首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 218 毫秒
1.
Nitric oxide (NO) generation by soybean (Glycine max, var ADM 4800) chloroplasts was studied by electron paramagnetic resonance (EPR) spin-trapping technique.1 Both nitrite and L-arginine (arg) are the required substrates for enzymatic activities considered as possible sources of NO in plants. Soybean chloroplasts showed a NO production of 3.2 ± 0.2 nmol min−1 mg−1 protein in the presence of 1 mM NaNO2. Chloroplasts incubated with 1 mM arg showed a NO production of 0.76 ± 0.04 nmol min−1 mg−1 protein. This production was inhibited when chloroplasts were incubated in presence of NOS-inhibitors L-NAME and L-NNA. In vitro exposure of chloroplasts to a NO-donor (GSNO) decreased both ascorbyl radical content and the activity of ascorbate peroxidase, without modification of the total ascorbate content. Exposure of the isolated chloroplasts to a NO-donor decreased lipid radical content in membranes, however, incubation in the presence of 25 µM peroxynitrite (ONOO) led to an increase in lipid-derived radicals (34%). The effect of ONOO on protein oxidation was determined by western blotting, showing an increase in carbonyl content either in stroma or thylakoid proteins as compared to control. Taken as a whole, NO seems to be an endogenous metabolite in soybean chloroplasts and reactive nitrogen species could exert either antioxidant or prooxidant effects on chloroplasts, since both a decreased lipid radical content in membranes and a decrease in the activity of ascorbate peroxidase were observed after exposure to a NO donor.Key Words: ascorbate, ascorbate peroxidase, chloroplasts, nitric oxide, peroxynitriteThe origin of nitric oxide (NO) in plants under aerobic conditions is currently under study. Although plants with low level of arginine shows arg-stimulated NO accumulation,2 the mechanism for arginine-dependent NO synthesis in plants is still unknown, because the detection of an animal-type NOS remains elusive to date.3,4 Even though assimilatory nitrate reductase is an enzymatic source of NO, its role in vivo would be limited by both its cytosolic localization which difficult the availability for nitrite, and the relative high Km for nitrite (100 µM).5Chloroplasts have been previously marked as NO sources based in nonquantitative studies employing fluorescence microscopy6,7 and immunogold electron microscopy.8 In our work we employed an specific technique (EPR, electron paramagnetic resonance with spin trap9) to detect NO as an endogenous metabolite and to quantify its generation in the presence of different substrates. In order to gain insight on the mechanism leading to NO production both nitrite-dependent and arg-dependent pathways were evaluated. In the presence of 1 mM arg and 0.1 mM NADPH the rate of NO generation was 0.76 ± 0.04 nmol min−1 mg−1 prot (arg-dependent synthesis). The synthesis of NO resulted completely blocked in the presence of arg analogs (L-NAME and L-NNA). It is important to point out that the content of arg in the chloroplasts stroma is high as compared to the content of other amino acids (56.7 ± 0.8 nmol mg−1 prot), suggesting that this pathway could be operative under physiological conditions.Soybean chloroplasts showed a NO production of 3.2 ± 0.2 nmol min−1 mg−1 prot in the presence of 1 mM NaNO2. Furthermore, NO generation was detected in the presence of nitrite concentrations as low as 25 mM. Since nitrite-dependent NO generation resulted inhibited by 50% by the addition of DCMU, and no NO generation was measured in the stroma fraction, thylakoidal electron transport seems to be a key feature in NO synthesis.According to this scenario and assuming that the two independent pathways for NO generation in chloroplasts are operative, the total rate of production of NO could be understood as the generation by the activity of an arg-dependent enzyme and by a NO2 dependent pathway, as indicated by eq. 1.d[NO]dt=(d[NO]dt)NOS like+(d[NO]dt)NO2(1)Regarding the NO disappearance, from a kinetic point of view, the rate of the reaction of NO with O2 to generate peroxynitrite seems to be the main pathway, since the reaction is diffusionally controlled. Thus, the rate of disappearance of NO could be estimated from the rate of generation of ONOO (eq. 2); however, other reactions should be considered under nonphysiological conditions.d[NO]dt=dONOOdt=k[NO][O2](2)NO generation rate should be equal to NO consumption rate in order to keep a physiological NO steady state concentration (eq. 3)d[NO]dt=d[NO]dt(3)Thus, replacing NO generation and disappearance rates by those rates indicated in equations 1 and 2, (d[NO]dt)NOS like+(d[NO]dt)NO2=k[NO][O2](4) The data obtained under unrestricted availability of substrates, indicate a generation rate of NO by the activity of a NOS-like enzyme of 13 × 10−9 M s−1. Chloroplastic NO generation rate in the presence of 100 µM NO2 was 14 × 10−9 M s−1. Thus, according to equation 1, the rate of generation of NO is approximately 3 × 10−8 M s−1. Assuming a steady state concentration for O2 of 1 nM in chloroplasts10 and a rate constant (k) of 6.9 × 109 M−1 s−1 for the reaction between O2 and NO,11 a steady state concentration of 4 nM for NO in the chloroplast could be estimated. Since under in vivo conditions chloroplasts may content the required substrates for the NO synthesis, the assays presented here strongly suggest that a feasible NO production could take place inside the chloroplasts. However, nonsupplemented chloroplasts did not show any NO-dependent EPR signal. This observation agrees with the fact that NO steady state concentration under physiological conditions as was calculated here (4 nM) is below the EPR detection limit (500 nM).12Further studies should be performed to characterize NO oxidative effects on chloroplasts. Scavenging of O2 and H2O2 is essential for chloroplasts to maintain their ability to fix CO2 since several enzymes in the CO2-reduction cycle are sensitive to active oxygen species.13 These organelles lacking catalase, contain a significant peroxidase activity.14 H2O2-reduction catalized by ascorbate peroxidase (AP) lead to ascorbate oxidation and produces ascorbyl radical (A.).15 In isolated chloroplast the content of A.. was evaluated in DMSO based extract by EPR16. Quantification of EPR signals indicated that A. content in control chloroplasts (123 ± 5 pmol mg−1 prot) decreased after exposure to NO (Fig. 1). The total content of ascorbate, assessed by an HPLC technique17 in chloroplasts isolated from soybean leaves exposed to NO was not significantly different from the measured content in chloroplasts not exposed to the NO donor (Fig. 1). The activity of AP was significantly decreased by 48, 53 and 54% after exposure of the chloroplasts to NO-donor. Previous data suggested that AP could be inactivated by NO via oxidation of functional thiols.18 Besides, the reversible inhibition of AP could be due to the formation of Fe-nitrosyl complexes between NO and the Fe atom of the heme group, as it was previously described for NO-mediated activation of guanylate cyclase and the inhibition of cytochrome P450 and catalase in mammals.19 The data presented here showed that in isolated chloroplasts exposed to a NO donor, there could be either a limited damage associated to the decrease in the content of A.. or an increased cellular deterioration by the decrease in the activity of the enzyme responsible for the scavenging of H2O2.Open in a separate windowFigure 1Ascorbate metabolism in soybean chloroplasts after NO exposure. A.. content (▪), ascorbate content (▪), and AP activity (*) as a function of the exposure of isolated chloroplasts to GSNO in the presence of 50 µM DTT. * = significantly different at p ≥ 0.05 from the value obtained in the absence of GSNO + 50 µM DTT.Thus, in situ generation of NO could play a protective role in preventing the oxidation of chloroplastic lipids; however, the reaction of NO with O2 leading to ONOO production may result in a potential source of damage or as it is shown here by the significant decrease of the AP activity that consumes H2O2. NO is a suitable candidate to modulate cellular H2O2 level through the chloroplast function, as an initial step to regulate complex metabolic pathways directed to activate physiological responses, defense pathways or deleterious effects in the cytosol. Furthermore, an integrated study on the effect of nitrogen reactive species is required under stress conditions to characterize the metabolic pathways involved in the resulting cellular damage.  相似文献   

2.
3.
The ability of biomolecules to fold and to bind to other molecules is fundamental to virtually every living process. Advanced experimental techniques can now reveal how single biomolecules fold or bind against mechanical force, with the force serving as both the regulator and the probe of folding and binding transitions. Here, we present analytical expressions suitable for fitting the major experimental outputs from such experiments to enable their analysis and interpretation. The fit yields the key determinants of the folding and binding processes: the intrinsic on-rate and the location and height of the activation barrier.Dynamic processes in living cells are regulated through conformational changes in biomolecules—their folding into a particular shape or binding to selected partners. The ability of biomolecules to fold and to bind enables them to act as switches, assembly factors, pumps, or force- and displacement-generating motors (1). Folding and binding transitions are often hindered by a free energy barrier. Overcoming the barrier requires energy-demanding rearrangements such as displacing water from the sites of native contacts and breaking nonnative electrostatic contacts, as well as loss of configurational entropy. Once the barrier is crossed, the folded and bound states are stabilized by short-range interactions: hydrogen bonds, favorable hydrophobic effects, and electrostatic and van der Waals attractions (2).Mechanistic information about folding and binding processes is detailed in the folding and binding trajectories of individual molecules: observing an ensemble of molecules may obscure the inherent heterogeneity of these processes. Single-molecule trajectories can be induced, and monitored, by applying force to unfold/unbind a molecule and then relaxing the force until folding or binding is observed (3–5) (Fig. 1). Varying the force relaxation rate shifts the range of forces at which folding or binding occurs, thus broadening the explorable spectrum of molecular responses to force and revealing conformational changes that are otherwise too fast to detect. The measured force-dependent kinetics elucidates the role of force in physiological processes (6) and provides ways to control the timescales, and even the fate, of these processes. The force-dependent data also provides a route to understanding folding and binding in the absence of force—by extrapolating the data to zero force via a fit to a theory.Open in a separate windowFigure 1Schematic of the output from a force-relaxation experiment. The applied force is continuously relaxed from the initial value F0 until the biomolecule folds or binds, as signified by a sharp increase in the measured force. From multiple repeats of this experiment, distributions of the folding or binding forces are collected (inset). Fitting the force distributions with the derived analytical expression yields the key parameters that determine the kinetics and energetics of folding or binding.In this letter, we derive an analytical expression for the distribution of transition forces, the major output of force-relaxation experiments that probe folding and binding processes. The expression extracts the key determinants of these processes: the on-rate and activation barrier in the absence of force. The theory is first developed in the context of biomolecular folding, and is then extended to cover the binding of a ligand tethered to a receptor. In contrast to unfolding and unbinding, the reverse processes of folding and binding require a theory that accounts for the compliance of the unfolded state, as well as the effect of the tether, to recover the true kinetic parameters of the biomolecule of interest.In a force-relaxation experiment, an unfolded biomolecule or unbound ligand-receptor complex is subject to a stretching force, which is decreased from the initial value F0 as the pulling device approaches the sample at speed V until a folding or binding transition is observed (Fig. 1) (3–5). Define S(t) as the probability that the molecule has not yet escaped from the unfolded (implied: or unbound) state at time t. When escape is limited by one dominant barrier, S(t) follows the first-order rate equationS˙(t)dS(t)dt=k(F(t))S(t),where k(F(t)) is the on-rate at force F at time t. Because, prior to the transition, the applied force decreases monotonically with time, the distribution of transition forces, p(F), is related to S(t) through p(F)dF=S˙(t)dt, yieldingp(F)=k(F)F˙(F)eF0Fk(F)F˙(F)dF.(1)Here F˙(F)dF(t)/dt<0 is the force relaxation rate. The proper normalization of p(F) is readily confirmed by integrating Eq. 1 from the initial force F0 to negative infinity, the latter accounting for transitions that do not occur by the end of the experiment. Note that the expression for the distribution of folding/binding forces in Eq. 1 differs from its analog for the unfolding process (7) by the limits of integration and a negative sign, reflecting the property of a relaxation experiment to decrease the survival probability S(t) by decreasing the force. Converting the formal expression in Eq. 1 into a form suitable for fitting experimental data requires establishing functional forms for k(F) and F˙(F) and analytically solving the integral. These steps are accomplished below.The on-rate k(F) is computed by treating the conformational dynamics of the molecule as a random walk on the combined free energy profile G(x,t) = G0(x) + Gpull(x,t) along the molecular extension x. Here G0(x) is the intrinsic molecular potential and Gpull(x,t) is the potential of the pulling device. When G(x,t) features a high barrier on the scale of kBT (kB is the Boltzmann constant and T the temperature), the dynamics can be treated as diffusive. The unfolded region of the intrinsic potential for a folding process, unlike that for a barrierless process (8), can be captured by the functionG0(x)=ΔGν1ν(xx)11νΔGν(xx),which has a sharp (if ν = 1/2, Fig. 2, inset) or smooth (if ν = 2/3) barrier of height ΔG and location x. The potential of a pulling device of stiffness κS is Gpull(x,t) = κS/2(X0Vtx)2 with an initial minimum at X0 (corresponding to F0). Applying Kramers formalism (9) to the combined potential G(x,t), we establish the analytical form of the on-rate at force F(t),k(F)=k0(1+κSκU(F))1ν12(1+νFxΔG)1ν1×eβΔG[1(1+κSκU(F))2ν1ν1(1+νFxΔG)1ν],where k0 is the intrinsic on-rate, β ≡ (kBT)−1, andκU(F)=ν(1ν)2ΔGx2(1+νFxΔG)21νis the stiffness of the unfolded biomolecule under force F (see the Supporting Material for details on all derivations). The full nonlinear form of Gpull(x,t) was necessary in the derivation because, in contrast to the typically stiff folded state, the unfolded state may be soft (to be exact, 1/2κS x‡2(F) << kBT may not be satisfied) and thus easily deformed by the pulling device. Because of this deformation, the folding transition faces an extra contribution (regulated by the ratio κS/κU(F)) to the barrier height, typically negligible for unfolding, that decreases the on-rate in addition to the applied force F.Open in a separate windowFigure 2Contributions to the free energy profile for folding (inset) and binding (main figure). The derived expression (Eq. 2) extracts the on-rate and the location and height of the activation barrier to folding. When applied to binding data, the expression extracts the parameters of the ligand-tether-receptor (LTR) potential G˜0 (x); the proposed algorithm (Eqs. 3 and 4) removes the contribution of the tether potential Gteth(x) to recover the parameters of the intrinsic ligand-receptor (LR) potential G0(x).The last piece required for Eq. 1, the loading rate F˙(F), is computed as the time derivative of the force F(t) on the unfolded molecule at its most probable extension at time t:F˙(F)=κSV1+κS/κU(F).Finally, we realize that the integral in Eq. 1 can be solved analytically exactly, both for ν = 1/2 and ν = 2/3, resulting in the analytical expression for the distribution of folding forces:p(F)=k(F)|F˙(F)|ek(F)β|F˙(F)|x(1+κSκU(F))νν1(1+νFxΔG)11ν.(2)Equation 2 can be readily applied to (normalized) histograms from force-relaxation experiments to extract the parameters of the intrinsic kinetics and energetics of folding. Being exact for ν = 1/2 and ν = 2/3, Eq. 2 is also an accurate approximation for any ν in the interval 1/2 < ν < 2/3 as long as κSκU (F) (see Fig. S1 in the Supporting Material). For simplicity, in Eq. 2 we have omitted the term containing F0 as negligible if F0 is large enough to prevent folding events.The solution in Eq. 2 reveals properties of the distribution of folding forces that distinguish it from its unfolding counterpart (7):
  • 1.The distribution has a positive skew (Fig. 3), as intuitively expected: the rare folding events occur at high forces when the barrier is still high.Open in a separate windowFigure 3Force histograms from folding (left) and binding (right) simulations at several values of the force-relaxation speed (in nanometers per second, indicated at each histogram). Fitting the histograms with the analytical expression in Eq. 2 (lines) recovers the on-rate and activation barrier for folding or binding (2.Increasing the relaxation speed shifts the distribution to lower forces (Fig. 3): faster force relaxation leaves less time for thermal fluctuations to push the system over a high barrier, causing transitions to occur later (i.e., at lower forces), when the barrier is lower.
  • 3.The stiffness κS and speed V enter Eq. 2 separately, providing independent routes to control the range of folding forces and thus enhance the robustness of a fit.
The application of the above framework to binding experiments on a ligand and receptor connected by a tether (3) involves an additional step—decoupling the effect of the tether—to reconstruct the parameters of ligand-receptor binding. Indeed, the parameters extracted from a fit of experimental histograms to Eq. 2 characterize the ligand-tether-receptor (LTR) potential (k˜0, x˜, ΔG˜, ν) (Fig. 2). The parameters of the natural ligand-receptor (LR) potential (k0, x, ΔG) can be recovered using three characteristics of the tether: contour length L; persistence length p; and extension Δℓ of the tether along the direction of the force in the LTR transition state. The values of L and p can be determined from the force-extension curve of the tether (10); these define the tether potential Gteth(x) (Fig. 2). The value of Δℓ can be found from an unbinding experiment (7) on LTR and the geometry of the tether attachment points (see Fig. S3). Approximating the region of the LR potential between the transition and unbound states as harmonic, with no assumptions about the shape of the potential beyond x, the ligand-receptor barrier parameters are thenx=α1α2x˜,ΔG=(α1)22(α2)x˜Fteth(Δ+x˜),(3)and the intrinsic unimolecular association rate isk0k˜0(βΔG)32(βΔG˜)1ν12(x˜x)2eβ(ΔG˜ΔG).(4)Here, the force value Fteth(Δ+x˜) is extracted from the force-extension curve of the tether at extension Δ+x˜ andα=2(ΔG˜Gteth(Δ)+Gteth(Δ+x˜))x˜Fteth(Δ+x˜),where Gteth(x) is the wormlike-chain potential (see Eq. S13 in the Supporting Material). Equations 3–4 confirm that a tether decreases the height and width of the barrier (see Fig. 2), thus increasing the on-rate.In Fig. 3, the developed analytical framework is applied to folding and binding force histograms from Brownian dynamics simulations at parameters similar to those in the analogous experimental and computational studies (3,5,11) (for details on simulations and fitting procedure, see the Supporting Material). For the stringency of the test, the simulations account for the wormlike-chain nature of the molecular unfolded and LTR unbound states that is not explicitly accounted for in the theory. With optimized binning (12) of the histograms and a least-squares fit, Eqs. 2–4 recover the on-rate, the location and the height of the activation barrier, and the value of ν that best captures how the kinetics scale with force (
  • 1.Multiple relaxation speeds,
  • 2.Folding/binding events at low forces, and
  • 3.A large number of events at each speed.
  • Table 1

    On-rate and the location and height of the activation barrier from the fit of simulated data to the theory in
    Eq. 2
    Foldingk0 (s−1)x (nm)ΔG (kBT)ν
     True9.5 × 1032.22.0
     Fit8 ± 2 × 1032.2 ± 0.21.8 ± 0.50.54a
    Binding (LTR)k˜0 (s−1)x˜ (nm)ΔG˜ (kBT)ν
     True281.561.7
     Fit24 ± 31.57 ± 0.091.8 ± 0.40.53a
    Binding (LR)k0 (s−1)x (nm)ΔG (kBT)
     True2.83.04.0
     Fit2.7 ± 0.22.9 ± 0.14.1 ± 0.1
    Open in a separate windowaFixed at value that minimized least-squares error.  相似文献   

    4.
    Recent development of titratable coions has paved the way for realizing all-atom molecular dynamics at constant pH. To further improve physical realism, here we describe a technique in which proton titration of the solute is directly coupled to the interconversion between water and hydroxide or hydronium. We test the new method in replica-exchange continuous constant pH molecular dynamics simulations of three proteins, HP36, BBL, and HEWL. The calculated pKa values based on 10-ns sampling per replica have the average absolute and root-mean-square errors of 0.7 and 0.9 pH units, respectively. Introducing titratable water in molecular dynamics offers a means to model proton exchange between solute and solvent, thus opening a door to gaining new insights into the intricate details of biological phenomena involving proton translocation.Solution pH is an important factor in biology. Although neutral pH in extracellular medium accounts for balanced electrostatics and proper folding of protein structures, pH gradients across cell membranes induce large conformational changes that are necessary for biological functions, such as ATP synthesis and efflux of small molecules out of the cell. To gain detailed insights into pH-dependent conformational phenomena, several constant pH molecular dynamics (pHMD) methods, based on either discrete or continuous titration coordinates, have been developed in the last decade (1–4). In the continuous pHMD (CpHMD) framework (2,4), a set of titration coordinates {λi} are simultaneously propagated along with the conformational degrees of freedom. Although the original CpHMD method based on the generalized Born (GB) implicit-solvent models (2,4) offers quantitative prediction of pKa values and pH dependence of folding and conformational dynamics of proteins (5), its accuracy and applicability to highly charged systems and those with dominantly hydrophobic regions are limited due to the approximate nature of the underlying implicit-solvent models.Motivated by the above-mentioned need, three groups have made efforts to develop a CpHMD method using exclusively the explicit-solvent models (6–8). In our development, the titration of acidic and basic sites is coupled with that of coions to level the total charge of the system (8). To further improve physical realism, here we replace the coions by titratable water molecules, which not only absorb the excess charge but also enable direct modeling of solute-solvent proton exchange in classical molecular dynamics simulations.To illustrate the utility of the new methodology, we applied it to the titration simulations of three proteins that were previously used to benchmark the GB-based CpHMD. Although this work does not explore specific interactions between titratable waters and proteins, the methodology can be further tested or improved to provide a rigorous way for modeling proton transfer in molecular dynamics, which is a computationally efficient alternative to the empirical valence-bond theory-based methodologies (9,10).We define titration of water as:
    • 1.Loss of a proton to give a negatively charged hydroxide,
    H2O ? OH? + H+, (1)or
    • 2.Gain of a proton to give a positively charged hydronium,
    H2O + H+ ? H3O+.(2)We now couple the titration of hydroxide (Eq. 1) with that of an acidic site of the solute in the CpHMD simulation,HA+OHKaA+H2O.(3)The use of hydronium is avoided here to prevent a potential artifact due to prolonged attraction with A. Analogously, we couple the titration of hydronium (Eq. 2) with that of a basic site,BH++H2OKbH3O++B.(4)Thus, effectively, a proton is transferred between the solute and solvent. However, we should note that in CpHMD simulations, titratable protons are represented by covalently attached dummies (2,4). Through varying the atomic charges and van der Waals interactions, they are seen by other atoms in the protonated state but not in the unprotonated state (see Table S1 in the Supporting Material). Furthermore, the solution proton concentration is implicitly modeled through a free energy term (2,4).In CpHMD, the reference potential of mean force (PMF) for titration is that of the model compound (blocked single amino acid in water) along λ (2,4). In the presence of cotitrating water molecules, it is necessary to add the PMF for the conversion of water to hydroxide or hydronium. One-nanosecond NPT simulations at ambient pressure and temperature were performed to calculate the average force, 〈dU/d,θ〉 at given θ-values, which are related to λ by λ = sin2 θ (see Fig. S1 in the Supporting Material). Thermodynamic integration was then applied to calculate the PMF. We found that the average force can be accurately fit when assuming the PMF is quadratic in λ (Fig. 1). The same applies to the PMFs for titration of models Asp, Glu, and His. After testing on the titration of model compounds (see Table S2), we performed 10-ns all-atom CpHMD simulations with the pH replica-exchange protocol for three proteins: HP36, BBL and HEWL (see the Supporting Material for details). Most of the calculated pKa values were converged in 10 ns per replica (see Fig. S3). Results are summarized in Fig. S4. Based on the 10-ns data, the root-mean-square (RMS) and average absolute errors are 0.9 and 0.7 pH units, respectively, while the largest absolute error is 2.5 (Glu35 of HEWL). Linear regression of the calculation versus experiment gives R2 of 0.8 and slope of 1.2.Open in a separate windowFigure 1Average force and potential of mean force for converting a water molecule to hydroxide (A) and hydronium. (B) (Data points) Average forces. (Dashed curves) Best fits using a linear function, 2A(λB). (Solid curves) Corresponding potential of mean force.

    Table 1

    Calculated and experimental pKa values of three proteins
    ResidueExperimenta
    GBa
    All-atom CpHMD
    Time (ns)b0–10–55–100–10
    HP36
     Asp443.10 (0.01)3.2 (0.1)2.03.02.6 (0.5)
     Glu453.95 (0.01)3.5 (0.1)4.34.54.4 (0.1)
     Asp463.45 (0.12)3.5 (0.1)2.43.73.1 (0.6)
     Glu724.37 (0.03)3.5 (0.1)4.44.44.4 (0.0)
    BBL
     Asp1293.88 (0.02)3.2 (0.0)2.23.22.7 (0.5)
     Glu1414.46 (0.04)4.3 (0.0)4.04.44.2 (0.2)
     His1426.47 (0.04)7.1 (0.0)5.95.85.8 (0.0)
     Asp1453.65 (0.04)2.8 (0.2)3.03.13.1 (0.0)
     Glu1613.72 (0.05)3.6 (0.3)4.23.94.0 (0.2)
     Asp1623.18 (0.04)3.4 (0.3)2.93.53.2 (0.3)
     Glu1644.50 (0.03)4.5 (0.1)5.74.65.2 (0.6)
     His1665.39 (0.02)5.4 (0.1)4.44.44.4 (0.0)
    HEWL
     Glu72.6 (0.2)2.6 (0.1)3.63.43.5 (0.1)
     His155.5 (0.2)5.3 (0.5)5.15.15.1 (0.0)
     Asp182.8 (0.3)2.9 (0.0)2.53.32.9 (0.4)
     Glu356.1 (0.4)4.4 (0.2)8.58.78.6 (0.1)
     Asp481.4 (0.2)2.8 (0.2)−0.11.10.6 (0.6)
     Asp523.6 (0.3)4.6 (0.0)5.45.65.5 (0.1)
     Asp661.2 (0.2)1.2 (0.4)−0.60.80.3 (0.7)
     Asp872.2 (0.1)2.0 (0.1)0.82.11.5 (0.7)
     Asp1014.5 (0.1)3.3 (0.3)6.15.75.9 (0.2)
     Asp1193.5 (0.3)2.5 (0.1)3.03.33.2 (0.1)
    Maximum absolute deviation1.82.42.62.5
    Average absolute deviation (RMS deviation)0.5 (0.7)1.0 (1.2)0.6 (0.9)0.7 (0.9)
    Linear fit R2 (slope)0.7 (0.8)0.8 (1.4)0.7 (1.1)0.8 (1.2)
    Open in a separate windowaTaken from Wallace and Shen (12). The pKa''s of BBL were recalculated.bSampling time per pH replica.Breaking the simulations in two halves, we noticed that the second 5-ns sampling gave better agreement with experiment. The RMS deviation is reduced from 1.2 to 0.9 pH units, while the average absolute deviation is reduced from 1.0 to 0.6 pH units. The linear regression against experimental data is also improved, with the slope decreasing from 1.4 to 1.1 although R2 remains the same. Comparing these second-half results with the GB-based simulations, we find that the RMS and average absolute deviations are about the same as the GB-CpHMD results; however, the all-atom simulations show a small systematic overestimation (regression slope >1), whereas GB simulations show a systematic underestimation (regression slope <1).The improvement in the second halves of the simulations are seen mainly for residues involved in attractive electrostatic interactions, including Asp44 and Asp46 of HP36, Asp129 of BBL, and Asp48, Asp66, and Asp87 of HEWL. These residues are initially locked in salt-bridges or hydrogen bonds. However, in the second 5 ns, the attractive interactions weakened, leading to a decrease in the calculated pKa shifts relative to the model values and better agreement with experiment. For instance, Asp44 was initially in a salt-bridge distance from Arg55. However, the salt-bridge positions were sampled less often in the second 5 ns (see Fig. S5), which explains the 1-unit reduction in the calculated pKa shift. Significant fluctuation in ion-pair interactions was also observed in the work by Alexov (11). The carboxyl oxygen of Asp46 was a hydrogen-bond acceptor with both the backbone amide and hydroxyl of Ser43. These hydrogen bonds were less frequently sampled in the second 5 ns (see Fig. S6), leading to a decrease of the pKa shift for Asp46 by 1.3 units. These results indicate that extensive conformational sampling is necessary to give an accurate estimate of the ratio between the charged and neutral populations.Limited conformational sampling is also a contributing factor to the overestimation of the pKa shifts for buried residues (Fig. S7 and Fig. S8). The increase in SASA is correlated with the more frequent sampling of the states with λ close to 1, i.e., the deprotonated form (see Fig. S9). However, because Glu35 was buried in the starting conformation and the transition between buried and exposed states is slow compared to the simulation length, the exposed state may not be sufficiently sampled, leading to overestimation of the pKa shift.In contrast to Glu35, the SASA of Asp52 in HEWL is almost identical for both protonation states. The lack of conformational fluctuation is due to the strong hydrogen bonding with the side-chain amino group of Asn46 and Asn59 (data not shown). Overestimation of the pKa shifts for buried residues can also be attributed to the limitation of the additive force field which underestimates dielectric response in protein environment (more discussion see Supporting Material) of the pKa shifts for buried residues.Finally, to ascertain if the presence of hydroxide/hydronium introduces artifacts, we studied the interaction between hydroxide/hydronium and the titratable sites/ions. Comparing the hydroxide/hydronium with respective chloride/sodium ions, we find that the spatial distributions are nearly identical (see plots of distance distributions and radial distribution functions in Figs. S10–S13). However, the relative occupancy of the hydroxide around the neutral Asp/Glu, positive histidine, or sodium ion is 2–3 times as that of a chloride. The water-bridged interaction between sodium and chloride ions becomes much weaker when chloride is replaced by hydroxide or sodium is replaced by hydronium. By contrast, the occupancy of the hydronium around the solute is similar to that of the sodium. Furthermore, similar pKa results for these proteins were obtained when coions were used instead of titratable waters (data not shown). Thus, we believe that potential artifacts related to the ionized forms of water are negligible. Work is underway to further understand the limitations of the methodology and to explore applications to protein dynamics coupled to proton transfer.In summary, we have developed and tested titratable water models for use in all-atom CpHMD simulations. Although the benchmark pKa calculations indicate a comparable accuracy as the GB-CpHMD method, the all-atom method offers physical rigor and most importantly, it is applicable to systems that cannot be studied with GB-based simulations such as lipids and nucleic acids. We anticipate that the accuracy of this methodology can be further improved by incorporating the new-generation force fields that account for polarization. The coupling between proton titration of water and solute offers a computationally efficient way to model proton transfer in molecular mechanics simulations.  相似文献   

    5.
    For a given graph G, ε(v) and deg(v) denote the eccentricity and the degree of the vertex v in G, respectively. The adjacent eccentric distance sum index of a graph G is defined as ξsv(G)=vV(G)ε(v)D(v)deg(v), where D(v)=uV(G)d(u,v) is the sum of all distances from the vertex v. In this paper we derive some bounds for the adjacent eccentric distance sum index in terms of some graph parameters, such as independence number, covering number, vertex connectivity, chromatic number, diameter and some other graph topological indices.  相似文献   

    6.
    7.
    Owing to their ability to break glycosidic bonds in recalcitrant crystalline polysaccharides such as cellulose, the catalysis effected by lytic polysaccharide monooxygenases (LPMOs) is of major interest. Kinetics of these reductant-dependent, monocopper enzymes is complicated by the insoluble nature of the cellulose substrate and parallel, enzyme-dependent, and enzyme-independent side reactions between the reductant and oxygen-containing cosubstrates. Here, we provide kinetic characterization of cellulose peroxygenase (oxidative cleavage of glycosidic bonds in cellulose) and reductant peroxidase (oxidation of the reductant) activities of the LPMO TrAA9A of the cellulose-degrading model fungus Trichoderma reesei. The catalytic efficiency (kcat/Km(H2O2)) of the cellulose peroxygenase reaction (kcat = 8.5 s−1, and Km(H2O2)=30μM) was an order of magnitude higher than that of the reductant (ascorbic acid) peroxidase reaction. The turnover of H2O2 in the ascorbic acid peroxidase reaction followed the ping-pong mechanism and led to irreversible inactivation of the enzyme with a probability of 0.0072. Using theoretical analysis, we suggest a relationship between the half-life of LPMO, the values of kinetic parameters, and the concentrations of the reactants.  相似文献   

    8.
    9.
    Several l-aminoacyl-tRNA synthetases can transfer a d-amino acid onto their cognate tRNA(s). This harmful reaction is counteracted by the enzyme d-aminoacyl-tRNA deacylase. Two distinct deacylases were already identified in bacteria (DTD1) and in archaea (DTD2), respectively. Evidence was given that DTD1 homologs also exist in nearly all eukaryotes, whereas DTD2 homologs occur in plants. On the other hand, several bacteria, including most cyanobacteria, lack genes encoding a DTD1 homolog. Here we show that Synechocystis sp. PCC6803 produces a third type of deacylase (DTD3). Inactivation of the corresponding gene (dtd3) renders the growth of Synechocystis sp. hypersensitive to the presence of d-tyrosine. Based on the available genomes, DTD3-like proteins are predicted to occur in all cyanobacteria. Moreover, one or several dtd3-like genes can be recognized in all cellular types, arguing in favor of the nearubiquity of an enzymatic function involved in the defense of translational systems against invasion by d-amino acids.Although they are detected in various living organisms (reviewed in Ref. 1), d-amino acids are thought not to be incorporated into proteins, because of the stereospecificity of aminoacyl-tRNA synthetases and of the translational machinery, including EF-Tu and the ribosome (2). However, the discrimination between l- and d-amino acids by aminoacyl-tRNA synthetases is not equal to 100%. Significant d-aminoacylation of their cognate tRNAs by Escherichia coli tyrosyl-, tryptophanyl-, aspartyl-, lysyl-, and histidyl-tRNA synthetases has been characterized in vitro (39). Recently, using a bacterium, transfer of d-tyrosine onto tRNATyr was shown to occur in vivo (10).With such misacylation reactions, the resulting d-aminoacyl-tRNAs form a pool of metabolically inactive molecules, at best. At worst, d-aminoacylated tRNAs infiltrate the protein synthesis machinery. Although the latter harmful possibility has not yet been firmly established, several cells were shown to possess a d-tyrosyl-tRNA deacylase, or DTD, that should help them counteract the accumulation of d-aminoacyl-tRNAs. This enzyme shows a broad specificity, being able to remove various d-aminoacyl moieties from the 3′-end of a tRNA (46, 11). Such a function makes the deacylase a member of the family of enzymes capable of editing in trans mis-aminoacylated tRNAs. This family includes several homologs of aminoacyl-tRNA synthetase editing domains (12), as well as peptidyl-tRNA hydrolase (13, 14).Two distinct deacylases have already been discovered. The first one, called DTD1, is predicted to occur in most bacteria and eukaryotes (see d-amino acids, including d-tyrosine (6). In fact, in an E. coli Δdtd strain grown in the presence of 2.4 mm d-tyrosine, as much as 40% of the cellular tRNATyr pool becomes esterified with d-tyrosine (10).

    TABLE 1

    Distribution of DTD1 and DTD2 homologs in various phylogenetic groupsHomologs of DTD1 and DTD2 were searched for using a genomic Blast analysis against complete genomes in the NCBI Database (www.ncbi.nlm.nih.gov). Values in the table are number of species. For instance, E. coli is counted only once in γ-proteobacteria despite the fact that several E. coli strains have been sequenced.
    DTD1DTD2DTD1 + DTD2None
    Bacteria
        Acidobacteria 2 0 0 0
        Actinobacteria 27 0 0 8
        Aquificae 1 0 0 0
        Bacteroidetes/Chlorobi 12 0 0 5
        Chlamydiae 1 0 0 6
        Chloroflexi 4 0 0 0
        Cyanobacteria 5 0 0 16
        Deinococcus/Thermus 4 0 0 0
        Firmicutes
            Bacillales 19 0 0 0
            Clostridia 19 0 0 0
            Lactobacillales 23 0 0 0
            Mollicutes 0 0 0 15
        Fusobacteria/Planctomycetes 2 0 0 0
        Proteobacteria
            α 6 0 0 55
            β 24 0 0 11
            γ 80 0 0 8
            δ 15 0 0 0
            ε 1 0 0 12
        Spirochaetes 0 0 0 7
        Thermotogae 5 0 0 0
    Archaea
        Crenarchaeota 0 13 0 0
        Euryarchaeota 1 26 0 2
        Nanoarchaeota 0 0 0 1
    Eukaryota
        Dictyosteliida 1 0 0 0
        Fungi/Metazoa
            Fungi 13 0 0 1
            Metazoa 19 0 0 0
        Kinetoplastida 3 0 0 0
        Viridiplantae 4 4 4 0
    Open in a separate windowHomologs of dtd/DTD1 are not found in the available archaeal genomes except that of Methanosphaera stadtmanae. A search for deacylase activity in Sulfolobus solfataricus and Pyrococcus abyssi led to the detection of another enzyme (DTD2), completely different from the DTD1 protein (15). Importing dtd2 into E. coli functionally compensates for dtd deprivation. As shown in 16).Several cells contain neither dtd nor dtd2 homologs (d-tyrosyl-tRNA deacylase (DTD3). This protein, encoded by dtd3, behaves as a metalloenzyme. Sensitivity of the growth of Synechocystis to external d-tyrosine is strongly exacerbated by the disruption of dtd3. Moreover, expression of the Synechocystis DTD3 in a Δdtd E. coli strain, from a plasmid, restores the resistance of the bacterium to d-tyrosine. Finally, using the available genomes, we examined the occurrence of DTD3 in the living world. The prevalence of DTD3-like proteins is surprisingly high. It suggests that the defense of protein synthesis against d-amino acids is universal.  相似文献   

    10.
    Presented here is the first report describing the detection of potentially diarrheal Vibrio parahaemolyticus strains isolated from cultured bivalves on the Mediterranean coast, providing data on the presence of both tdh- and trh-positive isolates. Potentially diarrheal V. parahaemolyticus strains were isolated from four species of bivalves collected from both bays of the Ebro delta, Spain.Gastroenteritis caused by Vibrio parahaemolyticus has been reported worldwide, though only sporadic cases have been reported in Europe (7, 14). The bacterium can be naturally present in seafood, but pathogenic isolates capable of inducing gastroenteritis in humans are rare in environmental samples (2 to 3%) (15) and are often not detected (10, 19, 20).The virulence of V. parahaemolyticus is based on the presence of a thermostable direct hemolysin (tdh) and/or the thermostable direct hemolysin-related gene (trh) (1, 5). Both are associated with gastrointestinal illnesses (2, 9).Spain is not only the second-largest producer in the world of live bivalve molluscs but also one of the largest consumers of bivalve molluscs, and Catalonia is the second-most important bivalve producer of the Spanish Autonomous Regions. Currently, the cultivation of bivalves in this area is concentrated in the delta region of the Ebro River. The risk of potentially pathogenic Vibrio spp. in products placed on the market is not assessed by existing legislative indices of food safety in the European Union, which emphasizes the need for a better knowledge of the prevalence of diarrheal vibrios in seafood products. The aim of this study was to investigate the distribution and pathogenic potential of V. parahaemolyticus in bivalve species exploited in the bays of the Ebro delta.Thirty animals of each species of Mytilus galloprovincialis, Crassostrea gigas, Ruditapes decussatus, and Ruditapes philippinarum were collected. They were sampled from six sites of the culture area, three in each bay of the Ebro River delta, at the beginning (40°37′112"N, 0°37′092"E [Alfacs]; 40°46′723"N, 0°43′943"E [Fangar]), middle (40°37′125"N, 0°38′570"E [Alfacs]; 40°46′666"N, 0°45′855"E [Fangar]), and end (40°37′309"N, 0°39′934"E [Alfacs]; 40°46′338"N, 0°44′941"E [Fangar]) of the culture polygon. Clams were sampled from only one site per bay as follows: in the Alfacs Bay from a natural bed of R. decussatus (40°37′44"N, 0°38′0"E) and in the Fangar Bay from an aquaculture bed of R. philippinarum (40°47′3"N, 0°43′8"E). In total, 367 samples were analyzed in 2006 (180 oysters, 127 mussels, 30 carpet shell clams, and 30 Manila clams) and 417 samples were analyzed in 2008 (178 oysters, 179 mussels, 30 carpet shell clams, and 30 Manila clams).All animals were individually processed and homogenized, and 1 ml of the homogenate was inoculated into 9 ml of alkaline peptone water (Scharlau, Spain). Following a 6-h incubation at 37°C, one loopful of the contents of each tube of alkaline peptone water was streaked onto CHROMagar vibrio plates (CHROMagar, France) and incubated for 18 h at 37°C. Mauve-purple colonies were purified, and each purified isolate was cryopreserved at −80°C (135 isolates in 2006 and 96 in 2008). From the initial homogenate portion, 100 μl was inoculated onto marine agar (Scharlau, Spain) and onto thiosulfate citrate-bile salts-sucrose agar (Scharlau, Spain) for total heterotrophic marine bacteria counts and total vibrio counts, respectively (Table (Table11).

    TABLE 1.

    Vibrio parahaemolyticus isolates, serotypes, and origins and total number of vibrios/heterotrophic bacteria contained in the bivalvea
    IsolateDate of collectionOrganism and site of originTemp (°C)Salinity (‰)Gene(s)SerotypeBacterial count using indicated medium (CFU ml−1)
    TCBS agarMarine agar
    I7458 August 2006Mg-F24.537tdhND1.5 × 1041.2 × 104
    I79314 August 2006Cg-A2535tdhND9.2 × 1028.5 × 103
    I80514 August 2006Cg-A2535tdhO2:KUT7.2 × 1029 × 103
    I80614 August 2006Cg-A2535tdh and trhO3:K331.9 × 1034.6 × 103
    I80914 August 2006Cg-A2535tdhO2:K288 × 1047.3 × 102
    I6784 July 2006Rd-A28.636tdhO2:K283.1 × 1052.5 × 105
    I6284 July 2006Rd-A28.636tdhO4:KUT2.9 × 1048.4 × 104
    I7758 August 2006Cg-A24.537tdhND4.21 × 1031.1 × 104
    I6914 July 2006Rd-A28.636trhO1:K322.2 × 1052.6 × 105
    I71227 July 2006Mg-A29.435.5trhO1:KUT8.6 × 1038.4 × 103
    I7658 August 2006Cg-F24.537trhO4:K341 × 104Uncountable
    I98022 July 2008Cg-A26.733.5tdhO1:K322.7 × 1041.3 × 104
    I98122 July 2008Cg-A26.733.5trhO1:KUT1 × 1042.2 × 104
    I99322 July 2008Cg-A26.733.5tdhO5:K173 × 1031.1 × 104
    I99429 July 2008Mg-A27.737trhO3:KUT3.4 × 1037 × 103
    I10315 August 2008Cg-F27.737tdhO5:KUT5.5 × 1043.3 × 104
    I10345 August 2008Cg-F27.737tdhO3:KUT8.7 × 1044 × 104
    I10405 August 2008Cg-F27.737tdhO3:KUT1.6 × 1043.2 × 104
    I10425 August 2008Cg-F27.737tdh and trhND2.8 ×1043 × 104
    I10505 August 2008Cg-F27.737tdhO1:KUT4.7 × 1047.3 × 104
    I106320 August 2008Mg-F25.936tdhO3:KUT7.9 ×1041.4 × 104
    I106520 August 2008Mg-F25.936tdhO2:KUT2.2 × 1031.2 × 104
    I106820 August 2008Mg-F25.936tdhO5:KUT2.6 × 1045.2 × 104
    I106920 August 2008Mg-F25.936tdhO3:KUT2.4 × 1035.3 × 104
    I107320 August 2008Mg-F25.936tdhO5:KUT2.3 × 1037.5 × 103
    I107420 August 2008Mg-F25.936tdhO3:KUT7.6 × 1046.9 × 104
    I107720 August 2008Mg-F25.936tdhO4:KUT1.7 × 1031.6 × 103
    I107920 August 2008Mg-F25.936trhO3:KUT2.5 × 1031.1 × 104
    I109220 August 2008Mg-F25.936tdhND1.7 × 1031.6 × 103
    I113025 August 2008Rd-A26.435tdhND1.7 × 1043.8 × 104
    I114325 August 2008Rd-A26.435tdhND1.1 × 1041.9 × 104
    I116525 August 2008Rd-A26.435trhO2:KUT4.4 × 1046.8 × 104
    I113325 August 2008Rp-F25.536.5tdhND3.4 × 1044 × 104
    I113425 August 2008Rp-F25.536.5tdhND3.9 × 1045.8 × 104
    I115825 August 2008Rp-F25.536.5trhO4:KUT6.6 × 1044.7 × 104
    I116125 August 2008Rp-F25.536.5trhO3:KUT2.2 × 1046.6 × 104
    Open in a separate windowaMg, Mytilus galloprovincialis; Cg, Crassostrea gigas; Rd, Ruditapes decussatus; Rp, R. phillipinarum; A, Alfacs; F, Fangar; ND, not determined; TCBS, thiosulfate citrate-bile salts-sucrose.Total DNA was extracted from each purified isolate using the Wizard genomic DNA purification kit (Promega), following the instructions of the manufacturer. A one-step PCR analysis was performed to identify/confirm which isolates were tl positive (species marker for V. parahaemolyticus). Further detection of the tdh or trh gene was carried out on all positive tl strains. All PCR analyses were carried out using the primers described by Bej et al. (2) with the following amplification conditions on the thermocycler (Eppendorf Mastercycler Personal): an initial denaturation at 95°C for 8 min, followed by 40 cycles of a 1-min denaturation at 94°C, annealing at 55°C for 1 min, elongation at 72° for 1 min, and a final extension of 10 min at 72°C. Positive and negative controls were included in all reaction mixtures: two positive controls, tl and tdh CAIM 1400 and trh CAIM 1772 (Collection of Aquatic Important Microorganisms [http://www.ciad.mx/caim/CAIM.html]), and negative control DNA-free molecular grade water (Sigma-Aldrich, Spain). Expected amplicons were visualized in 2% agarose gels stained with ethidium bromide.Fifty-eight isolates contained the gene tl in 2006 and 96 in 2008, which confirmed their identity as V. parahaemolyticus. In 2006, the distribution of the 58 isolates was as follows: 7 from 127 mussels, 34 from 180 oysters, and 17 from 30 R. decussatus clams. No tl-positive isolates were found in R. philippinarum. PCR analysis of the tl-positive isolates for the presence of the tdh or trh gene indicated that eight isolates contained the tdh gene and four contained the trh gene. In 2008, the source of the confirmed V. parahaemolyticus isolates was as follows: 31 from 88 oysters, 44 from 89 mussels, 9 from 30 R. decussatus clams, and 12 from 30 R. philippinarum clams. Of these, 17 were found to contain the tdh gene and 7 contained the trh gene. Two isolates (I806 and I1042) contained both toxigenic genes, tdh and trh.Putative tdh- and trh-positive PCR products were purified using the QIAquick PCR purification kit (Qiagen) following the manufacturer''s instructions and were sequenced bidirectionally by Macrogen Inc. Sequences were aligned using BioEdit (8) and analyzed using BLAST (National Center for Biotechnology Information). None of the toxigenic isolates was found positive by PCR analysis for the presence of open reading frame 8 of the phage 237 (16), a marker for the pandemic strain O3:K6.The isolates were fingerprinted by repetitive extragenic palindromic PCR (rep-PCR) as described previously (3), and the resulting electrophoretic band patterns were analyzed with the GelCompar II software (v4.5; Applied Maths). The similarity matrix was calculated with the Jaccard coefficient with a band position tolerance of 0.8%, and the dendrogram was constructed with the Ward algorithm. A high level of genomic diversity was found among the 32 toxigenic isolates characterized by rep-PCR. Three clonal groups were identified (those having identical rep-PCR band patterns) (Fig. 1a to c).Open in a separate windowFIG. 1.rep-PCR dendrogram of toxigenic isolates of V. parahaemolyticus isolated in the Ebro delta. Letters denote clonal groups of isolates.In vitro antibiotic susceptibility tests were performed using the diffusion disc test following a previously described protocol (18). The antibiotics used were gentamicin (10 μg), oxolinic acid (10 μg), amoxicillin (25 μg), polymyxin B (300 UI), vancomycin (30 μg), trimethoprim sulfamethoxazole (1.25/23.75 μg), nitrofurantoin (300 μg), doxycyclin (30 μg), ceftazidime (30 μg), streptomycin (10 μg), neomycin (30 UI), penicillin (6 μg), flumequine (30 μg), tetracycline (30 μg), ampicillin (10 μg), kanamycin (30 μg), ciprofloxacin (5 μg), and sulfonamide (300 μg). All tests were performed in duplicate. A Student t test for two samples with unequal variance was performed to compare the sensitivity of all 2006 isolates against the sensitivity of 2008 isolates for each antibiotic (Microsoft Office Excel 97-2003). Antibiogram results revealed a lower susceptibility in 2008 than in 2006, indicating a possible shift in overall susceptibility. Results from the t test indicated that significantly lower susceptibility in 2008 was detected (P ≤ 0.05; n = 36) for the following antibiotics: vancomycin, polymyxin B, ampicillin, amoxicillin, gentamicin, neomycin, trimethoprim sulfamethoxazole, nitrofurantoin, doxycyclin, ceftazidime, tetracycline, flumequine, and ciprofloxacin.The serological types for 27 strains were determined by the agglutination method using commercially available V. parahaemolyticus antisera (Denka Seiken Ltd.; Cosmos Biomedical Ltd, United Kingdom) following the manufacturer''s instructions. Potentially toxigenic V. parahaemolyticus isolates collected in 2006 were serologically heterogeneous (8 out of the 11 isolates) (Table (Table1).1). In isolates collected in 2008, results were more homogenous, with seven serotypes found among 19 isolates analyzed. The O3:K6 serotype was not detected in any of the strains analyzed, in agreement with the open reading frame 8 PCR results.The present study is the first to report the detection of potentially diarrheal V. parahaemolyticus strains isolated from cultured bivalves on Spanish Mediterranean coasts, providing data on the presence of both tdh- and trh-positive isolates. V. parahaemolyticus has previously been detected in several European countries (4, 13, 21, 22). A recent study carried out in Spain detected tdh-positive V. parahaemolyticus strains from patients who had consumed fresh oysters in a market in Galicia on the Atlantic coast of Spain (12) and potentially pathogenic V. parahaemolyticus strains have also been reported in France (17). These studies indicate that the risk of infections caused by V. parahaemolyticus in Europe is low compared to that in America or Asia (15). However, this risk could have been underestimated, since V. parahaemolyticus is not included in the current European surveillance programs, such as the European Network for Epidemiological Surveillance and Control of Communicable Diseases.Toxigenic V. parahaemolyticus strains detected in this study were genomically and serologically heterogeneous. The pandemic serotype O3:K6 was not detected, and although attempts to isolate O3:K6 from the environment and from seafood have not always been successful in previous studies reviewed by Nair and coauthors (15), this finding seems to be in agreement with the fact that no outbreak of diarrhea was observed in the area. Interestingly, isolates I806 and I1042 have been found positive for both tdh and trh in PCR tests. The coexistence of tdh and trh genes has already been reported in isolates from Japan, the United States, and Mexico (3, 6, 11, 19, 23). To our knowledge, no occurrence of an environmental isolate positive for both tdh and trh had previously been reported in Europe. All isolates tested were slightly different in their antibiotic resistance profiles. Typically, a high level of resistance could be determined. The detection of tdh- and/or trh-positive V. parahaemolyticus strains for the first time on the Mediterranean coast emphasizes the need to monitor for the presence of potentially diarrheal vibrios and bacterial gastroenteritis, and these data should be taken into consideration to revise the European legislation on the requirements for shellfish harvested for consumption in order to include the surveillance of these pathogens in Europe.  相似文献   

    11.
    12.
    Amylose-borate interaction has been analyzed by frontal gel chromatography, using the constituent velocity data alone. The constituent Velocity equation was reformulated in terms of elution volume for a type of interacting system described byA+iB=ABi(i=1,2,3n)

    Detailed examination of the binding data indicates that, in the complex formation between amylose and borate, this type of equilibria operates predominantly, if not solely. Use of the constituent elution volume equation enabled us, for the first time, to evaluate the association constant (K) and number of binding site pertaining to this system, i.e., K = 4.9 102 and n = 1. There was no evidence indicating the occurrence of the formation of inclusion complex.  相似文献   

    13.
    14.
    15.
    The tests of planktonic foraminifera recovered from deep-sea sediment are commonly observed to encapsulate fine grain carbonate sediment within their chambers. In sediment below the lysocline, the interstitial water within the chambers may not be as corrosive as the seawater in contact with the outer surface of the test due to slow continuous dissolution of the encapsulated carbonate. As a consequence, the pore walls of the foram dissolve more slowly than the outer surface of the test.Using published dissolution rate measurements for foraminifera and deep-sea sediment, the effect of diffusional reduction of pore wall dissolution was quantitatively estimated with a one-dimensional model for the steady state condition where diffusional flux out of the foraminifer's pores is balanced by the dissolution flux from the encapsulated fines and pore walls. The diffusional effect is found to principally depend on the structural parametric ratioTwdT/f, whereTw is the wall thickness,dT the test diameter andf is the test porosity. In the case of adult planktonic foraminifera, the ratio of the pore wall to outer surface dissolution flux is predicted to vary between 60% for the thin-walled porous species and 30% for thick-walled tests.Incorporation of the predicted pore to outer surface flux ratios into the morphologic index equation of Adelseck (1978) results in a very good prediction (73% of the variation) of the solution index of Berger (1975) obtained from ranking species counts from core tops. A simple empirical equation which may be useful for prediction of the resistance of extinct microfossils was found as follows:
    R=Twλ{1+[Aw/(As?Ap)][.74?.27log?(TwdT/f)]}
    is the measured ratio of pore wall area to outer surface of the test, andTwdT/f is in units of 104 μm2.  相似文献   

    16.
    With great interest, we read the article by Toms and colleagues [1] in the previous issue of Arthritis Research & Therapy, in which they assessed prevalences of metabolic syndrome (MetS) in rheumatoid arthritis (RA) patients. Moreover, they identified demographic and clinical factors that may be associated with MetS. Toms and colleagues found prevalences of up to 45% of MetS and demonstrated older age and health status (health assessment questionnaire) to be associated with MetS irrespectively of the definition used. Of most interest, an association between methotrexate (MTX) use and decreased presence of MetS was observed in patients more than 60 years of age. The investigators hypothesized that this may be attributed to a drug-specific effect (and not to an anti-inflammatory effect) either by changing levels of adenosine, which is known to interact with glucose and lipid metabolism, or by an indirect effect mediated through concomitant folic acid administration, thereby decreasing homocysteine levels.Recently, we also examined the prevalence of MetS in (a subgroup of) RA patients in the CARRÉ investigation, a prospective cohort study on prevalent and incident cardiovascular disease and its underlying cardiovascular risk factors [2]. The findings of Toms and colleagues stimulated us to perform additional analyses in our total study population (n = 353).The prevalences of MetS were 35% and 25% (Table (Table1)1) according to criteria of National Cholesterol Education Program (NCEP) 2004 and NCEP 2001, respectively. In multivariate backward regression analyses, we found significant associations between body mass index, pulse rate, creatinine levels, hypothyroidism and diabetes mellitus and the presence of MetS independently of the criteria used (Table (Table2).2). However, an independent association between single use of MTX or use of MTX in combination with other disease-modifying antirheumatic drugs, on the one hand, and a decreased prevalence of MetS, on the other hand, could not be demonstrated (even in the subgroup of patients over the age of 60).

    Table 1

    Characteristics of the study population
    MetS presentaMetS absentaMetS presentbMetS absentb
    n = 84n = 265n = 121n = 228P valueaP valueb
    Demographics
     Age, years63.8 (± 8)63.1 (± 7)64.3 (± 8)62.7 (± 7)0.460.045
     Female, percentage766374620.0220.028
    RA-related characteristics
     DAS284.2 (± 1.3)3.9 (± 1.4)4.1 (± 1.3)3.8 (± 1.4)0.210.062
     ESR, mm/hour22 (10-35)16 (9-30)20 (10-34)17 (9-31)0.0590.33
     CRP, mg/L11 (4-21)6 (3-16)8 (3-18)6 (3-19)0.0210.46
     RA duration, years7 (4-10)7 (4-10)7 (4-10)7 (5-10)0.830.19
     Erosion, percentage778379830.200.36
     Number of DMARDs1 (1-2)1 (1-1)1 (1-2)1 (1-1)0.260.43
     MTX current, percentage626063590.710.46
     MTX only, percentage393941380.950.67
     SSZ only, percentage8139140.230.22
     HCQ only, percentage14340.310.55
     Combination of DMARDs, percentage312529250.240.38
     TNF-blocking agent, percentage1191190.730.65
     Prednisolone only, percentage12311.000.42
    Cardiovascular risk factors
     Current smoker, percentage263125320.420.15
     Pack-years, years17 (0-34)19 (2-38)19 (0-35)18 (2-38)0.230.75
     BMI, kg/m230 (± 4)26 (± 5)29 (± 4)25 (± 5)< 0.001< 0.001
     Creatinine, μmol/L89 (± 21)89 (± 16)91 (± 22)87 (± 14)0.990.070
     Renal clearance, mL/minute81 (± 24)72 (± 19)77 (± 23)73 (± 19)0.0030.062
     Pulse, beats per minute76 (± 11)73 (± 9)75 (± 11)73 (± 9)0.0050.015
     Diabetes mellitus, percentage143123< 0.0010.001
     Hypothyroidism, percentage122920.0010.003
    Open in a separate windowaMetabolic syndrome (MetS) according to National Cholesterol Education Program (NCEP) 2001; bMetS according to NCEP 2004. Continuous variables are presented as means (± standard deviations) in cases of normal distribution or as medians (interquartile ranges) in cases of non-normal distribution. BMI, body mass index; CRP, C-reactive protein; DAS28, disease activity score using 28 joint counts; DMARD, disease-modifying antirheumatic drug; ESR, erythrocyte sedimentation rate; HCQ, hydroxychloroquine; MTX, methotrexate; RA, rheumatoid arthritis; SSZ, sulfasalazine; TNF, tumour necrosis factor.

    Table 2

    Variables associated with metabolic syndrome
    UnivariateMultivariatea


    OR95% CIP valueOR95% CIP value
    Body mass index1.21.1-1.3< 0.0011.21.1-1.3< 0.001
    Pulse1.031.01-1.060.0111.031.00-1.060.020
    Creatinine1.011.00-1.020.0801.021.00-1.030.017
    Hypothyroidism4.51.5-13.20.0074.71.5-15.00.009
    Diabetes mellitus4.81.8-12.90.0024.51.4-15.20.014
    Open in a separate windowaIn multivariate analyses, the following variables were used: gender, age, prednisolone only, methotrexate only, sulfasalazine only, hydroxychloroquine only, tumour necrosis factor-blocking agents, combination of disease-modifying antirheumatic drugs, pack-years, smoking, erosions, DAS28 (disease activity score using 28 joint counts), body mass index, pulse rate, creatinine levels, renal clearance, hypothyroidism and diabetes mellitus. CI, confidence interval; OR, odds ratio.Therefore, to get more support for a drug-specific effect, it is of interest to know whether or not in the study of Toms and colleagues the MTX effect was present only in the group of RA patients with single use of MTX or in the group of MTX-treated patients with other antirheumatic drugs. As patients with MetS were significantly older, it would give further information whether age was an independent risk factor for MetS in regression analyses. Moreover, as readers, we are not informed about comorbidities like diabetes and clinical hypothyroidism, which are notorious cardiometabolic risk factors. On the whole, we could not confirm a plausible protective role for the use of MTX and presence of MetS, and hence further investigation is required to explain the discrepancy between our findings and those of Toms and colleagues.  相似文献   

    17.
    Riboflavin significantly enhanced the efficacy of simulated solar disinfection (SODIS) at 150 watts per square meter (W m−2) against a variety of microorganisms, including Escherichia coli, Fusarium solani, Candida albicans, and Acanthamoeba polyphaga trophozoites (>3 to 4 log10 after 2 to 6 h; P < 0.001). With A. polyphaga cysts, the kill (3.5 log10 after 6 h) was obtained only in the presence of riboflavin and 250 W m−2 irradiance.Solar disinfection (SODIS) is an established and proven technique for the generation of safer drinking water (11). Water is collected into transparent plastic polyethylene terephthalate (PET) bottles and placed in direct sunlight for 6 to 8 h prior to consumption (14). The application of SODIS has been shown to be a simple and cost-effective method for reducing the incidence of gastrointestinal infection in communities where potable water is not available (2-4). Under laboratory conditions using simulated sunlight, SODIS has been shown to inactivate pathogenic bacteria, fungi, viruses, and protozoa (6, 12, 15). Although SODIS is not fully understood, it is believed to achieve microbial killing through a combination of DNA-damaging effects of ultraviolet (UV) radiation and thermal inactivation from solar heating (21).The combination of UVA radiation and riboflavin (vitamin B2) has recently been reported to have therapeutic application in the treatment of bacterial and fungal ocular pathogens (13, 17) and has also been proposed as a method for decontaminating donor blood products prior to transfusion (1). In the present study, we report that the addition of riboflavin significantly enhances the disinfectant efficacy of simulated SODIS against bacterial, fungal, and protozoan pathogens.Chemicals and media were obtained from Sigma (Dorset, United Kingdom), Oxoid (Basingstoke, United Kingdom), and BD (Oxford, United Kingdom). Pseudomonas aeruginosa (ATCC 9027), Staphylococcus aureus (ATCC 6538), Bacillus subtilis (ATCC 6633), Candida albicans (ATCC 10231), and Fusarium solani (ATCC 36031) were obtained from ATCC (through LGC Standards, United Kingdom). Escherichia coli (JM101) was obtained in house, and the Legionella pneumophila strain used was a recent environmental isolate.B. subtilis spores were produced from culture on a previously published defined sporulation medium (19). L. pneumophila was grown on buffered charcoal-yeast extract agar (5). All other bacteria were cultured on tryptone soy agar, and C. albicans was cultured on Sabouraud dextrose agar as described previously (9). Fusarium solani was cultured on potato dextrose agar, and conidia were prepared as reported previously (7). Acanthamoeba polyphaga (Ros) was isolated from an unpublished keratitis case at Moorfields Eye Hospital, London, United Kingdom, in 1991. Trophozoites were maintained and cysts prepared as described previously (8, 18).Assays were conducted in transparent 12-well tissue culture microtiter plates with UV-transparent lids (Helena Biosciences, United Kingdom). Test organisms (1 × 106/ml) were suspended in 3 ml of one-quarter-strength Ringer''s solution or natural freshwater (as pretreated water from a reservoir in United Kingdom) with or without riboflavin (250 μM). The plates were exposed to simulated sunlight at an optical output irradiance of 150 watts per square meter (W m−2) delivered from an HPR125 W quartz mercury arc lamp (Philips, Guildford, United Kingdom). Optical irradiances were measured using a calibrated broadband optical power meter (Melles Griot, Netherlands). Test plates were maintained at 30°C by partial submersion in a water bath.At timed intervals for bacteria and fungi, the aliquots were plated out by using a WASP spiral plater and colonies subsequently counted by using a ProtoCOL automated colony counter (Don Whitley, West Yorkshire, United Kingdom). Acanthamoeba trophozoite and cyst viabilities were determined as described previously (6). Statistical analysis was performed using a one-way analysis of variance (ANOVA) of data from triplicate experiments via the InStat statistical software package (GraphPad, La Jolla, CA).The efficacies of simulated sunlight at an optical output irradiance of 150 W m−2 alone (SODIS) and in the presence of 250 μM riboflavin (SODIS-R) against the test organisms are shown in Table Table1.1. With the exception of B. subtilis spores and A. polyphaga cysts, SODIS-R resulted in a significant increase in microbial killing compared to SODIS alone (P < 0.001). In most instances, SODIS-R achieved total inactivation by 2 h, compared to 6 h for SODIS alone (Table (Table1).1). For F. solani, C. albicans, ands A. polyphaga trophozoites, only SODIS-R achieved a complete organism kill after 4 to 6 h (P < 0.001). All control experiments in which the experiments were protected from the light source showed no reduction in organism viability over the time course (results not shown).

    TABLE 1.

    Efficacies of simulated SODIS for 6 h alone and with 250 μM riboflavin (SODIS-R)
    OrganismConditionaLog10 reduction in viability at indicated h of exposureb
    1246
    E. coliSODIS0.0 ± 0.00.2 ± 0.15.7 ± 0.05.7 ± 0.0
    SODIS-R1.1 ± 0.05.7 ± 0.05.7 ± 0.05.7 ± 0.0
    L. pneumophilaSODIS0.7 ± 0.21.3 ± 0.34.8 ± 0.24.8 ± 0.2
    SODIS-R4.4 ± 0.04.4 ± 0.04.4 ± 0.04.4 ± 0.0
    P. aeruginosaSODIS0.7 ± 0.01.8 ± 0.04.9 ± 0.04.9 ± 0.0
    SODIS-R5.0 ± 0.05.0 ± 0.05.0 ± 0.05.0 ± 0.0
    S. aureusSODIS0.0 ± 0.00.0 ± 0.06.2 ± 0.06.2 ± 0.0
    SODIS-R0.2 ± 0.16.3 ± 0.06.3 ± 0.06.3 ± 0.0
    C. albicansSODIS0.2 ± 0.00.4 ± 0.10.5 ± 0.11.0 ± 0.1
    SODIS-R0.1 ± 0.00.7 ± 0.15.3 ± 0.05.3 ± 0.0
    F. solani conidiaSODIS0.2 ± 0.10.3 ± 0.00.2 ± 0.00.7 ± 0.1
    SODIS-R0.3 ± 0.10.8 ± 0.11.3 ± 0.14.4 ± 0.0
    B. subtilis sporesSODIS0.3 ± 0.00.2 ± 0.00.0 ± 0.00.1 ± 0.0
    SODIS-R0.1 ± 0.10.2 ± 0.10.3 ± 0.30.1 ± 0.0
    SODIS (250 W m−2)0.1 ± 0.00.1 ± 0.10.1 ± 0.10.0 ± 0.0
    SODIS-R (250 W m−2)0.0 ± 0.00.0 ± 0.00.2 ± 0.00.4 ± 0.0
    SODIS (320 W m−2)0.1 ± 0.10.1 ± 0.00.0 ± 0.14.3 ± 0.0
    SODIS-R (320 W m−2)0.1 ± 0.00.1 ± 0.10.9 ± 0.04.3 ± 0.0
    A. polyphaga trophozoitesSODIS0.4 ± 0.20.6 ± 0.10.6 ± 0.20.4 ± 0.1
    SODIS-R0.3 ± 0.11.3 ± 0.12.3 ± 0.43.1 ± 0.2
    SODIS, naturalc0.3 ± 0.10.4 ± 0.10.5 ± 0.20.3 ± 0.2
    SODIS-R, naturalc0.2 ± 0.11.0 ± 0.22.2 ± 0.32.9 ± 0.3
    A. polyphaga cystsSODIS0.4 ± 0.10.1 ± 0.30.3 ± 0.10.4 ± 0.2
    SODIS-R0.4 ± 0.20.3 ± 0.20.5 ± 0.10.8 ± 0.3
    SODIS (250 W m−2)0.0 ± 0.10.2 ± 0.30.2 ± 0.10.1 ± 0.2
    SODIS-R (250 W m−2)0.4 ± 0.20.3 ± 0.20.8 ± 0.13.5 ± 0.3
    SODIS (250 W m−2), naturalc0.0 ± 0.30.2 ± 0.10.1 ± 0.10.2 ± 0.1
    SODIS-R (250 W m−2), naturalc0.1 ± 0.10.2 ± 0.20.6 ± 0.13.4 ± 0.2
    Open in a separate windowaConditions are at an intensity of 150 W m−2 unless otherwise indicated.bThe values reported are means ± standard errors of the means from triplicate experiments.cAdditional experiments for this condition were performed using natural freshwater.The highly resistant A. polyphaga cysts and B. subtilis spores were unaffected by SODIS or SODIS-R at an optical irradiance of 150 W m−2. However, a significant reduction in cyst viability was observed at 6 h when the optical irradiance was increased to 250 W m−2 for SODIS-R only (P < 0.001; Table Table1).1). For spores, a kill was obtained only at 320 W m−2 after 6-h exposure, and no difference between SODIS and SODIS-R was observed (Table (Table1).1). Previously, we reported a >2-log kill at 6 h for Acanthamoeba cysts by using SODIS at the higher optical irradiance of 850 W m−2, compared to the 0.1-log10 kill observed here using the lower intensity of 250 W m−2 or the 3.5-log10 kill with SODIS-R.Inactivation experiments performed with Acanthamoeba cysts and trophozoites suspended in natural freshwater gave results comparable to those obtained with Ringer''s solution (P > 0.05; Table Table1).1). However, it is acknowledged that the findings of this study are based on laboratory-grade water and freshwater and that differences in water quality through changes in turbidity, pH, and mineral composition may significantly affect the performance of SODIS (20). Accordingly, further studies are indicated to evaluate the enhanced efficacy of SODIS-R by using natural waters of varying composition in the areas where SODIS is to be employed.Previous studies with SODIS under laboratory conditions have employed lamps delivering an optical irradiance of 850 W m−2 to reflect typical natural sunlight conditions (6, 11, 12, 15, 16). Here, we used an optical irradiance of 150 to 320 W m−2 to obtain slower organism inactivation and, hence, determine the potential enhancing effect of riboflavin on SODIS.In conclusion, this study has shown that the addition of riboflavin significantly enhances the efficacy of simulated SODIS against a range of microorganisms. The precise mechanism by which photoactivated riboflavin enhances antimicrobial activity is unknown, but studies have indicated that the process may be due, in part, to the generation of singlet oxygen, H2O2, superoxide, and hydroxyl free radicals (10). Further studies are warranted to assess the potential benefits from riboflavin-enhanced SODIS in reducing the incidence of gastrointestinal infection in communities where potable water is not available.  相似文献   

    18.
    Some physicochemical properties and amino acid composition of alkaline proteinase from Aspergillus sojae were found to be as follows: The isoelectric point was at pH 5.1. The molecular weight was 25,500 using the Sheraga-Mandelkern’s formula, based upon the values of the sedimentation coefficient (s20,w°=?2.82?S), the intrinsic viscosity ([η] = 0.027 dl/g), and the partial specific volume (V¯?=?0.726?ml/g). The enzyme contains 16.8% of nitrogen and is composed of 250 residues of amino acid; Asp31 Glu19, Gly27, Ala32, Val18, Leu14, Ile14, Ser28, Thr18, (Cys C?ys)1, Met2, Pro6, Phe7, Tyr8, Trp2, His5, Lys14, Arg3, (amide-NH3)20.  相似文献   

    19.
    Kinase cascades, in which enzymes are sequentially activated by phosphorylation, are quintessential signaling pathways. Signal transduction is not always achieved by direct activation, however. Often, kinases activate pathways by deactivation of a negative regulator; this indirect mechanism, pervasive in Akt signaling, has yet to be systematically explored. Here, we show that the indirect mechanism has properties that are distinct from direct activation. With comparable parameters, the indirect mechanism yields a broader range of sensitivity to the input, beyond saturation of regulator phosphorylation, and kinetics that become progressively slower, not faster, with increasing input strength. These properties can be integrated in network motifs to produce desired responses, as in the case of feedforward loops.Phosphorylation of proteins and lipids, catalyzed by specific kinase enzymes, is ubiquitous in intracellular signal transduction. A classic example in eukaryotes is the canonical structure of the mitogen-activated protein kinase cascades, in which three kinases are sequentially activated by phosphorylation (1). Another example is the PI3K (phosphoinositide 3-kinase)/Akt pathway, which (like the mammalian mitogen-activated protein kinases) is prominently dysregulated in human cancers (2). Type-I PI3Ks phosphorylate a lipid substrate to produce the lipid second messenger, PIP3, which recruits the protein kinase Akt and mediates its activation by phosphorylation (3,4). In no small part because of these important pathways, we typically think of phosphorylation as a direct means of activating molecular interactions and reactions in signal transduction. This is not the only way to increase the flux through a signaling pathway, however. Consider signaling downstream of Akt, which phosphorylates a host of protein substrates to affect diverse functions. A survey of the Akt signaling hub shows that many of these reactions result in a decrease, rather than an increase, in activity/function of the substrates (3). And, among those substrates, the four listed in Fig. S1 in the Supporting Material). Whereas negative regulators are appreciated for their roles in feedback adaptation of signaling, the implications of deactivating a negative regulator as an indirect mechanism of pathway activation has yet to be explored.

    Table 1

    Survey of Akt substrates and downstream signaling
    Substrate (site)Effect on substrateOutcome
    TSC2 (T1462)GAP activity ↓Rheb, mTOR ↑
    PRAS40 (T246)mTOR binding ↓mTOR ↑
    GSK3α/β (S21/S9)kinase activity ↓β-catenin ↑
    BAD (S136)Bcl-2/xL binding ↓Bcl-2/xL ↑
    Open in a separate windowHere, we use simple kinetic models to elucidate the basic properties of pathway activation by deactivation of a negative regulator (hereafter referred to as mechanism II), as compared with the standard activation of a positive regulator (mechanism I). The analysis is presented in the context of protein phosphorylation, but the conclusions may be generalized to other reversible modifications or to allosteric binding interactions. The common first step is phosphorylation of the regulatory molecule by the kinase. The activity of the upstream kinase such as Akt may be represented by a dimensionless, time (t)-dependent input signal function, s(t). We assume that the total amount of regulator is constant and define its phosphorylated fraction as ϕ(t). Neglecting concentration gradients and saturation of the upstream kinase and of the opposing (constitutively active) phosphatase(s), the conservation of phosphorylated regulator is expressed as follows (see Text S1 in the Supporting Material):dϕdt=kp[s(1ϕ)ϕ];ϕ(0)=0.(1)The parameter kp is the pseudo-first-order rate constant of protein dephosphorylation. In the case of s = constant (i.e., subject to a step change at t = 0), the properties of this simplified kinetic equation are well known (5) and may be summarized as follows. As the magnitude of the signal strength s increases, the steady-state value of ϕ, ϕss, increases in a saturable fashion; when s >> 1, ϕss approaches its maximum value of 1 and is insensitive to further increases in s. The kinetics of ϕ(t) approaching ϕss become progressively faster as s increases, however.Next, we model the influence of the regulator on a downstream response. Defining the fractional response as ρ and following analogous assumptions as above, we formulate equations for mechanisms I and II as follows:dρdt={[ka,0+(ka,maxka,0)ϕ](1ρ)kd,0ρ(I)ka,0(1ρ)[kd,0(kd,0kd,min)ϕ]ρ(II).(2)In each equation, the first term on the right-hand side describes activation, and the second, deactivation. In mechanism I, the effective rate constant of activation increases linearly with ϕ, from a minimum value of ka,0 when ϕ = 0 up to a maximum value of ka,max when ϕ = 1; the deactivation rate constant is fixed at kd,0. Conversely, in mechanism II, the effective rate constant of deactivation decreases linearly with ϕ, from a maximum value of kd,0 when ϕ = 0 down to a minimum value of kd,min when ϕ = 1; in this mechanism, the activation rate constant is fixed at ka,0. The initial condition is assigned so that ρ is stationary when ϕ = 0. To further set the two mechanisms on a common basis, we define dimensionless parameters such that the maximum steady-state value of ρ (with ϕss = 1) is the same for both mechanisms I and II,gka,max/ka,0kd,0/kd,minKka,0/kd,0.(3)With these definitions, each conservation equation is reduced to the following dimensionless form:1kd,0dρdt={K[1+(g1)ϕ](1ρ)ρ(I)K(1ρ)[1(1g1)ϕ]ρ(II).(4)Mechanisms I and II (Fig. 1 a) are compared first at the level of their steady-state solutions, ρss, for stationary s. Equation 1 yields the familiar hyperbolic dependence of ϕss on s, and ρss(s) has the same shape for both mechanisms. However, whereas ρss of mechanism I shows saturation at a lower value of s than ϕss, the opposite is true of mechanism II (Fig. 1 b). Thus, mechanism II retains sensitivity to the input even while phosphorylation of the upstream regulator shows saturation. This is perhaps more readily seen when ϕss(s) is replaced with a sigmoidal Hill function (i.e., with s replaced by sn in Eq. 1) (Fig. 1 c). The key parameter that affects the relative sensitivities of mechanisms I and II and the disparity between them is the gain constant, g (see Text S1 in the Supporting Material). As this parameter is increased, ρss of mechanism I becomes increasingly saturable with respect to ϕss (Fig. 1 d), whereas ρss of mechanism II gains sensitivity as ϕss approaches 1 (Fig. 1 e). As an illustrative example, consider that when ϕss is increased from 0.90 to 0.95, or from 0.98 to 0.99, the amount of the negative regulator in the active state is reduced by a factor of 2 (see Fig. S2).Open in a separate windowFigure 1Steady-state properties of mechanisms I and II. (a) Schematics of direct (I) and indirect (II) activation. (b) Steady-state dose responses, ρss(s), of mechanisms I and II along with phosphorylation of the upstream regulator, ϕss(s) (Eq. 1 at steady state); K = 0.05, g = 100. (c) Same as panel b, except with a sigmoidal ϕss(s) (Hill function with n = 4). (d) Steady-state output, ρss, of mechanism I vs. ϕss for K = 0.05 and indicated values of the gain constant, g. (e) Same as panel d, but for mechanism II. To see this figure in color, go online.The two mechanisms also show distinct temporal responses. In the response of mechanism I to a step increase in s, ρ(t) approaches ρss with a timescale that generally becomes faster as s increases. Unless the kinetics of ϕ(t) are rate-limiting, the timescale is ∼kd,0–1(1–ρss) (Fig. 2 a; see also Text S1 and Fig. S3 in the Supporting Material). Conversely, the response of mechanism II generally becomes slower as s increases, inasmuch as the frequency of deactivation decreases whereas that of activation is constant, with a timescale of ∼ka,0–1ρss (Fig. 2 b). To approximate a transient input, we model s(t) as a step increase followed by a decay. For mechanism I, the response ρ(t) is such that the variation in the time of the peak, as a function of the step size, is modest. The subsequent decay is prolonged when ϕ(t) hovers close to saturation (Fig. 2 c). Such kinetic schemes have been analyzed in some detail previously (6,7). In contrast, the response of mechanism II to the transient input is such that the system retains sensitivity and consistent decay kinetics beyond the saturation of ϕ(t). The distinctive feature is that ρ(t) peaks noticeably later in time as the magnitude of the peak increases (Fig. 2 d).Open in a separate windowFigure 2Kinetic properties of mechanisms I and II. (a) Response of mechanism I to a step change in s from zero to the indicated s(0). Time is given in units of kpt; parameters are K = 0.05, g = 10, and kd,0 = 0.1kp. (b) Same as panel a, but for mechanism II. (c) Same as panel a, but for a transient input, s(t) = s(0)exp(–0.03kpt). d) Same as panel c, but for mechanism II. To see this figure in color, go online.Having established the basic steady state and kinetic properties of mechanism II as compared with the canonical mechanism I, we considered what outcomes could be achieved by linking these motifs in series or in parallel. Such schemes are identified in the Akt/mTOR signaling network, for example (see Fig. S4). In a standard kinase activation cascade, it is understood that the properties of saturation and sensitivity are compounded with each step of the cascade (8). Thus, two sequential steps of mechanism I yield progressive saturation of the steady-state output at lower s (Fig. 3 a), and the desaturating effect of mechanism II is likewise compounded (Fig. 3 b). By corollary it follows that a sequence of mechanisms I and II will show an intermediate dose response; that is, the mechanism II step offsets the saturation effect of mechanism I.Open in a separate windowFigure 3Serial and parallel schemes incorporating mechanism I or/and II. (a) Steady-state outputs of two response elements, ρ1 and ρ2, activated by mechanism I in series. At each level, K = 0.05, g = 100. (b) Same as panel a, but for mechanism II in series. (c) Incoherent feedforward loop (FFL) in which mechanisms I and II are activated in parallel to activate and inhibit, respectively, the terminal output. For both mechanisms I and II, K = 0.05, g = 100. The parameters for Eq. 5 are α = 2.5, β = 50. To see this figure in color, go online.A more complex scheme is to combine the two mechanisms in parallel, as in an incoherent feedforward loop (FFL) connected to an “AND NOT” output as follows:Output = αρI/(1 + αρIβρII).(5)Given the differential saturation properties of mechanisms I and II, this scheme readily yields the expected biphasic dose response (9) without the need for disparate values of the parameters (Fig. 3 c). Regarding the kinetics, the analysis shown in Fig. 2 makes it clear that mechanism II naturally introduces time delays in cascades or network motifs. Thus, for the incoherent FFL at high, constant s, activation of inhibition by mechanism II would tend to yield a dynamic response marked by a peak followed by adaptation (see Fig. S5). Analogous calculations were carried out for a coherent FFL as well (see Fig. S6).To summarize our conclusions and their implications for signaling downstream of Akt and other kinases, we have described a distinct, indirect signal transduction mechanism characterized by deactivation of a negative regulator. This motif shows steady-state sensitivity beyond saturation, and therefore the activity of the upstream kinase, such as Akt, can be relatively high. By comparison, the direct activation of signaling by phosphorylation requires that activity of the kinase be regulated, or specifically countered by high phosphatase activity, to maintain sensitivity and avoid saturation of the response. The mechanism described here also introduces relatively slow kinetics (for comparable parameter values). This property, together with its extended range of sensitivity, would allow the motif to be incorporated in signaling networks to yield desired steady and unsteady responses in a robust manner. Considering that key signaling processes mediated by Akt (notably activation of the mammalian target of rapamycin (mTOR) pathway) are achieved by deactivation of negative regulators, we assert that greater recognition of this mechanism and of its distinct properties is warranted.  相似文献   

    20.
    设为首页 | 免责声明 | 关于勤云 | 加入收藏

    Copyright©北京勤云科技发展有限公司  京ICP备09084417号