which enables us to simulate the bimodal effect of pH on the NOX2 complex activity, as observed experimentally. (B) Schematics shows the NOX2 mode of action of the drug GSK2795039 (GSK). The competitive inhibitor GSK binds to the NOX2 state (E1.X)T (i.e. competes for the NADPH binding site) in a rapid equilibrium fashion, which modifies the rate constants k1f and k3r by the GSK-dependent factors f1f and f3r; KGSK is the binding constant for GSK. (C) Schematics shows the NOX2 mode of action of the drug DPI (diphenyleneiodonium). The uncompetitive inhibitor DPI binds to the NOX2 state (E2.X)T (i.e. binds after NADPH is bound) in a rapid equilibrium fashion, which modifies the rate constants k1r and k2f by the DPI-dependent factors f1r and f2f; KDPI is the binding constant for DPI.

In Fig. 1, the NOX2 membrane-bound subunits gp91phox and p22phox are together represented as E, which encompasses FlavoCytB (flavocy-tochrome b245 or b558), but is in inactive form (Fig. 1A). Upon binding with the assembled cytosolic subunit complex p47phox.p40phox.p67phox.Rac-GTP (represented as X), the NOX2 enzyme gets assembled and activated to form the enzyme complex E.X (Fig. 1B), which mediates the transfer of electrons from substrate NADPH to molecular O2 in multiple steps through the action of FlavoCytB within the membrane-bound subunits gp91phox and p22phox (Fig. 1CF). In the first step, two electrons are transferred from NADPH to FAD to reduce it to FADH2. Subsequently, FADH2 provides two electrons to the heme sites of two CytB in two sequential steps via the intermediate radical FADH to reduce two CytB. In the final step, two electrons are transferred from the heme sites of two reduced CytB to two O2 molecules in two sequential steps to produce two O2 molecules. The process is referred to as the sequential one-electron O2 reduction mechanism by FlavoCytB, as proposed recently by Nisimoto et al. [38].

As detailed in the Appendix, the schematics of Fig. 1E, F are simplifications of the schematics of Fig. 1C, D as a result of the lumping of the two elementary electron transfer reactions from FADH2 to 2CytBox via FADH and from 2CytBred to 2O2, without affecting the overall thermodynamics of the electron transfer reactions. In Fig. 1D, F, the different states of the NOX2 complex E.X are represented by En.X, depending on the redox status of FlavoCytB within gp91phox and p22phox. Specifically, E1, E2, E3, E4, and E5 in Fig. 1D represent FAD.2CytBox, FADH2.2CytBox, FADH.CytBox.CytBred, FAD.2CytBred, and FAD.CytBred.CytBox, respectively, while E1, E2, and E3 in Fig. 1F represent FAD.2CytBox, FADH2.2CytBox, and FAD.2CytBred, respectively. Thus, for the simplified three-state model (Fig. 1F), FADH.CytBox.CytBred and FAD.CytBred.CytBox states in the complex five-state model (Fig. 1D) are eliminated. The assumption is that the electron transfer rate from FADH to CytBox is fast compared to that from FADH2 to CytBox, so that the contributions of the states FADH.CytBox.CytBred and FAD.CytBred.CytBox are accounted for in the forward or reverse rate constants of the relevant electron transfer reactions in Fig. 1F.

As schematized in Fig. 2A, the NOX2 kinetic model includes thermodynamically-balanced, pH-regulated electron transfer reactions from NADPH to O2 in producing O2, which actually occurs at multiple redox centers within the NOX2 complex (i.e. the flavin of adenine dinucleotide (FAD) and terminal oxidase cytochrome b245 (CytB)). It should be noted here that the resting (inactivated) NOX2 enzyme does not transfer electrons from NADPH to the flavin center at a significant rate, as evidenced by the absence of FAD or CytB reduction [29,33,35]. The pH regulation scheme (Fig. 2A) was developed to reproduce the experimentally-observed bimodal behavior of the NOX2 enzyme activity as a function of pH [3941]. Specifically, we assume rapid equilibrium binding of protons (H+) to En.X (inactive states) to form En.X.H (active states and promote electron transfer reactions) and En.X.H2 (inactive states), which are collectively termed as the total En.X ((En.X)T, n = 1, …, 3). Finally, the NOX2 kinetic model also incorporates differential modes of actions of the drugs GSK2795039 (denoted here as GSK) and diphenyleneiodonium (DPI) on the NOX2 enzyme activity, as schematized in Fig. 2B, C. Specifically, GSK acts as a competitive inhibitor (competes for the NADPH binding site), while DPI acts as an uncompetitive inhibitor (binds after NADPH binding), thereby regulating the NOX2 enzyme activity [45].

Flux expression for NOX2 complex-mediated superoxide production:

At steady state, the rate of change in the enzyme concentration at each state of the assembled and activated NOX2 complex is zero (Fig. 2AC), and hence the turnover of the enzyme (i.e. reaction flux per unit enzyme concentration; 1/time) can be calculated using any of the rate limiting elementary reaction steps (see Appendix for detailed derivation). This can also be easily computed using our KAPattern tool in MATLAB for generating rate equations for enzymatic reactions [47]. This results in the following expressions for the fractional enzyme states (F1, F2, F3) for the NOX2 complex:

F1=C(E1.X)TC(E.X)T=f1rk1rf2rk2rCNo+f1rk1rf3fk3fCNoCO2+f2fk2ff3fk3fCO2Den (2a)
F2=C(E2.X)TC(E.X)T=f1fk1ff3fk3fCNrCO2+f1fk1ff2rk2rCNr+f2rk2rf3rk3rCO2Den (2b)
F3=C(E3.X)TC(E.X)T=f1fk1ff2fk2fCNr+f2fk2ff3rk3rCO2+f1rk1rf3rk3rCNoCO2Den (2c)

where the common denominator (Den) is defined as:

Den=f1fk1ff2fk2fCNr+f1fk1ff2rk2rCNr+f1fk1ff3fk3fCNrCO2+f1rk1rf2rk2rCNo+f1rk1rf3fk3fCNoCO2+f2fk2ff3fk3fCO2+f2rk2rf3rk3rCO2+f2fk2ff3rk3rCO2+f1rk1rf3rk3rCNoCO2 (2d)

In the above equations, C(En.X)T denotes the total En.X concentrations (individual NOX2 complex states) and C(E.X)T denotes the total E.X concentration (NOX2 complex). The NOX2 reaction flux (concentration/time) can then be derived as the turnover rate of the activated NOX2 complex times the total activated NOX2 complex concentration, which is expressed as:

JNOX2=C(E.X)T(f1fk1ff2fk2ff3fk3fCNrCO2f1rk1rf2rk2rf3rk3rCNoCO2)(f1fk1ff2fk2fCNr+f1fk1ff2rk2rCNr+f1fk1ff3fk3fCNrCO2+f1rk1rf2rk2rCNo+f1rk1rf3fk3fCNoCO2+f2fk2ff3fk3fCO2+f2rk2rf3rk3rCO2+f2fk2ff3rk3rCO2+f1rk1rf3rk3rCNoCO2) (3)

where the total NOX2 complex concentration C(E.X)T is a measure of the FlavoCytB concentration CFlavoCytB; CNr and CNo denote NADPH and NADP+ concentrations, respectively; CO2 and CO2 denote oxygen and superoxide concentrations, respectively; knf and knr are the rate constants for the nth forward and reverse reactions, respectively; and fnf and fnr are the scaling factors for the rate constants knf and knr due to the binding of NOX2 complex with protons and drugs, as defined in Equation (7) below. The NOX2 reaction flux expression (Equation (3)) is based on sequential one-electron O2 reduction mechanism by FlavoCytB, similar to that proposal by Nisimoto et al. [38], and hence the flux expression involves CO2 and CO2, rather than CO22 and CO22 (see Appendix).

Simplified NOX2 flux expression in the absence of reaction products:

Most of the experimental data that are used in the present study for model development involve NOX2 complex-mediated NADPH and O2 dependent O2 production in the absence of the reaction products NADP+ and O2. Under such conditions, the NOX2 reaction flux expression can be simplified to:

JNOX2=C(C.X)Tf1fk1ff2fk2ff3fk3fCNrCO2(f1fk1ff2fk2fCNr+f1fk1ff2rk2rCNr+f2fk2ff3fk3fCO2+f1fk1ff3fk3fCNrCO2) (4)

which can be rewritten in the form of familiar Michaelis-Menten equation:

JNOX2=VmaxCNrCO2(KO2CNr+KNrCO2+CNrCO2) (5)

where V′max is the apparent maximal velocity, and K′Nr and K′O2 are the apparent Km (Michaelis-Menten constants) of NADPH and O2, respectively. These kinetic parameters are expressed in terms of the reaction rate constants as:

Vmax=C(C.X)Tf2fk2f (6a)
KNr=f2fk2ff1fk1f (6b)
KO2=(f2fk2f+f2rk2r)f3fk3f (6c)

NOX2 complex kinetic regulations by pH and inhibitory drugs:

The fractions of total active NOX2 complex in different states facilitating the electron transfer reactions due to regulations by protons and drug molecules can be expressed as:

fH=(CHKH1)/(1+CHKH1+CH2KH1KH2) (7a)
f1f=f3r=fH/(1+CGSKKGSK) (7b)
f2f=f1r=fH/(1+CDPIKDPI) (7c)
f3f=f2r=fH (7d)

where CGSK, CDPI and CH denote the concentrations of GSK, DPI and protons bound to NOX2 complex (different states), respectively; KGSK, KDPI, KH1 and KH2 represent the respective binding constants for GSK, DPI and protons, as represented in Fig. 2AC.

Thermodynamic and kinetic constraints for NOX2 model parameters:

The standard midpoint redox potentials Em,j0 0 for different redox pairs involved in the three lumped elementary electron transfer reactions (Fig. 1F) at pH=7 [29], and the equilibrium constants Keq,i of the electron transfer reactions derived from these midpoint redox potentials (see Appendix), which are also further corrected for variable pH, are given by:

Em,Nr/No0=320mV (8a)
Em,Fr/Fo0=280mV (8b)
Em,Cr/Co0=245mV (8c)
Em,O2/O20=160mV (8d)
Keq,1=Keq,E1E2=(CHKH)exp(2F(Em,Fr/Fo0Em,Nr/No0)RT) (8e)
Keq,2=Keq,E2E3=(KHCH)2exp(2F(Em,Cr/Co0Em,Fr/Fo0)RT) (8f)
Keq,3=Keq,E3E1=exp(2F(Em,O2/O20Em,Cr/Co0)RT) (8g)
Keq,NOX2=i=13Keq,i=(KHCH)exp(2F(Em,O2/O20Em,Nr/No0)RT) (8h)

where Nr and No denote NADPH (reduced) and NADP+ (oxidized), respectively; Fr and Fo denote FADH2 (reduced) and FAD (oxidized), respectively; Cr and Co denote the reduced and oxidized forms of FlavoCytB, respectively; KH = 1 × 10−7M is the reference proton concentration; F, R and T represent the Faradays constant (96485 J/mol/V), ideal gas constant (8.314 J/mol/K) and temperature (298.15 K), respectively. Ideally, the equilibrium constants in Equation 8(eh) should also be corrected for temperature variations using the Van’t-Hoff equation [48], which are expected to be small. However, these small corrections could not be accounted for, as the enthalpy changes associated with the individual electron transfer reactions are not known.

The pH-dependency of the rate constants are appropriately determined based on thermodynamics and whether H+ is a reactant or a product of the associated electron transfer reaction. In reaction 1 of Fig. 1F, H+ is a reactant, and in reaction 2, 2H+ are products. The model accounts for these H+ effects through modifications of the forward or reverse rate constants, which are constrained by equilibrium constants, which are functions of pH (Equation 8(eh)). As shown below, k1r and k2r are dependent on pH, in addition to k1f. The ratio k1f/k1r=Keq,1 is proportional to (CH/KH), while the ratio k2f/k2r=Keq,2 is proportional to (KH/CH)2. In addition, all the rate constants are further modified by pH-dependent function fH, defined in Equation (7a), as determined by the kinetic schematics of Fig. 2A. The pH-dependency of the rate constant k1f as (CH/KH)0.5 rather than (CH/KH) is motivated by the pH-dependent kinetic data (NOX2 Km of NADPH as a function of pH [39]); k1r is set appropriately to satisfy the thermodynamic constraint below. The rate constants are constrained by the following kinetic and thermodynamic constraints:

k1f=k2fKNr(CHKH)0.5 (9a)
k1r=k1fKeq,E1E2 (9b)
k2r=k2fKeq,E2E3 (9c)
k3f=(k2f+k2r)KO2 (9d)
k3r=k3fKeq,E3E1 (9e)

where KNr and KO2 are the intrinsic Km of NADPH and O2 for NOX2 at pH=7, respectively. The apparent Vmax and apparent Km of NADPH and O2 for NOX2 at variable pH and drug concentrations, defined in Equation (6), can then be expressed as:

Vmax=C(E.X)Tk2ffH/(1+CDPIKDPI) (10a)
KNr=KNr(KHCH)0.5(1+CGSKKGSK)/(1+CDPIKDPI) (10b)
KO2=KO2(k2f/(1+CDPIKDPI)+k2r)/(k2f+k2r) (10c)

It is clear from Equation (10) that the drug DPI decreases both the apparent Vmax and apparent Km of NADPH and O2 for NOX2, while the drug GSK only increases the apparent Km of NADPH for NOX2. In the absence of drugs, the apparent Vmax and apparent Km of NADPH and O2 for NOX2 at variable pH can be simplified as:

Vmax=C(E.X)Tk2ffH (11a)
KNr=KNr(KHCH)0.5 (11b)
KO2=KO2 (11c)

Temperature dependency of NOX2 complex kinetics:

The effect of temperature on the kinetics of NOX2 complex was modeled by multiplying the elementary electron transfer reaction rate constants by the following temperature correction coefficient based on the Arrhenius principle [48]:

Tfact=Q10(T2T1)10 (12)

where T1 = 25 °C at which the reaction rate constants are defined, T2 is the temperature (°C) at which the reaction rate constants are calculated, Q10 = 2.5 is the temperature correction factor for the increase of reaction rate for every 10 °C temperature change. This approach of accounting for the temperature effect was motivated by the results of a study by Morgan et al. [42] on the temperature dependency of the NOX2 enzyme activity in stimulated human eosinophils while changing the temperature rapidly during the peak of the respiratory burst. We corrected the kinetic parameter k2f for the temperature change effect as k2f = Tfact × 10, and this way all the rate constants have temperature effects as per Equation (9).

Numerical estimation of NOX2 model parameters:

The developed NOX2 kinetic model contains several unknown parameters (Table 1). We used the values of the standard midpoint redox potentials involved in the elementary electron transfer reactions and the corresponding equilibrium constants (Equation (8)) to constraint the reverse reaction rate constants k1r, k2r, and k3r (Equation (9)). In addition, we fixed some kinetic parameters at their stipulated values (i.e. k2f = 10 sec−1 which is a measure of the maximal reaction velocity Vmax and cannot be distinguished from CFlavoCytB; KH = 1 × 10−7 as the reference proton concentration). To estimate the remaining unknown kinetic parameters, we used a modular multi-step optimization technique (e.g. sequential use of the available experimental data [3742,45] related to electrons transfer, pH effect, drugs action, and temperature effect) of the combined least-square objective function defined below to fit the model solution to the experimental data:

minϕE(ϕ),E(ϕ)=k=1Nexp1Ndata,k(j=1Ndata,k(Jj,kdataJj,kmodel(ϕ)max(Jj,kdata))2) (13)

where Nexp is the number of experiments and Ndata,k is the number of data points in the kth experiment; Jj,kdata are the experimental data and Jj,kmodel(ϕ) are the corresponding model solutions that depend on the values of the model parameters ϕ. The accuracy and robustness of the model fitting to the experimental data were assessed based on the value of the mean square residual error E (ϕ) and the sensitivities of E (ϕ) or J (ϕ) to perturbations in the optimal parameter estimates. All the model parameters and their values (fixed or estimated) are summarized in Table 1.

Table 1.

Estimated parameter values for NOX2 kinetic model (NOX2 complex-mediated electron transfer and O2 production).

Parameter Definition/Comment Value Units Reference
C(E.X)T Total active NOX2 complex concentration (known from specific experimental conditions) Vary depending on the experiment/assay μM [3740,45]
KNr Km of NADPH for NOX2 complex at pH = 7 (estimated based on NADPH-dependent data on O2 production) Vary depending on the experiment/assay
40–300 (Cell-free assays)
5–15 (Cell-based assays)
μM [37,39,40]
KO2 Km of O2 for NOX2 complex at pH = 7 (estimated based on O2 dependent data on O2 46 (Cell-free assays)
60 (Cell-based assays)
μM [38]
CH Proton (H+) concentration (vary depending on experiment/assay) 10−pH M
KH Reference 10−pH at which Keq’s or Em0's of electron transfer redox reactions were obtained (fixed) 10–7.0 M [3740,45]
KH1 First H+ binding constant (estimated based on pH dependent data on O2 production) 10–7.52 M [40]
KH2 Second H+ binding constant (estimated based on pH dependent data on O2 production) 10–6.69 M [40]
k1f Forward rate constants from state (E1.X) to state (E2.X); dependent on CH/KH (calculated) (k2f /KNr )(CH /KH )0.5 1/(μM.s) Eq. 9a
k1r Reverse rate constants from state (E2.X) to state E1.X; proportional to (CH/KH)−0.5 (calculated) k1f/Keq,E1E2 1/(μM.s) Eq. 9b
k2f Forward rate constant from state (E2.X) to state E3.X (fixed) 10.0 1/s [3740,45]
k2r Reverse rate constant from state E3.X to state (E2.X); proportional to (CH/KH)2 (calculated) k2f/Keq,E2E3 1/s Eq. 9c
k3f Forward rate constant from state E3.X to state E1.X; dependent on CH/KH via k2r (calculated) (k2f+k2r)/KO2 1/(μM.s) Eq. 9d
k3r Reverse rate constant from state E1.X to state E3.X; dependent on CH/KH via k2r (calculated) k3f/Keq,E3E1 1/(μM.s) Eq. 9e
KDPI DPI binding constant for NOX2 complex (estimated) 0.78 μM [45]
KGSK GSK binding constant for NOX2 complex (estimated) 0.55 μM [45]

3. Results

In this section, we present the parameterization and simulation results of the integrated kinetic model of NOX2 complex-mediated electron transfer and O2 production, and their regulations by pH, temperature, and inhibitory drugs. As described below, we estimated the values of the unknown model parameters (Table 1) by fitting the model solution to a diverse set of published experimental data (Figs. 3 and 4) [3742,45]. The parameterized and validated/corroborated model was then used to predict emergent properties of the NOX2 enzyme system (i.e. NADPH and O2-dependent bimodal effects of pH on NOX2 enzyme activity, fractional enzyme states, and associated kinetic parameters as well as alkaline pH-dependent NADP+ inhibition of NADPH-dependent O2 production and the effect of temperature on NOX2 enzyme activity) (Figs. 58).

Fig. 3. Model fittings to available experimental data on NADPH and O2 dependent superoxide generations via the assembled and activated NOX2 complex, and their regulations by pH.

Fig. 3

(A–C) Model simulations (lines) are compared to experimental data (symbols) on the effects of varying concentrations of NADPH on relative NOX2 fluxes (JNOX2; the rates of O2 production via the NOX2 complex normalized with respect to the maximal rate), studied either in permeablized polymorphonuclear neutrophils (PMNs) (cell-based system) [37] (A) or in isolated plasma membrane (PM) preparations from PMNs (cell-free system) [37,40] (B,C) under different NOX2 stimulation conditions. In Bauldry et al.’s experiments [37], 1 ml cuvette buffer contained multiples of 1 × 106 cells or its equivalent PM protein. Since 1 × 106 cells is equivalent to approximately 1.5 pmol of FlavoCytB, its concentration in the cuvette was assumed to be multiples of 1.5 pmol/ml = 1.5 nM, which was used as the total activated NOX2 complex concentration for the model simulations. Similar assumptions were made for the analysis of Suzuki et al.’s experimental data [40]. (D) Model simulations (lines) are compared to experimental data (symbols) on the effects of varying concentrations of O2 on relative NOX2 fluxes (JNOX2; the rates of O2 production via the NOX2 complex normalized with respect to the maximal rate), studied either in intact PMNs (cell-based system) or in isolated PM preparations from PMNs (cell-free system) under different NOX2 stimulation conditions [38]. In Nisimoto et al.’s experiments [38], since 0.8 ml cuvette buffer contained about 4.2 × 105 cells or its equivalent PM protein and NOX2 was activated differently, FlavoCytB concentration in the cuvette was assumed to be multiples of (0.42/ 0.8) × 1.5 pmol/ml ≈ 0.8 nM, which was used as the total activated NOX2 complex concentration for the model simulations. Percentage O2 in the cuvette was first converted to partial pressure of O2 (PO2), and then to dissolved O2 using [O2] = αO2 × PO2, where αO2 = 1.65 μM/mmHg and PO2 = (150/21) × (%O2) mmHg. (E) Model simulation (line) is compared to experimental data (symbols) [40] demonstrating the bimodal behavior of the assembled and activated NOX2 complex activity, measured as the rate of O2 production (normalized with respect to the maximal rate) in isolated PM preparations from PMNs upon pH changes ranging from acidic to neutral to alkaline values (pH = 5–10), with high or saturating substrate NADPH and O2 concentrations (250 μM each) and FlavoCytB = 1.5 nM. (F) Model simulation (line) on the apparent Km of NADPH for the NOX2 complex is compared to experimental data (symbols) [39] for pH values ranging from 5 to 8. The model simulations are made with fixed NOX2 complex concentration (assembled and activated equivalent to 1.5 nM FlavoCytB), with unlimited O2 concentration (250 μM) in the cuvette. In all plots (A–F), JNOX2 is appropriately normalized to have unit NOX2 complex maximal activity for the purpose of model comparisons to the experimental data. The estimated apparent Km values and other experimental details are also indicated in the respective figure insets.

Fig. 4. Model fittings to available experimental data on NADPH dependent superoxide generations via the assembled and activated NOX2 complex, and their regulations by the inhibitory drugs DPI and GSK.

Fig. 4

(A) Model simulations (lines) are compared to experimental data (symbols) on the kinetic profiles of the drug DPI as an uncompetitive inhibitor of the NOX2 complex [45]. In Hirano et al.’s experiments [45], the effects of various concentrations of the drug DPI on the NADPH-dependent O2 production (JNOX2; normalized with respect to the maximal rate) were studied in the isolated cell membrane preparations (cell-free system; recombinant NOX2) at high O2 and pH = 7.4. (B) Model simulations (lines) are compared to experimental data (symbols) on the kinetic profiles of the drug GSK as a competitive inhibitor of the NOX2 complex [45], with the experimental details the same as that for Panel A. (C, D) Model predictions of the inhibitory profiles of the drugs DPI and GSK (relative NOX2 flux as % of control vs. log10 of drug concentrations) for different concentrations of NADPH in the experimental solution. (E, F) Model simulations of the time courses of NADPH depletions and O2 productions (simulated as decreasing concentrations of NADPH and increasing concentrations of O2) in response to different increasing concentrations of the uncompetitive inhibitor drug DPI and competitive inhibitor drug GSK. The model parameters are the same as that for panels (A, C) and (B, D). (G–J) Model simulations of NOX2 apparent Km of NADPH and Vmax with respect to NADPH as functions of the drug DPI (G, H) and GSK (I, J) concentrations at high O2 and pH = 7.4.

Fig. 5. Model predictions demonstrating multi-dimensional profiling of the NADPH, O2 and pH dependent superoxide generations and the corresponding NOX2 complex states.

Fig. 5

(A, B) Model simulations depict bimodal behavior of the NOX2 complex activity, simulated as the rate of O2 production (NOX2 flux; nM/s) upon pH changes ranging from acidic to neutral to alkaline values (pH = 5–9) with either the substrate NADPH or O2 varying ranging from limited to saturated levels (5–250 μM) in the absence of the products NADP+ and O2. When NADPH is varied, O2 is maintained at saturated level (250 μM) and vice-versa. (C, D) Model simulations depict transitions of the assembled and activated NOX2 complex states, expressed as fractions of the total NOX2 complex (F1, F2, F3), as functions of pH changes ranging from acidic to neutral to alkaline values (pH = 5–9) with the substrate NADPH ranging from limited to saturated levels (5–250 μM) at high saturated level of O2 (250 μM) in the absence of the products NADP+ and O2. (E, F) Model simulations depict transitions of the assembled and activated NOX2 complex states (F1, F2, F3) as functions of pH changes ranging from acidic to neutral to alkaline values (pH = 5–9) with the substrate O2 ranging from limited to saturated levels (5–250 μM) at high saturated level of NADPH (250 μM) in the absence of the products NADP+ and O2. The model simulations are made with fixed NOX2 complex concentrations (assembled and activated equivalent to 1.5 nM FlavoCytB) with parameters corresponding to the cell-free system, as reported in Table 1.

Fig. 8. Model predictions demonstrating the effects of temperature on the kinetic parameters of the NOX2 complex governing electron transfer and superoxide production.

Fig. 8

(A, B) Model simulations depict variations of the apparent Km of NADPH for the NOX2 complex as functions of varying O2 concentration (5–250 μM) and pH (5–9) for a range of temperature (T = 18–43 °C) in the absence of the reaction products NADP+ and O2. (C, D) Model simulations depict variations of the apparent Km of O2 for the NOX2 complex as functions of varying NADPH concentration (5–250 μM) and pH (5–9) for a range of temperature (T = 18–43 °C) in the absence of the reaction products NADP+ and O2. (E, F) Model simulations depict the effects of varying temperature (T = 18–43 °C) on the apparent Vmax with respect to O2 or NADPH for the NOX2 complex as functions of varying NADPH or O2 concentrations (5–250 μM) at pH = 7.4 (E) and varying pH (5–9) at high saturating levels of NADPH and O2 (250 μM) (F) in the absence of the reaction products NADP+ and O2. These apparent kinetic parameters are calculated with fixed NOX2 complex concentration (assembled and activated equivalent to 1.5 nM FlavoCytB) with the intrinsic kinetic parameters values, as reported in Table 1, corresponding to the cell-free system.

Kinetics of NOX2 complex-mediated electron transfer, superoxide production, and regulation by pH:

Fig. 3 shows that the NOX2 kinetic model (Equation (4) or (5)) fits well the published experimental data [3740] on NADPH, O2 and pH-dependent O2 generation via the assembled and activated NOX2 complex. These fits provided the values of the rate constants for the elementary electron transfer reactions in Fig. 2A, subject to thermodynamic and kinetic constraints with some parameters fixed appropriately (Table 1). Bauldry et al. [37] developed a cell-based assay system to activate NOX2 in situ in permeabilized neutrophils at pH = 7.4 to determine the substrate NADPH affinity for NOX2, which was more physiologically appropriate than that obtained using data from isolated membrane preparations (usually several fold greater). As shown in Fig. 3A, the NOX2 kinetic model was able to fit the above data very well for three different stimulation conditions, resulting in higher Vmax values (a measure of enzyme activity) with higher NOX2 stimulations (not apparent in the figure due to normalization of the rates) and appropriate Km values (3.9–7.3 μM at pH 7.4; a measure of the affinity of the enzyme for the substrate). Furthermore, Fig. 3B, C shows that the same model fits well published experimental data on NADPH-dependent O2 production via the activated NOX2 complex observed in isolated membrane preparations [37,40] at different pH with approximately 10–50 fold higher Km values for NADPH. Interestingly, under isolated membrane preparation conditions, the model was able to capture the increased velocities of NOX2 with higher stimulations, but with the same apparent Km value (Fig. 3C). In addition, we observed differences in the Km values for NADPH from one experiment (Fig. 3B) to the other (Fig. 3C) in isolated membrane preparations, suggesting differences in the conditions used for experiments (e.g. pH). Similarly, as shown in Fig. 3D, the developed NOX2 kinetic model also fitted very well the activated NOX2 enzyme kinetics in permeablized neutrophils and isolated membrane preparations with varying O2 concentrations when performed under saturated NADPH at pH = 7.4 [38]. The model fits revealed small variations in the apparent Km values of O2 for NOX2 enzyme prepared under different experimental conditions (see the figure caption for further descriptions of the experiments).

The published kinetic data from Suzuki et al. [40] in isolated membrane preparations was used to characterize the bimodal behavior of the NOX2 enzyme activity under varying pH conditions and identify the pH regulation mechanism and dependence of the rate constants on pH (Fig. 3E, F). The NOX2 kinetic model with postulated pH regulation mechanism (Fig. 2A; Equation (4) or (5)) was able to capture the bimodal behavior of the NOX2 enzyme activity with pH variations, with the optimal activity near pH 7 (Fig. 3E), as observed by Suzuki et al. [40]. The authors used stimulated granule rich fragments (GRF) from human neutrophils and found that the GRF activity was dependent on pH and the activation status (resting vs. stimulated) of the neutrophils. The same model was also able to fit well the effects of pH variations on NOX2 Km of NADPH (Fig. 3F), indicating that the affinity of NADPH for NOX2 changes significantly with pH, as observed experimentally by Babior et al. in isolated membrane preparations [39].

Regulation of NOX2 enzyme kinetics by inhibitory drugs:

The NOX2 kinetic model (Equation (4) or (5)) was then used to simulate the recently published experimental data [45] regarding the action of an uncompetitive inhibitor, DPI, as well as a competitive inhibitor, GSK, by properly accounting for their bindings to the NOX2 enzyme states (Fig. 2B, C). The model simulations are shown in Fig. 4. Hirano et al. [45] investigated the modes of inhibitory actions of the drugs DPI and GSK in cell-free membrane-isolated activated NOX2 complex by determining the concentration-dependent effects on NADPH-dependent O2 production. With proper DPI and GSK binding to the NOX2 enzyme states (i.e. GSK binding to (E1.X)T and DPI binding to (E2.X)T; Fig. 2B, C), the NOX2 kinetic model was able to describe well these kinetic data (Fig. 4A, B). In addition, these model fits provided appropriate values for the binding constants of these two drugs and the NOX2 enzyme states characterizing the drug inhibitions.

Using the estimated values of the model parameters, we simulated the resultant effects of these drugs on the NOX2 enzyme kinetics, as shown in Fig. 4C, D. Here, the model predictions plotted with logarithmic changes in the drug concentrations against the fraction of NOX2 enzyme activity clearly show a sigmoidal inhibitory effects of both drugs. However, the inhibition with GSK was markedly more dependent on the availability of NADPH (Fig. 4D) than that with DPI for higher NADPH, specifically at concentration greater than 10 μM (Fig. 4D). Furthermore, the experimental results of Hirano et al. [45] showed that DPI decreased both the Km of NADPH and Vmax with respect to NADPH for NOX2, while GSK only increased the Km value without affecting the Vmax value. The model was able to simulate such behavior as shown in Fig. 4G, H for DPI and Fig. 4I, J for GSK (also see Equation (10)). Finally, Fig. 4C, F shows the model predictions of the concentration-response curves of the drugs DPI and GSK on NADPH depletion in comparison with O2 production in a time-dependent manner (for 0–60 min) in a typical isolated membrane preparation experiment. These temporal profiles clearly show that the increased concentration of these drugs slowed down the consumption rate of NADPH utilized to produce O2, which agrees well with the raw fluorescent measurements of Hirano et al. [45].

Model predictions of key emergent properties of the NOX2 enzyme system:

Enzyme activities are sensitive to pH changes and operate optimally at certain pH. Several studies in the literature demonstrated the bimodal effects of pH on the NOX2 enzyme activity [3941]. One such study is that of Suzuki et al. [40], discussed above (Fig. 3E), and in a similar study, the NOX2 complex-generated electron current across the plasma membrane was reported to be maximal at around pH = 7.2, and to decrease at higher and lower pH values [41]. We have incorporated the effects of altered pH into the NOX2 kinetic model via its effects on the thermodynamics of electron transfer (redox) reactions as well as via its effects on the NOX2 enzyme states (En.X)T (Fig. 2A). Using the model parameter values estimated from the experimental data in Fig. 3 for the cell-free assay system, the NOX2 kinetic model was used to simulate the enzyme functionality over a wide range of pH values (5–9) as well as NADPH or O2 concentrations (5–250 μM), as shown in Fig. 5. As illustrated in Fig. 5A, B, the model simulations are consistent with the bimodal behavior of the NOX2 enzyme activity with respect to pH alterations over a wide range of NADPH and O2 variations with the optimal activity near pH = 7.0 at saturated NADPH and O2 levels, as has been experimentally observed [40,41]. Unexpectedly, however, the optimal activity was shifted to lower pH values at lower concentrations of NADPH or O2, indicating modifications in the optimal NOX2 enzyme activity in acidic environments (Fig. 5A, B).

We further used the NOX2 kinetic model to determine the effects of substrate variations and pH upon the transitions of intermediate enzyme states for the activated NOX2 complex. As shown in Fig. 5C, D, the model predicted that the NOX2 complex prefers to stay in the enzyme state (E2.X)T at a lower pH (acidic), whereas at a higher pH (alkaline) it prefers to be in the enzyme state (E1.X)T, transitioning from (E1.X)T to (E2.X)T state at a level close to the neutral pH of 7 depending on NADPH concentrations. We observed that under these conditions of varying pH and NADPH, and at a constant saturated levels of O2, the fractional levels of (E3.X)T remain minimal (shown in green). In contrast, when we used a constant saturating concentration of NADPH (250 μM) and varied pH and O2 concentrations, as shown in Fig. 5E, F, the fractional levels of the state (E3.X)T were drastically increased with O2 under alkaline conditions. These predictions show that the NOX2 enzyme activity highly depends on the availability of its substrates together with the environment under which it operates which has not been previously quantitatively appreciated.

Alterations in the apparent Km's of substrates NADPH (Fig. 6A, B) and O2 (Fig. 6C, D) were simulated to study their relationships to these substrates and pH, and their effects on NOX2 enzyme kinetics. This was done over a range of pH (5–9) in the absence of the reaction products NADP+ and O2. These simulations revealed that the apparent Km of NADPH remained low (high NADPH affinity) and largely independent of O2 under acidic conditions. However, this apparent Km was highly dependent on O2 when pH increased to basic conditions leading to increased apparent Km and higher NADPH required for the NOX2 complex binding (Fig. 6A, B). In contrast, the simulations for the apparent Km of O2 indicate that it remains high under acidic conditions and slightly drops as NADPH level decreases from high to low values. Interestingly, as pH increased to basic conditions, the model predicted an overall reduction in the apparent Km of O2 for NOX2 with its dependency on NADPH (Fig. 6C, D). The simulations of the dependency of apparent Vmax for NOX2 at different pH with varying O2 or NADPH suggest optimal enzyme activity near pH 7.0 (simulations similar to Fig. 5A). Furthermore, the maximum velocities decreased as pH was altered in either direction from the neutral value indicating that pH exerts bimodal effects on both the NOX2 reaction flux and maximal reaction velocity (Vmax).

Fig. 6. Model predictions demonstrating multidimensional profiling of the kinetic parameters of the NOX2 complex governing electron transfer and superoxide production.

Fig. 6

(A–B) Model simulations depict variations of the apparent Km of NADPH for the NOX2 complex as functions of varying pH (5–9) and O2 concentration (5–250 μM) in the absence of the reaction products NADP+ and O2. (C–D) Model simulations depict variations of the apparent Km of O2 for the NOX2 complex as functions of varying pH (5–9) and NADPH concentration (5–250 μM) in the absence of the reaction products NADP+ and O2. These apparent kinetic parameters are calculated with fixed NOX2 complex concentration (assembled and activated equivalent to 1.5 nM FlavoCytB) with the intrinsic kinetic parameters values, as reported in Table 1, corresponding to the cell-free system.

The full NOX2 kinetic model (Equation (3)) was also used to investigate the effects of product accumulation and the associated product inhibition of NOX2 reaction flux (i.e. NADPH-dependent O2 production) at different pH levels. As shown in Fig. 7A, C, the model simulations show that the NOX2 reaction flux was minimally affected by the concentration of the product NADP+ at pH of 6.3, but was strongly affected by an increase to an alkaline pH of 7.7, with an intermediate effect at neutral pH of 7.0. The corresponding effects on the individual NOX2 complex states (NOX2 enzyme fractions) are illustrated in Fig. 7B, D showing increased sensitivities to NADPH for the NOX2 complex state transitions from (E1.X)T state to (E2.X)T state at pH = 6.3 vs. that at pH = 7.7 with increased NADP+ concentrations. On the contrary, the generated O2 as a product did not show any effect on the NOX2 reaction flux in terms of the product inhibition at any pH (results not shown). This model observation could be due to the fact that the respective reaction of O2 production (reaction from (E3.X)T to (E1.X)T) is always thermodynamically favored in the forward direction (i.e. high Keq,E3E1).

Fig. 7. Model predictions demonstrating the effects of NADP+ on NADPH-dependent superoxide generation and the corresponding NOX2 complex states at high O2 and different pH.

Fig. 7

(A, B) Model simulations depict pH-dependent product NADP+ inhibition of the substrate NADPH-dependent O2 production via the assembled and activated NOX2 complex at high saturating level of O2 (NADPH = 0:5:250 μM; NADP+ = 0:10:500 μM; pH = 6.3, 7.0, 7.7; and O2 = 250 μM). (C, D) Model simulations depict transitions of the assembled and activated NOX2 complex states, expressed as fractions of the total NOX2 complex (F1, F2, F3), as functions of the substrate NADPH and product NADP+ variations ranging from limited to excess levels (NADPH = 0:5:250 μM; NADP+ = 0:10:500 μM) at high saturating level of O2 (250 μM) for three different levels of pH (6.3, 7.0, 7.7). The model simulations were generated using a fixed NOX2 complex concentration (assembled and activated equivalent to 1.5 nM FlavoCytB) and with values of model parameters corresponding to the cell-free system, as reported in Table 1.

Regulation of NOX2 complex kinetics by temperature:

Temperature affects the NOX2 enzyme activity by slightly modulating the availability of the intermediate enzyme states through the alterations of the elementary electron transfer reactions equilibrium constants (Equation (8)). In addition, motivated by the data of a study by Morgan et al. [42], we further incorporated the effect of temperature on the NOX2 enzyme activity by multiplying the electron transfer reaction rate constants with the temperature correction coefficient (Equation (12)). The model predictions for the effects of temperature ranging 18–43 °C on the NOX2 kinetic parameters (apparent Km of NADPH and O2 and apparent Vmax) governing the NOX2 enzyme kinetics are shown in Fig. 8. These predictions suggest that the apparent Km of NADPH and O2 for NOX2 increase with increasing concentrations of O2 (Fig. 8A) and NADPH (Fig. 8C), respectively, but are independent of temperature. Similarly, the apparent Km of NADPH and O2 are significantly increased and decreased with increasing pH, respectively, but are independent of temperature (Fig. 8B, D). Unlike the apparent Km of NADPH and O2, the apparent Vmax for NOX2 appears to be highly dependent on temperature with velocities increasing with increasing temperature. However, they are significantly dependent on the availability of substrates, with highly decreased enzyme velocities at high temperatures and low substrate concentrations, which increased as the substrate concentrations increased at a constant pH value of 7.4 (Fig. 8E). In contrast, when pH was altered from acidic to alkaline values, the enzyme maximum velocities were drastically decreased with increased temperature with the observation of optimal rates at pH 7 (Fig. 8F).

4. Discussion

The major ROS (O2, H2O2)-producing systems within cells are cytochrome P450 oxidases, uncoupled nitric oxide synthase (NOS), mitochondrial electron transport chain (ETC), xanthine oxidase (XO), and NOX [57,4951]. Production of ROS by ETC is the byproduct of oxidative phosphorylation, and is largely known to be deleterious to the cell when produced in excess [57]. In contrast, NOXs are the only known enzyme family in the plasma membrane with the sole function of generating ROS upon assembly and activation, and have a major role in both physiological and pathophysiological conditions [13,915]. Although several kinetic models explaining the plausible mechanisms for production and detoxification of mitochondrial ROS are well elucidated [5255], to the best of our knowledge, no such kinetic model exists for NOXs (e.g. NOX2; the highly orchestrated, widely expressed, and most studied NOXs) that can explain the detailed kinetics of ROS production from this major non-mitochondrial source. Therefore, in this study, we developed and validated a thermodynamically-constrained mathematical model for the NOX2 complex-mediated electron transfer, ROS production, and regulations by pH, temperature, and inhibitory drugs.

An important aspect of our NOX2 kinetic model is its ability to fit well published kinetic data from multiple studies under a range of experimental conditions with appropriate model parameter values. In this regard, we note here that a model cannot fit data from various sources under different experimental conditions with suitable model parameter values if it does not incorporate appropriate kinetic mechanisms, no matter how many parameters it contains. Interestingly, the effect of pH on the NOX2 enzyme activity has a unique bimodal behavior, while the effects of the two distinct drugs GSK and DPI have very different behaviors. As shown, the model simulations are consistent with the experimentally-observed bimodal behavior of the NOX2 enzyme activity upon pH variations, and suggested alkaline pH-dependent product inhibition of the NOX2 reaction flux due to NADP+ accumulation at high pH. The model enables predictions that are important for quantitative understanding of the function and behavior of this key enzyme under physiological and pathophysiological conditions. The model also enables testing of the modes of actions and kinetics of different inhibitory drugs targeting the NOX2 enzyme system. Such analyses enable examination of the effects of targeting specific enzymatic sources of pathological ROS, which could overcome the limitations of pharmacological efforts aimed at scavenging of ROS, including the poor outcomes of antioxidant therapies in clinical studies [56,57].

Model captures NOX2 complex-mediated electron flow and superoxide production, which is regulated by pH:

The model exhibits remarkable ability to simulate the NADPH and O2-dependent O2 generation via the assembled and activated NOX2 complex as characterized from diverse experimental data in different assay systems at varying levels of pH (Fig. 3) [3742]. The model predicted higher Vmax values with higher NOX2 stimulations, but with slightly different Km values for NADPH for three different stimulation conditions in a study by Bauldry et al. [37] in a cell-based assay system in situ permeabilized neutrophils at pH = 7.4 (Fig. 3A). Indeed, the estimated average Km value of ∼6 μM at pH = 7.4 for NADPH from this in situ experimental system appears to be more appropriate than those found in isolated membrane preparations (Km value usually 10–15 fold greater; Fig. 3B), as suggested by Bauldry and colleagues [37]. Similarly, Suzuki et al. [40] used activated GRF to measure Km of > 0.25 mM NADPH and a Vmax of > 2.0 nmol NADP+/107 PMN-eq/minute. We were able to capture these observations with a similar apparent Km value for NADPH (0.32 mM) (Fig. 3C) indicating differences in the NOX2 enzyme kinetics in isolated membrane conditions despite the same mechanism of action. These differences in the Km values may be attributed to different affinities of the enzyme for the substrate NADPH due to differences in the molecular interactive forces between the enzyme and the substrate in these different experimental systems (cell-free vs cell-based assay systems).

The ability of NOX2 kinetic model to capture the NOX2 enzyme kinetics under different experimental conditions was tested in other ways, specifically by varying pH and reaction substrates. For example, unlike the experiments discussed above which were performed at a saturating O2 level, Nisimoto et al. [38] used stimulated plasma membrane-enriched fractions and cytosolic fractions from human neutrophils equilibrated with N2/O2 gas mixtures to explore the regulation of NOX2 enzyme activity during alterations of O2 availability. The experimentally observed Km of O2 was different for different experimental systems with the Km of PMN intact cells averaging 3.1% (22 mmHg) and that of cell-free system averaging 2.3% (16 mmHg). As illustrated in the results above, our model was able to very well capture the NOX2 enzyme kinetics under these conditions with varying concentrations of O2 at high NADPH and at pH 7.4, and provided appropriate estimates for the apparent Km of O2 (59 and 43 μM, respectively).

The NOX2 complex-mediated O2 production occurs through transfer of electrons from NADPH to O2 across the plasma membrane, where these electrons alkalinize the external solution while the protons acidify it. It has been found using human eosinophils that NOX2 can translocate electrons, generating a trans-membrane current of ∼30–40 pA at 37 °C at pH ≈ 7.2, and this current is reduced at both lower and higher pH [42]. Similarly, our model predicted bimodal behavior of NOX2 enzyme activity upon pH variations. This is of considerable relevance since there are inconsistencies in studies describing the effects of pH due to differences in the techniques used to quantify these effects and the use of cell-free vs. cell-based assay systems [58,59]. It has been reported that the NOX2 enzyme activity in cell-free assay systems is inhibited at both high and low pH, and functions optimally near a neutral pH of 7.0 [39,40]. This aspect (bimodal pH effect) was modeled by hypothesizing that protons bind to the (En.X)T state of the NOX2 complex such that (En.X.H)T is the active state and (En.X)T and (En.X.H2)T are inactive states (Fig. 2A). It was then determined if the model prediction with this pH regulation mechanism is consistent with the published kinetic data of Suzuki et al. [40] obtained from studies of isolated membranes activated by PMA with high NADPH and O2 (Fig. 3E). In addition to capturing the results of those studies, our model of the protonation mechanism for the NOX2 complex was able to capture the experimental data from Babior et al. [39] who evaluated the variation of the apparent Km of NADPH with respect to varying pH (Fig. 3F). The computed model results of these different conditions have nicely validated our proposed kinetic mechanism of the NOX2 complex-mediated O2 generation.

The NOX2 kinetic model as parameterized using the values corresponding to the cell-free NOX2 complex systems (shown in Table 1) was used to further understand the kinetic behavior of NOX2 under a wide range of experimental conditions. The results of these model simulations not only fit the observed bimodal behavior of the NOX2 enzyme activity with pH variations, but also predicted that the pH at which the optimal NOX2 complex activity occurs depends on the concentrations of its participating substrates (Fig. 5A, B). The model results were also consistent with the simulated fractional enzyme states which revealed that these intermediate enzyme states undergo transformation based on the availability of substrates NADPH and O2 with the majority of NOX2 residing in a particular state depending on the condition (Fig. 5CF). Finally, the NOX2 enzyme behavior was also characterized with respect to alterations in the apparent Km and Vmax as functions of varying pH (pH = 5–9), O2 (0–250 μM), and NADPH (5–250 μM) in the absence of the reaction products NADP+ and O2 (Fig. 6). Together, these simulations clearly demonstrated that pH plays an important role in the NOX2 enzyme kinetics and its affinity for substrates. We conclude that the optimal activity occurs at pH 7 under saturating substrate conditions, and can shifts to the acidic pH at low substrate concentrations.

Despite others having studied the NOX2 enzyme kinetics under different experimental conditions, none of these studies have elucidated the effects of the reaction products NADP+ and O2 on the enzyme activity. We therefore used the present model to evaluate the role of these reaction products under different substrate and medium conditions, in particular pH. As shown, the model predicted that the pH-dependent product (NADP+) inhibition of the NADPH-dependent O2 production is significantly higher at alkaline pH and saturated O2. The model predicts that optimal product inhibition occurs at a pH of 7.7, rather than 7.0 (Fig. 7A, C). The observed changes in the intermediate enzyme fractions with product therefore indicate that NADP+ alters the enzyme transition states to alter the enzyme activity under different simulation conditions (Fig. 7B, D).

Regulation of NOX2 enzyme kinetics by inhibitory drugs:

One of the important features of the NOX2 kinetic model is the incorporation of mechanisms of several inhibitors, specifically the drug molecules GSK and DPI (Fig. 4). Based on the experimental data of Hirano et al. [45] and the hypothesized inhibitory mechanisms for these drugs, the molecule GSK was shown to bind with the enzyme state (E1.X)T, whereas DPI was shown to bind to the enzyme state (E2.X)T in the NOX2 complex model to accurately describe the kinetic data. The model simulations showed decreased Km of NADPH and Vmax with respect to NADPH using DPI as an inhibitor, while GSK increased the Km value without affecting the Vmax value for NOX2 (Fig. 4GJ; Equation (10)). These results are consistent with a mechanism of uncompetitive inhibition with DPI and competitive inhibition with GSK. Interestingly, the model also predicts that the percentage inhibition of the NOX2 reaction flux via GSK is dependent upon NADPH while DPI is not (Fig. 4C, D). An earlier study by Doussiere et al. [60] suggests the heme of CytB is a potential target for the drug DPI for its inhibition of the NOX2 enzyme activity. However, this study did not include kinetic data of the type shown in Fig. 4A (i.e. NOX2 enzyme activity as a function of O2 level for different doses of DPI) which is needed to assess the possibility that the mechanism of action of DPI is by acting on the (E3.X)T state (FAD-2CytBred) of the NOX2 complex. Thus, our NOX2 kinetic model could not test this alternative action of DPI on the NOX2 enzyme kinetics.

The NOX inhibitors have established the roles of the NOX enzyme in both normal and pathological states [4345]. Unfortunately, human antioxidant supplementation trials have not been found to be therapeutically very effective [56,57], and other approaches that can better target specific sites or pathways of ROS production (e.g. NOX) are being sought [4345,6163]. As emphasized in those studies, an ideal NOX inhibitor would be active and functional in both cell-free and cell-based systems, does not have an intrinsic antioxidant activity, and is NOX isoform selective with minimal toxicity. The significance of the proposed NOX2 kinetic model lies in its capacity to mechanistically and quantitatively test the modes of actions and kinetics of different pharmacological drug inhibitors that could be used to target the NOX enzymes for therapeutic purposes.

Regulation of NOX2 enzyme kinetics by temperature:

From the aforementioned discussions, it is apparent that, mostly steady-state measurements at fixed temperatures examining the activated NOX2 complex-mediated O2 production (or the equivalent electron current generation across the plasma membrane) have been conducted. Only few studies have examined the effects of varying temperature on the activated NOX2 complex kinetics [42,6468]. The NOX2 enzyme is shown to be optimally active at sub-physiological temperatures, also showing a bimodal temperature dependence with loss of function at higher temperatures and significant inhibition at lower temperatures. The temperature at which the optimal activity occurs depends on the experimental system used in the study. The results of the most recent study by Morgan et al. [42] suggested that the steady-state NOX2 enzyme activity is strongly dependent on temperature between 20 and 30 °C, with the Arrhenius activation energy Ea of 25.1 kcal mol−1 (Q10 = 4.2). In contrast, the NOX2 enzyme activity was found weaker in response to a rapid temperature increase to 37 °C with Ea only 14 kcal mol−1 (Q10 = 2.2). The developed NOX2 kinetic model was used to simulate the effects of varying temperature (T = 18–43 °C) on the NOX2 apparent Km of NADPH and O2 and apparent Vmax with respect to NADPH and O2 with varying levels of NADPH and O2 (0–250 μM) and pH (5–9) with a fixed physiologically-relevant Q10 value of 2.5 (Fig. 8). These simulations found that the apparent Km of NADPH and O2 was independent of temperature at all pH, but the apparent Vmax increased with increasing temperature. The model cannot simulate the loss of NOX2 function scenario at high temperatures (above 37 °C), since it does not account for the denaturation process of the NOX2 enzyme at high temperatures.

Significance of the present study and future directions:

The main objective of the present study was to develop and parameterize a thermodynamically-constrained mechanistic kinetic model of the NOX2 enzyme system, as no such model exists in the literature. The NOX2 enzyme plays an important role in the cellular oxidative stress, one of the major factors in several human diseases. A well-characterized kinetic model of the NOX2 enzyme system will allow us to better understand the critical role of ROS in cellular oxidative stress, together with the other contributors to ROS production and detoxification. Therefore, this NOX2 kinetic model will be an important component (module) of an integrated computational model of the mitochondrial and cellular ROS handling systems (production and scavenging) that will be essential for furthering our understanding the critical roles of ROS (and NOX2) in mitochondria and cell physiology and pathophysiology.

The parameterized and validated/corroborated NOX2 kinetic model provides several important insights into the functioning (i.e. emergent properties) of the NOX2 enzyme system upon assembly and activation (Figs. 58), which can be experimentally tested. These include (i) the bimodal behavior of the NOX2 enzyme activity upon pH variations ranging from alkaline values to acidic values, (ii) NADPH-dependent shifting of the peak NOX2 enzyme activity upon pH variations, (iii) alkaline pH-dependent NADP+ product inhibition of the NOX2 enzyme activity, and (iv) temperature effects on the NOX2 enzyme activity in cell-free and cell-based assay systems. Experimental verification of these model predictions is important, and hence could be undertaken in future studies, which will provide further understanding of the NOX2 enzyme system.

The NOX2 enzyme activity is regulated in part by its multistep assembly and activation process, involving a group of plasma membrane proteins (i.e. gp91phox and p22phox) and several cytosolic proteins (i.e. p47phox, p40phox, p67phox and Rac-GTP) [2632] (Fig. 1A, B). However, in the current study, our primary objective was the development and validation of a mechanistic and quantitative understanding of the NOX2 complex-mediated electron transfer and O2 production, and the regulatory effects of pH, inhibitory drugs, and temperature under the assumption that the NOX2 enzyme is already assembled and activated. Incorporating the kinetics of NOX2 assembly and activation process into the proposed NOX2 electron flow and O2 production model will enhance the mechanistic and quantitative understanding of the NOX2 enzyme system, but is beyond the scope of the present study, and will be considered in future studies.

Summary:

The NOX2 kinetic model described in this manuscript is the first thermodynamically-constrained mechanistic mathematical model that incorporates diverse kinetic details on the NOX2 complex function. The model accounts for the effects of pH, inhibitory drugs, and temperature in addition to thermodynamics of electron transfer and subsequent O2 production. The key model hypotheses and simulation outcomes illustrated in this manuscript include (i) competitive and uncompetitive inhibitory effects of the drugs GSK and DPI on O2 production enabling detailed pharmacological profiling of these drugs on the NOX2 enzyme activity; (ii) altered NOX2 enzyme activity upon pH and temperature variations; and (iii) product inhibition of the NOX2 enzyme activity via NADP+ accumulation at higher pH levels without having such inhibitory effect with O2 accumulation. Among many potential applications, the model provides a mechanistic and quantitative framework to investigate the critical role of the NOX2 enzyme system in O2 production regulating diverse cellular mechanisms in health and disease, such as NOX2 and mitochondrial crosstalk/interaction in ROS induced ROS release in oxidative stress and diseases.

Acknowledgement

We thank the Reviewer for providing insightful comments of our manuscript, which has significantly improved the clarity of the manuscript. This work was supported by the National Institute of Health grants P01-GM066730, P01-HL116264, U01-HL122199, and R01-HL122662.

APPENDIX

This Appendix provides the mathematical derivations of the NOX2 reaction flux expression and the regulations by pH and inhibitory drugs in the presence of the reaction products. We rationalize the assumptions and mathematical derivations through the following steps.

Overall NOX2 Electron Transfer Reaction and Thermodynamics:

The overall two-electron transfer reaction for NADPH oxidation and O2 reduction towards NADP+ and superoxide (O2) productions, mediated by the assembled and activated NOX2 enzyme complex, is given by:

Nr+2O2krkfNo+2O2+H+ (A1)

where Nr and No denote NADPH (reduced) and NADP+ (oxidized), respectively; kf and kr (both are in the units of 1/M2.s) denote, respectively, the forward and reverse rate constants of the overall reaction. The equilibrium constant (Keq, unitless) for the overall reaction is defined by:

Keq=kfkr=(KHCH)exp(2F(Em,Nr/No0Em,O2/O20)RT) (A2)

where CH = 10−pH M is the proton concentration and KH = 10−7 M is the proton binding constant; Em,Nr/No0 and Em,O2/O20 denote, respectively, the standard midpoint redox potentials for half-reactions involving NADPH oxidation and O2 reduction at pH=7; F, R and T denote the Faradays constant (96485 J/mol/V), ideal gas constant (8.314 J/mol/K), and temperature (298.15 K), respectively.

Elementary NOX2 Electron Transfer Reactions and Thermodynamics:

The two-electron transfer from NADPH to 2O2 occurs via different redox centers of the NOX2 enzyme complex, as schematized in Fig. 1C. The elementary electron transfer reactions based on the kinetic scheme of Fig. 1D can be written as:

Reaction1:Nr+Fo.2Co+H+k1rk1fNo+Fr.2Co (A3)
Reaction2a:Fr.2Cok2ark2afFs.Co.Cr+H+ (A4)
Reaction2b:Fs.Co.Crk2brk2bfFo.2Cr+H+ (A5)
Reaction3a:Fo.2Cr+O2k3ark3afFo.Cr.Co+O2 (A6)
Reaction3b:Fo.Co.Cr+O2k3brk3bfFo.2Co+O2 (A7)

where Fr, Fs and Fo denote FADH2 (reduced), FADH· (semiradical) and FAD (oxidized), respectively; Cr and Co denote the reduced and oxidized forms of CytB, respectively; F.2C denotes the FlavoCytB (flavocytochrome b); knf and knr denote, respectively, the forward and reverse rate constants for the individual elementary electron transfer reactions; k1f and k1r are in the units of 1/M.s, k2af, k2bf, k2ar and k2br are in the units of 1/s, and k3af, k3bf, k3ar and k3br are in the units of 1/M.s. The equilibrium constants (unitless) for the five elementary electron transfer reactions are defined by:

Keq,1=k1fk1r=(CHKH)exp(2F(Em,Nr/No0Em,Fr/Fo0)RT) (A8)
Keq,2a=k2afk2ar=(KHCH)exp(F(Em,Fr/Fs0Em,Cr/Co0)RT) (A9)
Keq,2b=k2bfk2br=(KHCH)exp(F(Em,Fs/Fo0Em,Cr/Co0)RT) (A10)
Keq,3a=k3afk3ar=exp(F(Em,Cr/Co0Em,O2/O20)RT) (A11)
Keq,3b=k3bfk3br=exp(F(Em,Cr/Co0Em,O2/O20)RT) (A12)

where Em,red/ox0’s denote the standard midpoint redox potentials for half-reactions involving different reduced-oxidized pairs at pH=7. Note that Em,Fr/Fo0=(Em,Fr/Fs0+Em,Fs/Fo0)/2. It can be easily shown from Equation (A2) and (A8-A12) that:

Keq=kfkr=k1f.k2af.k2bf.k3af.k3bfk1r.k2ar.k2br.k3ar.k3br=Keq,1.Keq,2a.Keq,2b.Keq,3a.Keq,3b (A13)

Lumped NOX2 Electron Transfer Reactions and Thermodynamics:

Clearly, the kinetic model of the NOX2 enzyme complex based on the five elementary electron transfer reactions (Equation (A3A7)) would contain 10 kinetic parameters (k1f, k1r, k2af, k2ar, k2bf, k2br, k3af, k3ar, k3bf, k3br), subject to 5 thermodynamic constraints (Equation (A8A12)), resulting in 5 unknown kinetic parameters for estimation. In order to reduce the complexity of the model and the number of unknown kinetic parameters for estimation, we can lump reactions 2a and 2b and reactions 3a and 3b, resulting in the following two lumped two-electron transfer reactions:

Reaction2:Fr.2Cok2rk2fFo.2Cr+2H+ (A14)
Reaction3:Fo.2Cr+2O2k3rk3fFo.2Co+2O2 (A15)

The forward and reverse rate constants for these two lumped two-electron transfer reactions (k2f and k2r are in the units of 1/s and k3f and k3r are in the units of 1/M2.s) are such that the corresponding equilibrium constants (unitless) are defined by:

Keq,2=k2fk2r=k2af.k2bfk2ar.k2br=(KHCH)2exp(2F(Em,Fr/Fo0Em,Cr/Co0)RT) (A16)
Keq,3=k3fk3r=k3af.k3bfk3ar.k3br=exp(2F(Em,Cr/Co0Em,O2/O20)RT) (A17)

The lumped electron transfer reactions 1–3 (Equation (A3, A14, A15)) constitute the kinetic scheme of Fig. 1E, or equivalently Fig. 1F, which contain 6 kinetic parameters (k1f, k1r, k2f, k2r, k3f, k3r), subject to 3 thermodynamic constraints (Equation (A8, A16, A17)), resulting in 3 unknown kinetic parameters for estimation. Thus, the electron transfer reactions of Equation (A3, A14, A15) (Fig. 1E or 1F) was to derive the kinetic model of the NOX2 enzyme complex.

Derivations of NOX2 Reaction Flux Expression:

Correlating Fig. 1E and F, it is easy to see that Fo. 2Co, Fr. 2Co, and Fo. 2Cr denote E1.X, E2.X and E3.X, respectively. The following ordinary differential equations (ODEs), describing the rate of change of enzyme concentration in each of the three enzyme states, are derived from the reactions of Equation (A3, A14 and A15) based on the kinetic scheme of Fig. 1F:

dCE1.Xdt=k1fCNrCE1.Xk3rCO22CE1.X+k1rCNoCE2.X+k3fCO22CE3.X (A18)
dCE2.Xdt=k1fCNrCE1.Xk1rCNoCE2.Xk2fCE2.X+k2rCE3.X (A19)
dCE3.Xdt=k3rCO22CE1.X+k2fCE2.Xk2rCE3.Xk3fCO22CE3.X (A20)

The total enzyme concentration CE.X based on the principle of mass conservation is given by:

CE.X=CE1.X+CE2.X+CE3.X (A21)

where CE1.X, CE2.X, and CE3.X denote the concentrations of E1. X or Fo. 2Co, E2. X or Fr. 2Co, and E3. X or Fo. 2Cr, respectively, which are the three different enzyme states representing different redox status of the NOX2 enzyme complex; CNr and CNo are the concentrations of Nr and No, respectively; CO2 and CO2· are the concentrations of O2 and O2, respectively; knf and knr (n = 1, 2, 3) denote the forward and reverse rate constants for the three elementary rate limiting electron transfer reactions (Equation (A3, A14, A15)).

At steady state, the rate of change of relative portion of the NOX2 enzyme complex at each state is zero. Thus, the flux of the overall NOX2 reaction (Equation (A1)) can be calculated from any of the three elementary rate limiting electron transfer reactions. Now, Equation (A18A21) at steady state can be rewritten in the following normalized form as:

(k1fCNr+k3rCO22)CE1.XCE.X+k1rCNoCE2.XCE.X+k3fCO22CE2.XCE.X=0 (A22)
k1fCNrCE1.XCE.X(k1rCNo+k2f)CE2.XCE.X+k2rCE3.XCE.X=0 (A23)
k3rCO22CE1.XCE.X+k2fCE2.XCE.X(k2r+k3fCO22)CE3.XCE.X=0 (A24)
CE1.XCE.X+CE2.XCE.X+CE3.XCE.X=1 (A25)

In matrix notation, Equation (A22A25) can be rewritten as:

[k1fCNrk3rCO22K1rCNok3fCO22k1fCNrk1rCNok2fk2rk3rCO22k2fk2rk3fCO22111][CE1.XCE.XCE2.XCE.XCE3.XCE.X]=[0001] (A26)

Based on the solution of the above system of linear equations (Equation (A26)), the relative portion of the NOX2 enzyme complex concentration in each state is:

F1=CE1.XCE.X=(k1rk2rCNo+k2fk3fCO22+k1rk3fCNoCO22)/Den (A27)
F2=CE2.XCE.X=(k1fk2rCNr+k2rk3rCO22+k1fk3fCNrCO22)/Den (A28)
F3=CE3.XCE.X=(k1fk2fCNr+k2fk3rCO22+k1rk3rCNoCO22)/Den (A29)

where the common denominator (Den) in Equation (A27A29) is given by:

Den=k1fk2fCNr+k1fk2rCNr+k1fk3fCNrCO22+k1rk2rCNo+k1rk3fCNoCO22+k2fk3fCO22+k2rk3rCO22+k2fk3rCO22+k1rk3rCNoCO22 (A30)

The elementary rate limiting electron transfer reaction fluxes can be written as the flux of forward reaction – flux of reverse reaction, based on reactions in Equation (A3, A14 and A15):

J1=k1fCNrCE1.Xk1rCNoCE2.X (A31)
J2=k2fCE2.Xk2rCE3.X (A32)
J3=k3fCO22CE3.Xk3rCO22CE1.X (A33)

The overall NOX2 reaction flux can then be obtained from any of the rate limiting reaction fluxes as (J1 = J2 = J3 = JNOX2):

JNOX2=CE.X(k1fk2fk3fCNrCO22k1rk2rk3rCNoCO22)(k1fk2fCNr+k1fk2rCNr+k1fk3fCNrCO22+k1rk2rCNo+k1rk3fCNoCO22+k2fk3fCO22+k2rk3rCO22+k2fk3rCO22+k1rk3rCNoCO22) (A34)

where CE.X = CFlavoCytB represents the concentration of the total active NOX2 enzyme complex.

Simplifications for Sequential O2 Reduction by CytB:

During O2 reduction and O2 production, two electrons are transferred from two reduced CytB to two molecules of O2 in two sequential steps without cooperativity. However, our derivation above is based on the assumption that two electrons are transferred from two reduced CytB to two molecules of O2 in a single step with full cooperativity. To account for this discrepancy, we assume the following O2 and O2 dependencies of the forward and reverse rate constants k3f and k3r:

k3f=k3f/CO2andk3r=k3r/CO2 (A35)

Substituting Equation (A35) in Equation (A34) and dropping out the prime notation, the NOX2 reaction flux can then be expressed as:

JNOX2=CE.X(k1fk2fk3fCNrCO2k1rk2rk3rCNoCO22)(k1fk2fCNr+k1fk2rCNr+k1fk3fCNrCO2+k1rk2rCNo+k1rk3fCNoCO2+k2fk3fCO2+k2rk3rCO2+k2fk3rCO2+k1rk3rCNoCO2) (A36)

The corresponding NOX2 enzyme complex states (Equation A27A29) can be expressed as:

F1=CE1.XCE.X=(k1rk2rCNo+k2fk3fCO2+k1rk3fCNoCO2)/Den (A37)
F2=CE2.XCE.X=(k1fk2rCNr+k2rk3rCO2+k1fk3fCNrCO2)/Den (A38)
F3=CE3.XCE.X=(k1fk2fCNr+k2fk3rCO2+k1rk3rCNoCO2)/Den (A39)

where the common denominator (Den) in Equation (A37A39) is given by:

Den=k1fk2fCNr+k1fk2rCNr+k1fk3fCNrCO2+k1rk2rCNo+k1rk3fCNoCO2+k2fk3fCO2+k2rk3rCO2+k2fk3rCO2+k1rk3rCNoCO2 (A40)

Regulations by pH and Inhibitory Drugs:

Considering the postulated proton and drug bindings to the three different NOX2 enzyme complex states E1.X, E2.X and E3.X, the forward and reverse reaction rate constants knf and knr can be scaled appropriately to account for the effects of the proton and drug binding factors fnf and fnr, as schematized in Fig. 2A, B and 2C, and as detailed in the main texts:

knf=knffnfandknr=knrfnr (A41)

Substituting Equation (A41) into Equation (A36) and dropping out the prime notation, the NOX2 reaction flux can then be expressed as:

JNOX2=C(E.X)T(f1fk1ff2fk2ff3fk3fCNrCO2f1rk1rf2rk2rf3rk3rCNoCO2)(f1fk1ff2fk2fCNr+f1fk1ff2rk2rCNr+f1fk1ff3fk3fCNrCO2+f1rk1rf2rk2rCNo+f1rk1rf3fk3fCNoCO2+f2fk2ff3fk3fCO2+f2rk2rf3rk3rCO2+f2fk2ff3rk3rCO2+f1rk1rf3rk3rCNoCO2) (A42)

The corresponding NOX2 enzyme complex states (Equation (A37A39)) can be expressed as:

F1=C(E1.X)TC(E.X)T=(f1rk1rf2rk2rCNo+f2fk2ff3fk3fCO2+f1rk1rf3fk3fCNoCO2)/Den (A43)
F2=C(E2.X)TC(E.X)T=(f1fk1ff2rk2rCNr+f2rk2rf3rk3rCO2+f1fk1ff3fk3fCNrCO2)/Den (A44)
F3=C(E3.X)TC(E.X)T=(f1fk1ff2fk2fCNr+f2fk2ff3rk3rCO2+f1rk1rf3rk3rCNoCO2)/Den (A45)

where the common denominator (Den) in Equation (A43A45) is given by:

Den=f1fk1ff2fk2fCNr+f1fk1ff2rk2rCNr+f1rk1rf3fk3fCNrCO2+f1rk1rf2rk2rCNo+f1rk1rf3fk3fCNoCO2+f2fk2ff3fk3fCO2+f2rk2rf3rk3rCO2+f2fk2ff3rk3rCO2+f1rk1rf3rk3rCNoCO2 (A46)

References

  • [1].Bae YS, Oh H, Rhee SG, Yoo YD, Regulation of reactive oxygen species generation in cell signaling, Mol. Cell 32 (6) (2011) 491–509. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [2].Ray PD, Huang BW, Tsuji Y, Reactive oxygen species (ROS) homeostasis and redox regulation in cellular signaling, Cell. Signal 24 (5) (2012) 981–990. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [3].Brown DI, Griendling KK, Regulation of signal transduction by reactive oxygen species in the cardiovascular system, Circ. Res 116 (3) (2015) 531–549. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [4].Yarosz EL, Chang CH, The role of reactive oxygen species in regulating T cell-mediated immunity and disease, Immune Netw 18 (1) (2018) e14. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [5].Andreyev AY, Kushnareva YE, Starkov AA, Mitochondrial metabolism of reactive oxygen species, Biochemistry (Mosc.) 70 (2) (2005) 200–214. [DOI] [PubMed] [Google Scholar]
  • [6].Murphy MP, How mitochondria produce reactive oxygen species, Biochem. J 417 (1) (2009) 1–13. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [7].Stowe DF, Camara AK, Mitochondrial reactive oxygen species production in excitable cells: modulators of mitochondrial and cell function [invited comprehensive review], Antioxidants Redox Signal 11 (6) (2009) 1373–1414. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [8].Nita M, Grzybowski A, The role of the reactive oxygen species and oxidative stress in the pathomechanism of the age-related ocular diseases and other pathologies of the anterior and posterior eye segments in adults, Oxid Med Cell Longev (2016) 3164734 2016. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [9].Bokoch GM, Knaus UG, NADPH oxidases: not just for leukocytes anymore!, Trends Biochem. Sci 28 (9) (2003) 502–508. [DOI] [PubMed] [Google Scholar]
  • [10].Lambeth JD, NOX enzymes and the biology of reactive oxygen, Nat. Rev. Immunol 4 (3) (2004) 181–189. [DOI] [PubMed] [Google Scholar]
  • [11].Lambeth JD, Neish AS, Nox enzymes and new thinking on reactive oxygen: a double-edged sword revisited, Annual review of pathology 9 (2014) 119–145. [DOI] [PubMed] [Google Scholar]
  • [12].Bedard K, Krause KH, The NOX family of ROS-generating NADPH oxidases: physiology and pathophysiology, Physiol. Rev 87 (1) (2007) 245–313. [DOI] [PubMed] [Google Scholar]
  • [13].Griendling KK, Sorescu D, Ushio-Fukai M, NAD(P)H oxidase: role in cardiovascular biology and disease, Circ. Res 86 (5) (2000) 494–501. [DOI] [PubMed] [Google Scholar]
  • [14].Lambeth JD, Nox enzymes, ROS, and chronic disease: an example of antagonistic pleiotropy, Free Radic. Biol. Med 43 (3) (2007) 332–347. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [15].Lassegue B, Griendling KK, NADPH oxidases: functions and pathologies in the vasculature, Arterioscler. Thromb. Vasc. Biol 30 (4) (2010) 653–661. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [16].Sedeek M, Nasrallah R, Touyz RM, Hebert RL, NADPH oxidases, reactive oxygen species, and the kidney: friend and foe, J. Am. Soc. Nephrol 24 (10) (2013) 1512–1518. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [17].Thallas-Bonke V, Jandeleit-Dahm KA, Cooper ME, Nox-4 and progressive kidney disease, Curr. Opin. Nephrol. Hypertens 24 (1) (2015) 74–80. [DOI] [PubMed] [Google Scholar]
  • [18].Jha JC, Banal C, Okabe J, Gray SP, Hettige T, Chow BSM, Thallas-Bonke V, De Vos L, Holterman CE, Coughlan MT, Power DA, Skene A, Ekinci EI, Cooper ME, Touyz RM, Kennedy CR, Jandeleit-Dahm K, NADPH oxidase Nox5 accelerates renal injury in diabetic nephropathy, Diabetes 66 (10) (2017) 2691–2703. [DOI] [PubMed] [Google Scholar]
  • [19].Baehner RL, Gilman N, Karnovsky ML, Respiration and glucose oxidation in human and Guinea pig leukocytes: comparative studies, J. Clin. Invest 49 (4) (1970) 692–700. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [20].Rada B, Leto TL, Oxidative innate immune defenses by Nox/Duox family NADPH oxidases, Contrib. Microbiol 15 (2008) 164–187. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [21].Brown DI, Griendling KK, Nox proteins in signal transduction, Free Radic. Biol. Med 47 (9) (2009) 1239–1253. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [22].Thomas DC, The phagocyte respiratory burst: historical perspectives and recent advances, Immunol. Lett 192 (2017) 88–96. [DOI] [PubMed] [Google Scholar]
  • [23].Schrenzel J, Serrander L, Banfi B, Nusse O, Fouyouzi R, Lew DP, Demaurex N, Krause KH, Electron currents generated by the human phagocyte NADPH oxidase, Nature 392 (6677) (1998) 734–737. [DOI] [PubMed] [Google Scholar]
  • [24].DeCoursey TE, Cherny VV, Zhou W, Thomas LL, Simultaneous activation of NADPH oxidase-related proton and electron currents in human neutrophils, Proc. Natl. Acad. Sci. U. S. A 97 (12) (2000) 6885–6889. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [25].Murphy R, DeCoursey TE, Charge compensation during the phagocyte respiratory burst, Biochim. Biophys. Acta Bioenerg 1757 (8) (2006) 996–1011. [DOI] [PubMed] [Google Scholar]
  • [26].Uhlinger DJ, Inge KL, Kreck ML, Tyagi SR, Neckelmann N, Lambeth JD, Reconstitution and characterization of the human neutrophil respiratory burst oxidase using recombinant p47-phox, p67-phox and plasma membrane, Biochem. Biophys. Res. Commun 186 (1) (1992) 509–516. [DOI] [PubMed] [Google Scholar]
  • [27].Uhlinger DJ, Tyagi SR, Inge KL, Lambeth J, The respiratory burst oxidase of human neutrophils. Guanine nucleotides and arachidonate regulate the assembly of a multicomponent complex in a semirecombinant cell-free system, J. Biol. Chem 268 (12) (1993) 8624–8631. [PubMed] [Google Scholar]
  • [28].Uhlinger DJ, Taylor KL, Lambeth JD, p67-phox enhances the binding of p47-phox to the human neutrophil respiratory burst oxidase complex, J. Biol. Chem 269 (35) (1994) 22095–22098. [PubMed] [Google Scholar]
  • [29].Cross AR, Curnutte JT, The cytosolic activating factors p47phox and p67phox have distinct roles in the regulation of electron flow in NADPH oxidase, J. Biol. Chem 270 (12) (1995) 6543–6548. [DOI] [PubMed] [Google Scholar]
  • [30].Cross AR, Erickson RW, Curnutte JT, Simultaneous presence of p47(phox) and flavocytochrome b-245 are required for the activation of NADPH oxidase by anionic amphiphiles. Evidence for an intermediate state of oxidase activation, J. Biol. Chem 274 (22) (1999) 15519–15525. [DOI] [PubMed] [Google Scholar]
  • [31].Cross AR, p40(phox) Participates in the activation of NADPH oxidase by increasing the affinity of p47(phox) for flavocytochrome b(558), Biochem. J 349 (Pt 1) (2000) 113–117. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [32].Nauseef WM, Assembly of the phagocyte NADPH oxidase, Histochem. Cell Biol 122 (4) (2004) 277–291. [DOI] [PubMed] [Google Scholar]
  • [33].Cross AR, Parkinson JF, Jones OT, Mechanism of the superoxide-producing oxidase of neutrophils. O2 is necessary for the fast reduction of cytochrome b-245 by NADPH, Biochem. J 226 (3) (1985) 881–884. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [34].Babior BM, NADPH oxidase: an update, Blood 93 (5) (1999) 1464–1476. [PubMed] [Google Scholar]
  • [35].Babior BM, Lambeth JD, Nauseef W, The neutrophil NADPH oxidase, Arch. Biochem. Biophys 397 (2) (2002) 342–344. [DOI] [PubMed] [Google Scholar]
  • [36].Gabig TG, Kipnes RS, Babior BM, Solubilization of the O2(−)-forming activity responsible for the respiratory burst in human neutrophils, J. Biol. Chem 253 (19) (1978) 6663–6665. [PubMed] [Google Scholar]
  • [37].Bauldry SA, Nasrallah VN, Bass DA, Activation of NADPH oxidase in human neutrophils permeabilized with Staphylococcus aureus alpha-toxin. A lower Km when the enzyme is activated in situ, J. Biol. Chem 267 (1) (1992) 323–330. [PubMed] [Google Scholar]
  • [38].Nisimoto Y, Diebold BA, Cosentino-Gomes D, Lambeth JD, Nox4: a hydrogen peroxide-generating oxygen sensor, Biochemistry 53 (31) (2014) 5111–5120. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [39].Babior BM, Curnutte JT, McMurrich BJ, The particulate superoxide-forming system from human neutrophils. Properties of the system and further evidence supporting its participation in the respiratory burst, J. Clin. Invest 58 (4) (1976) 989–996. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [40].Suzuki Y, Lehrer RI, NAD(P)H oxidase activity in human neutrophils stimulated by phorbol myristate acetate, J. Clin. Invest 66 (6) (1980) 1409–1418. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [41].Morgan D, Cherny VV, Murphy R, Katz BZ, DeCoursey TE, The pH dependence of NADPH oxidase in human eosinophils, J. Physiol 569 (Pt 2) (2005) 419–431. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [42].Morgan D, Cherny VV, Murphy R, Xu W, Thomas LL, DeCoursey TE, Temperature dependence of NADPH oxidase in human eosinophils, J. Physiol 550 (Pt 2) (2003) 447–458. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [43].Cifuentes-Pagano E, Csanyi G, Pagano PJ, NADPH oxidase inhibitors: a decade of discovery from Nox2ds to HTS, Cell. Mol. Life Sci. : CMLS 69 (14) (2012) 2315–2325. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [44].Altenhofer S, Radermacher KA, Kleikers PW, Wingler K, Schmidt HH, Evolution of NADPH oxidase inhibitors: selectivity and mechanisms for target engagement, Antioxidants Redox Signal 23 (5) (2015) 406–427. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [45].Hirano K, Chen WS, Chueng AL, Dunne AA, Seredenina T, Filippova A, Ramachandran S, Bridges A, Chaudry L, Pettman G, Allan C, Duncan S, Lee KC, Lim J, Ma MT, Ong AB, Ye NY, Nasir S, Mulyanidewi S, Aw CC, Oon PP, Liao S, Li D, Johns DG, Miller ND, Davies CH, Browne ER, Matsuoka Y, Chen DW, Jaquet V, Rutter AR, Discovery of GSK2795039, a novel small molecule NADPH oxidase 2 inhibitor, Antioxidants Redox Signal 23 (5) (2015) 358–374. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [46].Green TR, Shangguan X, Stoichiometry of O2 metabolism and NADPH oxidation of the cell-free latent oxidase reconstituted from cytosol and solubilized membrane from resting human neutrophils, J. Biol. Chem 268 (2) (1993) 857–861. [PubMed] [Google Scholar]
  • [47].Qi F, Dash RK, Han Y, Beard DA, Generating rate equations for complex enzyme systems by a computer-assisted systematic method, BMC Bioinf 10 (2009) 238. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [48].Alberty RA, Thermodynamics of Biochemical Reactions, John Wiley & Sons, Hoboken, N.J., 2003. [Google Scholar]
  • [49].McNally JS, Davis ME, Giddens DP, Saha A, Hwang J, Dikalov S, Jo H, Harrison DG, Role of xanthine oxidoreductase and NAD(P)H oxidase in endothelial superoxide production in response to oscillatory shear stress, Am. J. Physiol. Heart Circ. Physiol 285 (6) (2003) H2290–H2297. [DOI] [PubMed] [Google Scholar]
  • [50].Zhao K, Huang Z, Lu H, Zhou J, Wei T, Induction of inducible nitric oxide synthase increases the production of reactive oxygen species in RAW264.7 macrophages, Biosci. Rep 30 (4) (2010) 233–241. [DOI] [PubMed] [Google Scholar]
  • [51].Hrycay EG, Bandiera SM, Involvement of cytochrome P450 in reactive oxygen species formation and cancer, Adv. Pharmacol 74 (2015) 35–84. [DOI] [PubMed] [Google Scholar]
  • [52].Pannala VR, Bazil JN, Camara AK, Dash RK, A biophysically based mathematical model for the catalytic mechanism of glutathione reductase, Free Radic. Biol. Med 65 (2013) 1385–1397. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [53].Pannala VR, Bazil JN, Camara AK, Dash RK, A mechanistic mathematical model for the catalytic action of glutathione peroxidase, Free Radic. Res 48 (4) (2014) 487–502. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [54].Bazil JN, Pannala VR, Dash RK, Beard DA, Determining the origins of super-oxide and hydrogen peroxide in the mammalian NADH:ubiquinone oxidoreductase, Free Radic. Biol. Med 77 (2014) 121–129. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [55].Pannala VR, Dash RK, Mechanistic characterization of the thioredoxin system in the removal of hydrogen peroxide, Free Radic. Biol. Med 78 (2015) 42–55. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [56].Bjelakovic G, Nikolova D, Gluud LL, Simonetti RG, Gluud C, Mortality in randomized trials of antioxidant supplements for primary and secondary prevention: systematic review and meta-analysis, J. Am. Med. Assoc 297 (8) (2007) 842–857. [DOI] [PubMed] [Google Scholar]
  • [57].Dotan Y, Pinchuk I, Lichtenberg D, Leshno M, Decision analysis supports the paradigm that indiscriminate supplementation of vitamin E does more harm than good, Arterioscler. Thromb. Vasc. Biol 29 (9) (2009) 1304–1309. [DOI] [PubMed] [Google Scholar]
  • [58].Simchowitz L, Intracellular pH modulates the generation of superoxide radicals by human neutrophils, J. Clin. Invest 76 (3) (1985) 1079–1089. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [59].Rotstein OD, Nasmith PE, Grinstein S, The Bacteroides by-product succinic acid inhibits neutrophil respiratory burst by reducing intracellular pH, Infect. Immun 55 (4) (1987) 864–870. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [60].Doussiere J, Gaillard J, Vignais PV, The heme component of the neutrophil NADPH oxidase complex is a target for aryliodonium compounds, Biochemistry 38 (12) (1999) 3694–3703. [DOI] [PubMed] [Google Scholar]
  • [61].Gorin Y, Nox4 as a potential therapeutic target for treatment of uremic toxicity associated to chronic kidney disease, Kidney Int 83 (4) (2013) 541–543. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [62].Rastogi R, Geng X, Li F, Ding Y, NOX activation by subunit interaction and underlying mechanisms in disease, Front. Cell. Neurosci 10 (2016) 301. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [63].Teixeira G, Szyndralewiez C, Molango S, Carnesecchi S, Heitz F, Wiesel P, Wood JM, Therapeutic potential of NADPH oxidase ¼ inhibitors, Br. J. Pharmacol 174 (12) (2017) 1647–1669. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [64].Cross AR, Parkinson JF, Jones OT, The superoxide-generating oxidase of leucocytes. NADPH-dependent reduction of flavin and cytochrome b in solubilized preparations, Biochem. J 223 (2) (1984) 337–344. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • [65].Severns C, Collins-Lech C, Sohnle PG, Effect of temperature on production of hypochlorous acid by stimulated human neutrophils, J. Lab. Clin. Med 107 (1) (1986) 29–35. [PubMed] [Google Scholar]
  • [66].Umeki S, DeLisle DM, A microtechnique for neutrophil respiratory burst oxidase in a cell-free system–characterization of oxidase activation system, Comp. Biochem. Physiol. B 96 (3) (1990) 461–464. [DOI] [PubMed] [Google Scholar]
  • [67].Maridonneau-Parini I, Malawista SE, Stubbe H, Russo-Marie F, Polla BS, Heat shock in human neutrophils: superoxide generation is inhibited by a mechanism distinct from heat-denaturation of NADPH oxidase and is protected by heat shock proteins in thermotolerant cells, J. Cell. Physiol 156 (1) (1993) 204–211. [DOI] [PubMed] [Google Scholar]
  • [68].Salman H, Bergman M, Bessler H, Alexandrova S, Djaldetti M, Ultrastructure and phagocytic activity of rat peritoneal macrophages exposed to low temperatures in vitro, Cryobiology 41 (1) (2000) 66–71. [DOI] [PubMed] [Google Scholar]

RESOURCES