Next Article in Journal
Synthesis, Complexation Properties, and Evaluation of New Aminodiphosphonic Acids as Vector Molecules for 68Ga Radiopharmaceuticals
Next Article in Special Issue
The Interaction of Human Glutathione Transferase GSTA1-1 with Reactive Dyes
Previous Article in Journal
Macrocystis pyrifera Extract Residual as Nutrient Source for the Production of Sophorolipids Compounds by Marine Yeast Rhodotorula rubra
Previous Article in Special Issue
Boronic Acids as Prospective Inhibitors of Metallo-β-Lactamases: Efficient Chemical Reaction in the Enzymatic Active Site Revealed by Molecular Modeling
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Good-Practice Non-Radioactive Assays of Inorganic Pyrophosphatase Activities

by
Alexander A. Baykov
1,*,
Viktor A. Anashkin
1 and
Anssi M. Malinen
2,*
1
Belozersky Institute of Physico-Chemical Biology, Lomonosov Moscow State University, 119899 Moscow, Russia
2
Department of Life Technologies, University of Turku, FIN-20014 Turku, Finland
*
Authors to whom correspondence should be addressed.
Molecules 2021, 26(8), 2356; https://doi.org/10.3390/molecules26082356
Submission received: 6 April 2021 / Revised: 15 April 2021 / Accepted: 16 April 2021 / Published: 18 April 2021
(This article belongs to the Special Issue Study on Enzyme Kinetics for Biochemistry)

Abstract

:
Inorganic pyrophosphatase (PPase) is a ubiquitous enzyme that converts pyrophosphate (PPi) to phosphate and, in this way, controls numerous biosynthetic reactions that produce PPi as a byproduct. PPase activity is generally assayed by measuring the product of the hydrolysis reaction, phosphate. This reaction is reversible, allowing PPi synthesis measurements and making PPase an excellent model enzyme for the study of phosphoanhydride bond formation. Here we summarize our long-time experience in measuring PPase activity and overview three types of the assay that are found most useful for (a) low-substrate continuous monitoring of PPi hydrolysis, (b) continuous and fixed-time measurements of PPi synthesis, and (c) high-throughput procedure for screening purposes. The assays are based on the color reactions between phosphomolybdic acid and triphenylmethane dyes or use a coupled ATP sulfurylase/luciferase enzyme assay. We also provide procedures to estimate initial velocity from the product formation curve and calculate the assay medium’s composition, whose components are involved in multiple equilibria.

1. Introduction

Inorganic pyrophosphatase (inorganic diphosphatase; EC 3.6.1.1; PPase) is a constitutive, highly specific enzyme that converts pyrophosphate (PPi), a byproduct and regulator of numerous biosynthetic reactions [1], into a metabolizable phosphate. Ubiquitous soluble PPases dissipate PPi energy as heat, whereas less common membrane-bound PPases use the energy to transport H+ or Na+ across lipid membranes [2,3]. Both PPase types can also catalyze the reverse reaction of PPi synthesis from Pi, and this activity of membrane-bound PPases is also considered physiologically important [4,5]. Soluble PPases are additionally divided into two nonhomologous families—family I, known for nearly a hundred years and found in all kingdoms of life, and family II found in 1998 in prokaryotes [6,7]. Several nonspecific phosphatases also exhibit PPase-like activity [8,9,10].
Although the primary structures of the three groups of specific PPases are completely different, their mechanisms and active sites reveal close similarity, making these enzymes remarkable examples of convergent evolution. All PPases are Mg2+-dependent enzymes, but family II PPases additionally require a transition metal ion (Mn2+ or Co2+) for maximal activity [11]. In total, three to four metal ions per active site are required for PPi conversion [7]. These metal ions, coordinated by numerous protein carboxylates and substrate phosphates, play key roles in catalysis. PPi hydrolysis involves a direct attack of an activated water molecule on a phosphorus atom and a stepwise release of two phosphate molecules. The value of kcat for soluble PPases reaches 104 s−1 for PPi hydrolysis and 102 s−1 for PPi synthesis [3,4,11]. Km values for most PPases with PPi as substrate lie in the micromolar range.
PPase is generally assayed by measuring phosphate, the product of PPi hydrolysis. Depending on the task, PPase activity is assayed at sub-Km PPi concentrations (in mechanistic studies) or a saturating PPi concentration (in high-throughput screening procedures). The former assay is more demanding concerning sensitivity because of the low Km value, especially with family I and membrane PPases. Sensitivity is also the limiting factor in PPi synthesis studies because the equilibrium PPi ⇆ 2Pi is largely shifted to the right. Here, we summarize our experience in sensitive PPase activity measurements in both directions and present protocols suitable for high-throughput screening and mechanistic studies.

2. Continuous Low-Substrate Assay of PPi-Hydrolyzing Activity

The low value of the Michaelis constant of PPases makes it necessary to use highly sensitive assays in kinetic studies conducted at sub-Km substrate concentrations. The assay currently used in our labs combines two approaches to achieve high sensitivity: first, it measures the formation of the intensively colored complex of 12-molibdophosphoric acid with a triphenylmethane dye, methyl green, and, second, it is continuous. The PPase reaction mixture and two-color reagents are continuously pumped off by a peristaltic pump and mixed, and the absorbance of the resulting solution is measured at 620 or 660 nm in a flow photometer. Even though the background absorbance is relatively high, it is automatically subtracted in this continuous assay mode, allowing the precise recording of much smaller changes in absorbance (Pi concentration) compared to a fixed-time assay. The principal advantage of the methyl green dye in this application is its low tendency to deposit on the optical cuvette [12].
The phosphate analyzer consists of three main parts: a four-channel peristaltic pump, flow photometer, and paper recorder (Figure 1A). Alternatively, the photometer’s output can be directed to a computer, which calculates initial velocity from the absorbance time-course. However, we found that the manual procedure, using paper recordings, is less sensitive to signal fluctuations and provides more accurate data. We have used a four-channel Gilson Minipuls pump equipped with Tygon tubings. Three pump channels are used to deliver a sample, acid/molybdate, and dye/Triton X-305 solutions. Their initial mixing occurs in tubing T-joints and is completed in a mixing device consisting of a glass tube of 1.3 mm inner diameter with 12–15 bubbles of 3.5 mm inner diameter. The fourth pump channel eliminates any air bubbles from the final mixed solution before it is directed to the flow photometer. To prevent excessive air from entering the flow system, the pump is stopped when the sample inlet tubing is transferred between the samples and water. To ensure that the color reaction proceeds to completion, the time intervals between the two mixing events and between the second mixing event and entry to the photometer cuvette should be adjusted to 12 and 32 s, respectively, by using connecting tubings of appropriate diameter and length. The small nonlinearity of the calibration plot at low phosphate content is eliminated by adding a small amount of phosphate to the stock acid/molybdate solution. We have been using ISCO UA-5 and 229 flow photometers. These instruments are no longer produced but can be replaced by any flow photometer operating at 620–660 nm wavelengths and a flow rate of approximately 10 mL/min. Alternatively, one can use a standard spectrophotometer with a flow cuvette, such as a Helma model 178-010-10-40 (10 mm pathlength, 80 µL internal volume).
Three setups of the phosphate analyzer were found useful for standard, high-sensitivity, and low dead-time measurements (Figure 1A–C). A five-fold increase in sensitivity was achieved in version B by interchanging the sample and molybdate tubings on the pump (Figure 1B). As the response is linear up to approximately 0.5 absorbance unit, the photometer sensitivity is adjusted to 0.5 absorbance unit per recorder scale in the standard mode (version A) and down to 0.1 unit in the high-sensitivity mode (version B). These settings correspond to approximately 100 and 4 µM Pi, respectively, per recorder scale. As each PPi molecule yields 2 molecules of Pi, this sensitivity is sufficient for a reliable measurement of initial velocities at PPi concentrations down to 0.5 µM (see Appendix A for further details). PPase concentration in the assay is typically in the pM range when activity is determined at saturating substrate concentrations and optimal pH and temperature values. We have successfully used the phosphate analyzer to determine PPase reaction kinetics over a broad temperature range (reaction mixture temperature 20–60 °C).
The dead-time between withdrawing the sample and mixing it with acid/molybdate is 10 s in the standard mode but can be decreased to 1 s by shifting the first mixing point to the pump inlet (Figure 1C). In this version, the sample is withdrawn from the reaction vessel at a rate of 1.2 mL/min due to differences in the two lowest pump tubes’ flow rates. This setup helps monitor reactions demonstrating nonlinear progress curves because of enzyme activation or inactivation during the reaction. Reagent concentrations were adjusted in versions B and C as indicated in Table 1 to ensure the same optimal final concentrations in the photometer cuvette.
The continuous Pi assay is robust and tolerates the presence of many biochemical compounds, including magnesium ions at high concentrations (up to 40 mM). In contrast, Mg2+ interferes with the Pi assay based on 12-molibdophosphoric acid reduction [17]. Nevertheless, it is wise to calibrate the instrument by measuring phosphate standard against background each time when a new component is added to the sample assayed, especially when working in the high-sensitivity mode (version B). Importantly, PPi at high concentrations was found to suppress the Pi signal (by 25% for 1.5 mM PPi) in this mode. This should be taken into consideration in measurements of PPase activity at a variable substrate concentration. The effect is proportional to PPi concentration and likely results from the competitive formation of 12-molibdopyrophosphoric acid, detected by Raman spectroscopy [18].
Other known continuous assays of PPase activity monitor, depending on pH, uptake or release of hydrogen ions during PPi conversion to Pi [19,20] or monitor PPi concentration using synthetic PPi sensors [21,22,23,24,25]. These assays are less sensitive and less convenient in that the signal is not proportional to the degree of conversion.
Substances that modulate PPase activity will interfere with any assay. Among them, Tris and other amine buffers were found to dramatically increase the Michaelis constant for family I and membrane PPases [26,27,28]. Accordingly, zwitterionic buffers, such as Tes, Mops, and HEPES, are preferable for work at sub-Km substrate concentrations. Tetramethylammonium hydroxide was found to be useful as the second component of the zwitterionic buffers in experiments measuring K+ and Na+ effects on K+-dependent membrane PPase. However, some batches of this strongly basic compound supplied in glass containers were found to nonspecifically inactivate this enzyme when used as a buffer component instead of KOH.

3. Sensitive Fixed-Time Assay of PPi-Hydrolyzing Activity

Fixed-time assays measuring the product Pi are perhaps most suitable for high-throughput screening of compound libraries for the effects on PPase activity. Most published procedures determine Pi as reduced 12-molybdophosphoric acid [29] and perform well with 10−5–10−4 M Pi concentrations. Although this sensitivity suffices for most applications, the identification of samples with low PPase activity or assays performed at low PPi concentrations (<10 µM) would require a more sensitive Pi assay, such as the assay based on the shift of malachite green spectrum upon binding to 12-molybdophosphoric acid. However, low dye solubility, leading to its precipitation, was a serious drawback of the original malachite green-based procedure [30] and its many later variants.
A solution to this problem was finding that increased acid concentration (2.5 M H2SO4) in stock dye solution surprisingly increases dye solubility and stability [15]. This finding allowed the formulation of a single stable color reagent that stops the enzymatic reaction and produces a green-blue color (630 nm), whose intensity is proportional to Pi concentration. The color reagent is made daily by adding the nonionic detergent Tween-20 (color stabilizer) to the stock dye/molybdate solution (Table 1). The signal produced by 2 µM Pi (corresponding to 1 µM PPi hydrolyzed) is 0.15 absorbance units in a 1 cm cuvette. The assay can be performed in a 96-well plate (Figure 2) and quantified with a plate reader.

4. Continuous and Fixed-Time Assays of PPi-Synthesizing Activity

Like any catalyst, PPase accelerates the attainment of thermodynamic equilibrium PPi ⇆ 2Pi in both directions. Although the equilibrium is shifted to the right (∆G0 = −20–25 kJ/mol), it allows the formation of measurable amounts of PPi under physiological conditions. For instance, the concentration of PPi at equilibrium with 10 mM Pi (pH 7, 25 °C, 1 mM Mg2+) is approximately 1 µM [31], which is below the detection limits of most analytical methods for PPi determination. However, micromolar sensitivity has been claimed for several assays based on fluorogenic and electrochemical PPi sensors [21,22,23,24,25]. Some of them are commercially available. However, they are of limited use in measurements of initial velocities of PPi synthesis, which require that PPi concentration in the medium does not exceed 10% of the equilibrium concentration to prevent product inhibition.
This problem is elegantly obviated in the coupled-enzyme luminescent procedure of Nyrén and Lundin [32], which later became a core element of “pyrosequencing”, a well-established DNA sequencing method [33]. In their procedure, PPi is converted to ATP by reaction with adenosine-5′-phosphosulfate (APS), catalyzed by ATP sulfurylase (E.C. 2.7.7.4), and the ATP formed is detected with a firefly luciferase/luciferin system. This method is used in three versions in PPase studies (Table 2 and Figure 3). In the first version, all components of the detection system, including ATP sulfurylase, luciferase, APS, and luciferin, are added to the PPase assay mix, and PPi formation is recorded continuously. This became possible because neither APS nor luciferin affects PPase activity, and both ATP sulfurylase and PPase need Mg2+ as a cofactor. The synthesized PPi is continuously converted into ATP at such a high rate that the steady-state PPi level is not inhibitory for PPase and luciferase activities. This is ensured by maintaining high ATP sulfurylase concentration, such that its doubling does not increase the measured rate of ATP accumulation. In the second assay version, the PPase reaction mix contains ATP sulfurylase/APS, which continuously converts the formed PPi into ATP, whose concentration is determined at fixed time points with luciferase/luciferin in a separate tube. Importantly, PPase does not hydrolyze ATP in the presence of Mg2+. Both assay versions combine the high sensitivity of the luciferase ATP determination with the possibility to accumulate detectable amounts of ATP while keeping the steady-state concentration of PPi at a sub-equilibrium, non-inhibitory level. ATP sulfurylase and luciferase act thus as amplifiers of the PPi-generated signal and allow the measurement of the initial velocities of PPi synthesis from Pi.
The third version of the PPi assay accesses the equilibrium of the PPi ⇆ 2Pi reaction in the active site of PPase. This information is required to estimate the rate constants for individual steps of PPase catalysis [37,38]. This equilibrium is markedly shifted to the left by comparing the equilibrium in solution—up to 20% of enzyme-bound Pi is converted to PPi [37,38]. To assay, the enzyme-bound PPi, the enzyme (20–100 µM) is incubated with Pi and Mg2+, inactivated by trifluoroacetic acid to release enzyme-bound PPi into solution, and an aliquot is taken to assay PPi using the coupled ATP sulfurylase/luciferase procedure.
Several comments on the assay procedure are appropriate. First, phosphoric acid and its salts are often contaminated with PPi, leading to high background luminescence. A low-PPi potassium phosphate can be prepared in the following way: phosphoric acid is diluted to 0.2 M with water, boiled for 3 h, and neutralized with KOH [39]. Second, low but significant background luminescence results from APS being a poor substrate for luciferase. Third, as Pi is added in high concentrations, care should be taken to ensure the assay mixture’s desired pH. It is not enough to adjust to this value the pH of the stock Pi solution because the formation of the MgHPO4 complex from H2PO4 upon mixing stock Pi and Mg2+ solutions would cause acidification of the medium (by 0.10 pH unit with 50 mM Pi, 5 mM Mg2+ in 0.1 M MOPS–KOH buffer, pH 7.2). This effect is compensated by adding an appropriate amount of KOH to the assay mixture. An Excel subroutine to calculate this amount, which depends on Pi, Mg2+, and H+ concentrations and reaction volume, is available (Supplementary Material). Alternatively, one can adjust the pH of the stock Pi solution used at each Mg2+ concentration to the value (above the desired assay pH) determined in a trial titration.
A different coupled-enzyme assay with similar characteristics has been described [40]. In this assay, PPi is converted to adenosine 5’-triphosphate (ATP) by pyruvate phosphate dikinase in the presence of its substrates, pyruvate phosphate and AMP. The ATP formed is similarly determined by the firefly luciferase reaction. The detection limit for PPi is approximately 10−8 M. The use of this assay in PPase studies has not yet been reported.

5. Ionic Equilibria in the Assay Media

Magnesium, the essential cofactor of PPases, has a dual role in catalysis—it activates the enzyme and substrates by forming complexes with both. Because the enzyme can sense only the actual substrate species in the medium, but not total substrate concentration, the magnesium complexes of the substrates PPi and Pi are named “true substrates.” A meaningful analysis of enzyme kinetics with the Michaelis–Menten equation should be done in terms of these true substrates. Furthermore, the substrate–concentration dependencies of enzyme activity should be measured at a fixed free Mg2+ ion concentration because it, not the total magnesium concentration, determines the fraction of the metal-complexed enzyme and, hence, its activity.
Pyrophosphoric acid undergoes a four-step dissociation, of which only the last two steps with the pKa3 of 6.38 and pKa4 of 9.11 should be considered in enzyme studies. Magnesium forms four complexes with three PPi species: MgP2O72−, Mg2P2O7, MgHP2O7 and MgH2P2O7 (Scheme 1). Furthermore, alkali metal ions, often present in the assay medium, form additional complexes: MP2O73−, M2P2O72− and MHP2O72− (M = K+ or Na+). At fixed pH and alkali metal ion concentrations, the complexation in this system can be described by simple Scheme 2, in which each species is the sum of all differently protonated individual species of the same magnesium:PPi stoichiometry.
The values of Kmpp and Km2pp can be calculated at each pH value and alkali metal concentration from the dissociation constants of the individual PPi complexes with H+, Mg2+, K+ and Na+ (tabulated in a previous publication [41]). Knowing the values of Kmpp and Km2pp allows calculating the total magnesium and PPi concentrations needed to maintain the required concentrations of MgPPi or Mg2PPi complex and free Mg2+. This algorithm was embedded in an Excel subroutine that calculates the composition of the assay medium at 25 °C and chosen pH and alkali metal concentrations (Supplementary Material).
The choice between MgPPi and Mg2PPi as the true substrate is largely arbitrary because the ratio of Mg2PPi and MgPPi concentrations is constant at a fixed Mg2+ concentration, and switching from one complex to the other would only change the apparent value of the Michaelis constant. In most reported studies, MgPPi has been assumed to be the true substrate for PPases of families I and II and Mg2PPi to be such for membrane PPases. However, one should keep in mind that steady-state kinetic studies can define only stoichiometry of the statistically significant enzyme species, but not the ways of their formation because schemes assuming different true substrates will lead to the same rate law and are, therefore, indistinguishable. These considerations are especially important to remember when working with the kinetic schemes considering both substrate and Mg2+ binding to the enzyme.
Pi is represented by a mixture of H2PO4 and HPO42− forms in the pH range 4–10, and only the latter species forms 1:1 complexes with Mg2+, K+, and Na+ (Scheme 3). The MgPi complex is generally assumed to be the true substrate in the PPase-catalyzed PPi synthesis. The total concentrations of Mg2+ and Pi needed to maintain the required concentrations of the MgPi complex and free Mg2+ ions are calculated using the apparent dissociation constant for MgPi. The latter is obtained similarly at each pH value and alkali metal concentration from the dissociation constants of the individual Pi complexes with H+, Mg2+, K+ and Na+ [42,43,44,45]. This algorithm was also embedded in an Excel subroutine (Supplementary Material), which additionally calculates the amount of the alkali needed to compensate for the acidification of the assay medium because of the selective complexation of the HPO42− species by Mg2+.

Supplementary Materials

The following is available: Microsoft Excel file: Calculations of medium composition.

Author Contributions

Writing—original draft preparation, A.A.B.; software and visualization, V.A.A.; writing—review and editing, A.M.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Academy of Finland (grant numbers 307775 and 335377) and by Lomonosov Moscow State University.

Data Availability Statement

The Excel subroutine that calculates the composition of the assay medium at chosen pH and alkali metal concentrations is available as Supplementary Material.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Estimation of Initial Velocities from Product Formation Curves

If substrate (PPi) concentration remains saturating during the recording time, resulting in a linear time-course, the initial velocity is obtained straightforwardly. However, if substrate consumption exceeds 20%, drawing the tangent to the product accumulation curve by eye may result in a large error in the initial velocity, more commonly underestimating [46] (pp. 80–82). There are two ways to increase the accuracy of such analysis. One can derive the explicit equation or system of equations that describe the integral reaction kinetics and fit it to the product formation curve. A simpler alternative is to use a semi-empirical procedure to manually draw the zero-time tangent to the curve. The procedure described below determines the second point, though, which the tangent must pass, in addition to the zero-time point. The required initial information includes the instrument signal (absorbance increase in recorder paper divisions) upon complete substrate conversion (Pinf). This parameter is calculated from the expected Pi concentration upon complete substrate conversion (twice the initial PPi concentration) and instrument sensitivity. One will also need an approximate Km value in terms of total PPi.
In a reaction with substrate exhaustion, the product accumulation curve will deviate from the tangent at the zero-time point, corresponding to initial velocity (Figure A1). The difference between the two curves (∆P) increases with time, and its knowledge at a single value of P suffices to draw the zero-time tangent. The smaller the correction (∆P), the more accurate is the initial velocity estimate. Obviously, the ∆P value will decrease with increasing Pinf and decreasing P and Km values. The equation for ∆P is obtained by combining Equation A1 and Equation A2, where V is the maximal velocity, t is time, Km is the Michaelis constant, and Pinf is the initial substrate concentration. Equation A1 describes product (P1) accumulation in time in a zero-order reaction proceeding with the rate equal to the enzyme-catalyzed reaction’s initial velocity. Equation A2 describes the true integral kinetics of the enzyme-catalyzed reaction [46] (p. 78). At any time point t, P1 = P + ∆P, and, accordingly, Equation A3 is valid. After its rearrangement, one obtains Equation A4, which links ∆P to Pinf, P, and Km values.
P1 = Vt/(1 + Km/Pinf)
AiP/Km + Km ln[Pinf/(PinfP)] = Vt
P/Km + Km ln[Pinf/(PinfP)] = (P + ∆P)(1 + Km/Pinf).
P = {P + Km ln[Pinf/(PinfP)]}/(1 + Km/Pinf) − P
Figure A1. Time-course of product accumulation in the enzyme-catalyzed reaction (solid line). The dashed straight line is the zero-time tangent to the solid curve, with the slope equal to the initial velocity.
Figure A1. Time-course of product accumulation in the enzyme-catalyzed reaction (solid line). The dashed straight line is the zero-time tangent to the solid curve, with the slope equal to the initial velocity.
Molecules 26 02356 g0a1
The dependencies of ∆P on Pinf calculated at several fixed P and Km values are presented in Figure A2. These graphs are used to determine ∆P at given Pinf and P values and draw the tangent for any reaction with substrate depletion. The top left panel applies to a first-order reaction. As expected, the ∆P value decreases, and the accuracy of the initial velocity estimate increases accordingly with increasing Pinf and decreasing P. Optimally, P should be chosen between 20 and 50% of Pinf, also keeping in mind that the signal-to-noise ratio becomes unfavorable at low P values. Although this treatment involves visual interpolation and is approximate, it decreases the error in the estimated initial velocity to less than 10%, which is sufficient for most uses. The simplicity of the procedure compared to that of Cornish-Bowden [47] and other known ones [46] (pp. 80–85) has made it a useful aid in the kinetic studies of PPase. Although we have been mainly using this correction procedure in the continuous assays, it is also applicable to and is especially important for the fixed-time assay carried out at non-saturating substrate concentration.
Figure A2. The dependence of ∆P in Figure A1 on Pinf at seven fixed values of P (shown on one panel and kept same on the other panels) and six values of Km (shown on the panels) for an enzyme-catalyzed reaction. For simplicity, the values of ∆P, Pinf, P and Km are shown in terms of the same arbitrary unit, for instance, recorder paper division. These graphs were created for an enzyme that obeys simple Michaelis–Menten kinetics with the product to substrate ratio of one. To use these graphs for PPase, which produces two Pi molecules from one PPi molecule, one must divide by two the measured values of P and Pinf if they are expressed in terms of Pi. Note that axis scaling is different on the two bottom panels.
Figure A2. The dependence of ∆P in Figure A1 on Pinf at seven fixed values of P (shown on one panel and kept same on the other panels) and six values of Km (shown on the panels) for an enzyme-catalyzed reaction. For simplicity, the values of ∆P, Pinf, P and Km are shown in terms of the same arbitrary unit, for instance, recorder paper division. These graphs were created for an enzyme that obeys simple Michaelis–Menten kinetics with the product to substrate ratio of one. To use these graphs for PPase, which produces two Pi molecules from one PPi molecule, one must divide by two the measured values of P and Pinf if they are expressed in terms of Pi. Note that axis scaling is different on the two bottom panels.
Molecules 26 02356 g0a2

References

  1. Heinonen, J.K. Biological Role of Inorganic Pyrophosphate; Kluwer Academic Publishers: London, UK, 2001. [Google Scholar]
  2. Baykov, A.A.; Malinen, A.M.; Luoto, H.H.; Lahti, R. Pyrophosphate-fueled Na+ and H+ transport in prokaryotes Microbiol. Mol. Biol. Rev. 2013, 77, 267–276. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Tsai, J.-Y.; Kellosalo, J.; Sun, Y.-J.; Goldman, A. Proton/sodium pumping pyrophosphatases: The last of the primary ion pumps. Curr. Opin. Struct. Biol. 2014, 27, 38–47. [Google Scholar] [CrossRef] [PubMed]
  4. Gaxiola, R.A.; Regmi, K.; Paez-Valencia, J.; Pizzio, G.; Zhang, S. Plant H+-PPases: Reversible enzymes with contrasting functions dependent on membrane environment. Mol. Plant. 2016, 9, 317–319. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Scholz-Starke, J.; Primo, C.; Yang, J.; Kandel, R.; Gaxiola, R.A.; Hirschi, K.D. The flip side of the Arabidopsis type I proton-pumping pyrophosphatase (AVP1): Using a transmembrane H+ gradient to synthesize pyrophosphate. J. Biol Chem. 2019, 294, 1290–1299. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Baykov, A.A.; Cooperman, B.S.; Goldman, A.; Lahti, R. Cytoplasmic inorganic pyrophosphatases. Progr. Mol. Subcell. Biol. 1999, 23, 127–150. [Google Scholar]
  7. Kajander, T.K.; Kellosalo, J.; Goldman, A. Inorganic pyrophosphatases: One substrate, three mechanisms. FEBS Lett. 2013, 587, 1863–1869. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Chappelet-Tordo, D.; Fosset, M.; Iwatsuba, M.; Gache, C.; Lazdunski, M. Intestinal alkaline phosphatase: Catalytic properties and half of the sites reactivity. Biochemistry 1974, 13, 1788–1795. [Google Scholar] [CrossRef]
  9. Lee, H.S.; Cho, Y.; Kim, Y.-J.; Lho, T.O.; Cha, S.-S.; Lee, J.-H.; Kang, S.G. A novel inorganic pyrophosphatase in Thermococcus onnurineus NA1. FEMS Microbiol. Lett. 2009, 300, 68–74. [Google Scholar] [CrossRef] [Green Version]
  10. Huang, H.; Patskovsky, Y.; Toro, R.; Farelli, J.D.; Pandya, C.; Almo, S.C.; Allen, K.N.; Dunaway-Mariano, D. Divergence of structure and function in the haloacid dehalogenase enzyme superfamily: Bacteroides thetaiotaomicron BT2127 is an inorganic pyrophosphatase. Biochemistry 2011, 50, 8937–8949. [Google Scholar] [CrossRef] [Green Version]
  11. Baykov, A.A.; Anashkin, V.A.; Salminen, A.; Lahti, R. Inorganic pyrophosphatases of Family II—two decades after their discovery. FEBS Lett. 2017, 591, 3225–3234. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Van Belle, H. New and sensitive reaction for automatic determination of inorganic phosphate and its application to serum. Anal. Biochem. 1970, 33, 132–142. [Google Scholar] [CrossRef]
  13. Baykov, A.A.; Avaeva, S.M. A simple and sensitive apparatus for continuous monitoring of orthophosphate in the presence of acid-labile compounds. Anal. Biochem. 1982, 116, 1–4. [Google Scholar] [CrossRef]
  14. Baykov, A.A.; Alexandrov, A.P.; Smirnova, I.N. A two-step mechanism of fluoride inhibition of rat liver inorganic pyrophosphatase. Arch. Biochem. Biophys. 1992, 294, 238–243. [Google Scholar] [CrossRef]
  15. Baykov, A.A.; Evtushenko, O.A.; Avaeva, S.M. A malachite green procedure for orthophosphate determination and its use in alkaline phosphatase-based enzyme immunoassay. Anal. Biochem. 1988, 171, 266–270. [Google Scholar] [CrossRef]
  16. Baykov, A.A.; Kasho, V.N.; Avaeva, S.M. Inorganic pyrophosphatase as a label in heterogeneous enzyme immunoassay. Anal. Biochem. 1988, 171, 271–276. [Google Scholar] [CrossRef]
  17. Vidilaseris, K.; Kellosalo, J.; Goldman, A. A high-throughput method for orthophosphate determination of thermostable membrane-bound pyrophosphatase activity. Anal. Methods 2018, 10, 646–651. [Google Scholar] [CrossRef] [Green Version]
  18. Himeno, S.; Ueda, T.; Shiomi, M.; Hori, T. Raman studies on the formation of 12-molybdopyrophosphate. Inorg. Chim. Acta 1997, 262, 219–223. [Google Scholar] [CrossRef]
  19. Cooperman, B.S.; Chiu, N.Y.; Bruckmann, R.H.; Bunick, G.J.; McKenna, G.P. Yeast inorganic pyrophosphatase. I. New methods of purification, assay, and crystallization. Biochemistry 1973, 12, 1665–1669. [Google Scholar] [CrossRef]
  20. Shakhov, Y.A.; Nyrén, P. A sensitive and rapid method for determination of pyrophosphatase activity. Acta Chem. Scand. 1982, B36, 689–694. [Google Scholar] [CrossRef] [Green Version]
  21. Kim, S.K.; Lee, D.H.; Hong, J.-I.; Yoon, J. Chemosensors for pyrophosphate. Acc. Chem. Res. 2009, 42, 23–31. [Google Scholar] [CrossRef]
  22. Zeng, Z.; Torriero, A.A.J.; Bond, A.M.; Spiccia, L. Fluorescent and electrochemical sensing of polyphosphate nucleotides by ferrocene functionalised with two Zn(II)(TACN)(pyrene) complexes. Chem. Eur. J. 2010, 16, 9154–9163. [Google Scholar] [CrossRef] [PubMed]
  23. Liu, D.J.; Credo, G.M.; Su, X.; Wu, K.; Lim, H.C.; Elibol, O.H.; Bashir, R.; Varma, M. Surface immobilizable chelator for label-free electrical detection of pyrophosphate. Chem. Commun. 2011, 47, 8310–8312. [Google Scholar] [CrossRef] [PubMed]
  24. Terenteva, E.A.; Apyari, V.V.; Dmitrienko, S.G.; Garshev, A.V.; Volkov, P.A.; Zolotov, Y.A. Determination of pyrophosphate and sulfate using polyhexamethylene guanidine hydrochloride-stabilized silver nanoparticles. Talanta 2018, 180, 346–351. [Google Scholar] [CrossRef]
  25. Lin, Y.; Hu, L.; Li, L.; Wang, K.; Ji, Y.; Zou, H. Electrochemical determination of pyrophosphate at nanomolar levels using a gold electrode covered with a cysteine nanofilm and based on competitive coordination of Cu(II) ion to cysteine and pyrophosphate. Microchim. Acta 2015, 182, 2069–2075. [Google Scholar] [CrossRef]
  26. Baykov, A.A.; Hyytia, T.; Turkina, M.V.; Efimova, I.S.; Kasho, V.N.; Goldman, A.; Cooperman, B.S.; Lahti, R. Functional characterization of Escherichia coli inorganic pyrophosphatase in zwitterionic buffers. Eur. J. Biochem. 1999, 260, 308–317. [Google Scholar] [CrossRef]
  27. Baykov, A.A.; Sergina, N.V.; Evtushenko, O.A.; Dubnova, E.B. Kinetic characterization of Rhodospirillum rubrum H+-pyrophosphatase in membrane-bound and isolated states. Eur. J. Biochem. 1996, 236, 121–127. [Google Scholar] [CrossRef]
  28. Gordon-Weeks, R.; Korenkov, V.D.; Steele, S.H.; Leigh, R.A. Tris is a competitive inhibitor of K+ activation of the vacuolar H+-pumping pyrophosphatase. Plant. Physiol. 1997, 114, 901–905. [Google Scholar] [CrossRef] [Green Version]
  29. Nagul, E.A.; McKelviea, I.D.; Worsfold, P.; Kolev, S.D. The molybdenum blue reaction for the determination of orthophosphate revisited: Opening the black box. Anal. Chim. Acta 2015, 890, 60–82. [Google Scholar] [CrossRef] [Green Version]
  30. Itaya, K.; Ui, M. A new micromethod for the colorimetric determination of inorganic phosphate. Clin. Chim. Acta 1966, 14, 361–366. [Google Scholar] [CrossRef]
  31. Flodgaard, H.; Fleron, P. Thermodynamic parameters for the hydrolysis of inorganic pyrophosphate at pH 7.4 as a function of [Mg2+], [K+], and ionic strength determined from equilibrium studies of the reaction. J. Biol. Chem. 1974, 249, 3465–3474. [Google Scholar] [CrossRef]
  32. Nyrén, P.; Lundin, A. Enzymatic method for continuous monitoring of inorganic pyrophosphate synthesis. Anal. Biochem. 1985, 151, 504–509. [Google Scholar] [CrossRef]
  33. Ronaghi, M.; Uhlén, M.; Nyrén, P. A sequencing method based on real-time pyrophosphate. Science 1998, 281, 363–365. [Google Scholar] [CrossRef]
  34. Baykov, A.A.; Fabrichniy, I.P.; Pohjanjoki, P.; Zyryanov, A.B.; Lahti, R. Fluoride effects along the reaction pathway of pyrophosphatase. Evidence for a second enzyme⋅pyrophosphate intermediate. Biochemistry 2000, 39, 11939–11947. [Google Scholar] [CrossRef]
  35. Belogurov, G.A.; Malinen, A.M.; Turkina, M.V.; Jalonen, U.; Rytkönen, K.; Baykov, A.A.; Lahti, R. Membrane-bound pyrophosphatase of Thermotoga maritima requires sodium for activity. Biochemistry 2005, 44, 4004–4010. [Google Scholar] [CrossRef]
  36. Fabrichniy, I.P.; Kasho, V.N.; Hyytia, T.; Salminen, T.; Halonen, P.; Dudarenkov, V.Y.; Heikinheimo, P.; Chernyak, V.Y.; Goldman, A.; Lahti, R.; et al. Structural and functional consequenses of substitutions at the tyrosine 55-lysine 104 hydrogen bond in Escherichia coli inorganic pyrophosphatase. Biochemistry 1997, 36, 7746–7753. [Google Scholar] [CrossRef]
  37. Springs, B.; Welsh, K.M.; Cooperman, B.S. Thermodynamics, kinetics, and mechanism in yeast inorganic pyrophosphatase catalysis of inorganic pyrophosphate: Inorganic phosphate equilibration. Biochemistry 1981, 20, 6384–6391. [Google Scholar] [CrossRef]
  38. Baykov, A.A.; Shestakov, A.S.; Kasho, V.N.; Vener, A.V.; Ivanov, A.H. Kinetics and thermodynamics of catalysis by the inorganic pyrophosphatase of Escherichia coli in both directions. Eur. J. Biochem. 1990, 194, 879–887. [Google Scholar] [CrossRef]
  39. Smirnova, I.N.; Kasho, V.N.; Volk, S.E.; Ivanov, A.H.; Baykov, A.A. Rates of elementary steps catalyzed by rat liver cytosolic and mitochondrial inorganic pyrophosphatases in both directions. Arch. Biochem. Biophys. 1995, 318, 340–348. [Google Scholar] [CrossRef] [PubMed]
  40. Arakawa, H.; Karasawa, K.; Igarashi, T.; Suzuki, S.; Goto, N.; Maeda, M. Detection of cariogenic bacteria genes by a combination of allele-specific polymerase chain reactions and a novel bioluminescent pyrophosphate assay. Anal. Biochem. 2004, 333, 296–302. [Google Scholar] [CrossRef] [PubMed]
  41. Baykov, A.A.; Bakuleva, N.P.; Rea, P.A. Steady-state kinetics of substrate hydrolysis by vacuolar H+-pyrophosphatase. A simple three-state model. Eur. J. Biochem. 1993, 217, 755–762. [Google Scholar] [CrossRef] [PubMed]
  42. Smith, R.M.; Alberty, R.A. The apparent stability constants of ionic complexes of various adenosine phosphates with divalent cations. J. Am. Chem. Soc. 1956, 78, 2376–2380. [Google Scholar] [CrossRef]
  43. Childs, C.W. Potentiometric study of equilibriums in aqueous divalent metal orthophosphate solutions. Inorg. Chem. 1970, 9, 2465–2469. [Google Scholar] [CrossRef]
  44. Smirnova, I.N.; Shestakov, A.S.; Dubnova, E.B.; Baykov, A.A. Spectral and kinetic studies of phosphate and magnesium ion binding to yeast inorganic pyrophosphatase. Eur. J. Biochem. 1989, 182, 451–456. [Google Scholar] [CrossRef] [PubMed]
  45. Daniele, P.G.; De Robertis, A.; De Stefano, C.; Gianguzza, A.; Sammartano, S. Salt effects on the protonation of orthophosphate between 10 and 50 degrees Celsius in aqueous solution. A complex formation model. J. Solut. Chem. 1991, 20, 495–515. [Google Scholar] [CrossRef]
  46. Bisswanger, H. Enzyme Kinetics. Principles and Methods, 2nd ed.; Wiley-VCH Verlag: Weinheim, Germany, 2008; pp. 78–85. [Google Scholar]
  47. Cornish-Bowden, A. The use of the direct linear plot for determining initial velocities. Biochem. J. 1975, 149, 305–312. [Google Scholar] [CrossRef] [PubMed]
Figure 1. The phosphate analyzer is used to assay PPase activity in a continuous way. (A) Flow diagram for the phosphate analyzer in standard mode; (B) Tubing connections on the pump in the high-sensitivity mode; (C) tubing connections in the low dead-time mode. Numbers on the pump refer to flow rate in mL/min (before the slash) and tubing diameter in mm (panel A) or flow rate in mL/min (panels B and C). (D) Actual Pi accumulation recordings in setup A with photometer sensitivity of 1 absorbance unit per recorder scale. The calibration data shown at the beginning of the recording was obtained by adding 0–200 µM Pi to the reaction buffer. (E) Actual Pi accumulation recordings in setup B with photometer sensitivity of 0.1 absorbance unit per recorder scale. The assay mixture of 40 mL volume contained 140 µM PPi, 5 mM MgCl2, 50 mM MOPS–KOH, pH 7.2, and 0.03 nM Streptococcus gordonii PPase with a specific activity of 480 s−1. (F) Actual recordings of Pi accumulation in the setup C for rat liver PPase in the presence (a) or absence (b) of slow-binding inhibitor (10 mM fluoride). The arrow marks the moment of enzyme addition. Part of the figure was taken with permission from references [13] (panels A and D) and [14] (panels C and F).
Figure 1. The phosphate analyzer is used to assay PPase activity in a continuous way. (A) Flow diagram for the phosphate analyzer in standard mode; (B) Tubing connections on the pump in the high-sensitivity mode; (C) tubing connections in the low dead-time mode. Numbers on the pump refer to flow rate in mL/min (before the slash) and tubing diameter in mm (panel A) or flow rate in mL/min (panels B and C). (D) Actual Pi accumulation recordings in setup A with photometer sensitivity of 1 absorbance unit per recorder scale. The calibration data shown at the beginning of the recording was obtained by adding 0–200 µM Pi to the reaction buffer. (E) Actual Pi accumulation recordings in setup B with photometer sensitivity of 0.1 absorbance unit per recorder scale. The assay mixture of 40 mL volume contained 140 µM PPi, 5 mM MgCl2, 50 mM MOPS–KOH, pH 7.2, and 0.03 nM Streptococcus gordonii PPase with a specific activity of 480 s−1. (F) Actual recordings of Pi accumulation in the setup C for rat liver PPase in the presence (a) or absence (b) of slow-binding inhibitor (10 mM fluoride). The arrow marks the moment of enzyme addition. Part of the figure was taken with permission from references [13] (panels A and D) and [14] (panels C and F).
Molecules 26 02356 g001
Figure 2. PPase assay using the malachite green procedure. Two bottom rows show a duplicate series of phosphate dilutions from 0 µM (left) to 10 µM (right). Two top rows show typical results of a duplicate screening test of a library of potential inhibitors of Escherichia coli PPase. The yellow color indicates strong inhibition, dark green color—no inhibition.
Figure 2. PPase assay using the malachite green procedure. Two bottom rows show a duplicate series of phosphate dilutions from 0 µM (left) to 10 µM (right). Two top rows show typical results of a duplicate screening test of a library of potential inhibitors of Escherichia coli PPase. The yellow color indicates strong inhibition, dark green color—no inhibition.
Molecules 26 02356 g002
Figure 3. Schematics of the assays to measure PPase-catalyzed PPi synthesis. (A) A continuous assay of the medium PPi synthesis; (B) fixed-time assay of the medium PPi synthesis; (C) determination of enzyme-bound PPi. The assayed PPase is added to all far-left tubes. Three other major components are shown as colored spots. The principal analytes transferred between the tubes are indicated above the arrows. The blue star refers to the luminescence signal. (D) Actual PPi accumulation recordings in the assay version A for baker’s yeast PPase in the presence of slow-binding inhibitor (fluoride; its concentrations in mM are indicated on the curves). Panel D was taken with permission from reference [34].
Figure 3. Schematics of the assays to measure PPase-catalyzed PPi synthesis. (A) A continuous assay of the medium PPi synthesis; (B) fixed-time assay of the medium PPi synthesis; (C) determination of enzyme-bound PPi. The assayed PPase is added to all far-left tubes. Three other major components are shown as colored spots. The principal analytes transferred between the tubes are indicated above the arrows. The blue star refers to the luminescence signal. (D) Actual PPi accumulation recordings in the assay version A for baker’s yeast PPase in the presence of slow-binding inhibitor (fluoride; its concentrations in mM are indicated on the curves). Panel D was taken with permission from reference [34].
Molecules 26 02356 g003
Scheme 1. Complex formation between PPi, H+, Mg2+, K+ and Na+ ions under physiological conditions. The number above or beside arrows refers to the minus logarithm of the respective dissociation constant [41].
Scheme 1. Complex formation between PPi, H+, Mg2+, K+ and Na+ ions under physiological conditions. The number above or beside arrows refers to the minus logarithm of the respective dissociation constant [41].
Molecules 26 02356 sch001
Scheme 2. A simplified description of complexation between PPi and Mg2+ at fixed pH and alkali metal ion concentrations.
Scheme 2. A simplified description of complexation between PPi and Mg2+ at fixed pH and alkali metal ion concentrations.
Molecules 26 02356 sch002
Scheme 3. Complex formation between Pi, H+, Mg2+, K+, and Na+ ions under physiological conditions. The number above or beside arrows refers to the minus logarithm of the respective dissociation constant [42,43,44,45].
Scheme 3. Complex formation between Pi, H+, Mg2+, K+, and Na+ ions under physiological conditions. The number above or beside arrows refers to the minus logarithm of the respective dissociation constant [42,43,44,45].
Molecules 26 02356 sch003
Table 1. Assays measuring PPi hydrolysis. Instrument setups for the continuous assay versions A–C are shown in Figure 1A–C.
Table 1. Assays measuring PPi hydrolysis. Instrument setups for the continuous assay versions A–C are shown in Figure 1A–C.
Assay Type and ProcedureComments
Continuous, version A [13]
Stock dye solution contains 80 mg/L methyl green and 3.4 g/L Triton X-305. Stock acid/molybdate solution contains 0.55 M H2SO4 (32 mL concentrated H2SO4 per 1 L final solution 1), 12 g/L ammonium molybdate, and 2 µM Pi. The mixing ratio sample: acid/molybdate:dye is 1.5:7:2. Washing solution: 0.1 M NaOH and 2 g/L Triton X-305. The enzymatic reaction is performed in a thermostated vessel with magnetic stirring at 10–60 °C; the reaction volume is typically 5 mL.
Methyl green and Triton X-305 are from Sigma-Aldrich, ammonium molybdate from Fisher Scientific. Phosphate-free sulfuric acid (1.83 g/cm3) is used. When switching to a new batch of Triton X-305, its concentration may need adjustment to ensure linearity of the calibration plot and low baseline drift because of dye deposition on the photometer cuvette.
Continuous, version B [13]
Stock dye solution contains 86 mg/L methyl green and 3.7 g/L Triton X-305. Stock acid/molybdate solution contains 1.8 M H2SO4 (105 mL concentrated H2SO4 per 1 L final solution), 39 g/L ammonium molybdate, and 7 µM Pi. The mixing ratio sample:acid/molybdate:dye is 7:2.3:2. The PPase reaction volume is typically 25–40 mL.
Acid-resistant Iso-Versinic tubing is preferable on the pump for the acid/molybdate solution in this case.
Continuous, version C [14]
Stock dye solution contains 30 mg/L methyl green and 1.3 g/L Triton X-305. Stock acid/molybdate solution contains 1.0 M H2SO4 (59 mL concentrated H2SO4 per 1 L final solution), 22 g/L ammonium molybdate, and 4 µM Pi. The mixing ratio sample: acid/molybdate:dye is 1.2:3.2:4.4.
Fixed-time [15,16]
Stock dye/molybdate solution (stable for months in a refrigerator): 115 mg of malachite green are dissolved in 100 mL of 2.5 M H2SO4, followed by 1.4 g of ammonium molybdate. On the day of use, 0.25 mL of 10% Tween-20 (w/w) is added per 10 mL of the dye/molybdate solution. The PPase reaction is performed in a total volume of 0.2 mL in 96-well plates and terminated by adding 0.05 mL of the color reagent. Absorbance at 630 nm is measured after 10 min.
Malachite green is from Sigma-Aldrich, Tween-20 from Ferak Berlin. Tween-20 (0.05%, w/w) can be added instead to the PPase assay mixture, which typically contains 0.05 M Tris-HCl or another buffer, 0.05 mM PPi, and 5 mM MgCl2.
1 Note: dilution of sulfuric acid generates much heat, which may cause boiling/spilling. It is advised to add the acid slowly to water in a Duran/Pyrex glass beaker in a fume hood.
Table 2. Assays measuring PPi formation.
Table 2. Assays measuring PPi formation.
Assay Type and ProcedureComments
PPi synthesis (continuous) [34]
Assay mixture contains calculated amounts of potassium phosphate and MgCl2, 0.7 U/mL ATP-sulfurylase, 10 µM APS, 5 µL of luciferin/luciferase solution (Sigma ATP assay mix, catalog no. FLAAM, reconstituted with 5 mL of water), 1 mM dithiothreitol, 0.8 mg/mL bovine serum albumin and 0.1 M MOPS–KOH buffer in a total volume of 0.2 mL. The reaction is initiated by adding PPase, and the time-course of luminescence is followed with a luminometer (e.g., LKB model 1250). After the PPi-generated signal stabilizes, 2 µL of 10 µM ATP are added for internal calibration.
ATP sulfurylase is from BioLabs (UK). Its concentration should be sufficiently high so that its doubling does not change the measured rate of ATP accumulation. The luminescence-versus-time dependencies are slightly curved because of the slow inactivation of luciferase in the reaction medium.
PPi synthesis (fixed-time) [35]
Assay mixture contains calculated amounts of potassium phosphate and MgCl2, 0.7 U/mL ATP-sulfurylase, 10 µM APS, and 0.1 M MOPS–KOH buffer in a total volume of 100 µL. The reaction is initiated by adding PPase and carried out for 10 min. Aliquots (15 µL) are withdrawn at various times, quenched with 15 µL of 1 M trifluoroacetic acid, incubated for 4 min at room temperature, and neutralized with 15 µL of 1.5 M Tris. The ATP formed is quantitated by adding 10 µL of the mixture to 200 µL of 0.2 M Tris-HCl buffer, pH 8.0 containing 6 µL of the luciferin/luciferase solution (Sigma ATP assay mix, catalog No. FLAAAM, reconstituted with 5 mL of water) and measuring the luminescence. After the signal stabilizes, 2 µL of 10 µM ATP are added for internal calibration.
Crystalline phosphoric acid (Fluka) is freed from PPi contamination as described in the text.
Enzyme-bound PPi formation [36]
PPase (typically 20–100 µM) is preequilibrated with Pi and Mg2+ under appropriate conditions in a 50 µL volume and quenched with 10 µL of 5 M trifluoroacetic acid. After several minutes, precipitated protein is removed by centrifugation, a 10 µL aliquot of the supernatant is added to 0.2 mL of the PPi assay cocktail (0.2 M Tris-HCl, pH 8.0, 0.4 U/mL ATP sulfurylase, 10 µM APS, 6 µL of the luciferase/luciferin solution, 1 mM dithiothreitol, 0.8 mg bovine serum albumin, 30 µM EGTA), and the luminescence is recorded. After the signal stabilizes, 2 µL of 10 µM PPi are added for internal calibration.
Determined PPi content refers to the sum of enzyme-bound and medium PPi. The latter is measured similarly but using a 100 times lower enzyme concentration and subtracted.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Baykov, A.A.; Anashkin, V.A.; Malinen, A.M. Good-Practice Non-Radioactive Assays of Inorganic Pyrophosphatase Activities. Molecules 2021, 26, 2356. https://doi.org/10.3390/molecules26082356

AMA Style

Baykov AA, Anashkin VA, Malinen AM. Good-Practice Non-Radioactive Assays of Inorganic Pyrophosphatase Activities. Molecules. 2021; 26(8):2356. https://doi.org/10.3390/molecules26082356

Chicago/Turabian Style

Baykov, Alexander A., Viktor A. Anashkin, and Anssi M. Malinen. 2021. "Good-Practice Non-Radioactive Assays of Inorganic Pyrophosphatase Activities" Molecules 26, no. 8: 2356. https://doi.org/10.3390/molecules26082356

Article Metrics

Back to TopTop