Protein film voltammetry

Last updated

In electrochemistry, protein film voltammetry (or protein film electrochemistry, or direct electrochemistry of proteins) is a technique for examining the behavior of proteins immobilized (either adsorbed or covalently attached) on an electrode. The technique is applicable to proteins and enzymes that engage in electron transfer reactions and it is part of the methods available to study enzyme kinetics.[ citation needed ]

Contents

Provided that it makes suitable contact with the electrode surface (electron transfer between the electrode and the protein is direct) and provided that it is not denatured, the protein can be fruitfully interrogated by monitoring current as a function of electrode potential and other experimental parameters.

Various electrode materials can be used. [1] Special electrode designs are required to address membrane-bound proteins. [2]

Experiments with redox proteins

Small redox proteins such as cytochromes and ferredoxins can be investigated on condition that their electroactive coverage (the amount of protein undergoing direct electron transfer) is large enough (in practice, greater than a fraction of pmol/cm2).

Electrochemical data obtained with small proteins can be used to measure the redox potentials of the protein's redox sites, [3] the rate of electron transfer between the protein and the electrode, [4] or the rates of chemical reactions (such as protonations) that are coupled to electron transfer. [5]

Interpretation of the peak current and peak area

Fig 1. Voltammetric signals for a one-electron (A) or two-electron (B) redox species adsorbed onto an electrode in the limit of slow scan rate. The current plotted here is normalized by
F
2
n
A
G
/
R
T
{\displaystyle F^{2}\nu A\Gamma /RT}
, implying that the measured current increases in proportion to scan rate (
n
{\displaystyle \nu }
). Voltammetric signals for a redox species adsorbed onto an electrode in the limit of slow scan rate.svg
Fig 1. Voltammetric signals for a one-electron (A) or two-electron (B) redox species adsorbed onto an electrode in the limit of slow scan rate. The current plotted here is normalized by , implying that the measured current increases in proportion to scan rate ().
Fig 2. Effect of scan rate on the voltammetry of a one-electron redox species adsorbed onto an electrode (the current shown here is normalized by
F
2
n
A
G
/
R
T
{\displaystyle F^{2}\nu A\Gamma /RT}
, implying that the measured current increases in proportion to scan rate (
n
{\displaystyle \nu }
)) calculated for
a
=
0.5
{\displaystyle \alpha =0.5}
,
T
=
298
{\displaystyle T=298}
K. The scan rate increases from blue to red. A: voltammograms. B: peak potentials Laviron2.svg
Fig 2. Effect of scan rate on the voltammetry of a one-electron redox species adsorbed onto an electrode (the current shown here is normalized by , implying that the measured current increases in proportion to scan rate ()) calculated for , K. The scan rate increases from blue to red. A: voltammograms. B: peak potentials

In a cyclic voltammetry experiment carried out with an adsorbed redox protein, the oxidation and reduction of each redox site shows as a pair of positive and negative peaks. Since all the sample is oxidised or reduced during the potential sweep, the peak current and peak area should be proportional to scan rate (observing that the peak current is proportional to scan rate proves that the redox species that gives the peak is actually immobilised). [3] The same is true for experiments performed with non-biological redox molecules adsorbed onto electrodes. The theory was mainly developed by the French electrochemist Etienne Laviron in the 1980s [4] , [6] ,. [7]

Since both this faradaic current (which results from the oxidation/reduction of the adsorbed molecule) and the capacitive current (which results from electrode charging) increase in proportion to scan rate, the peaks should remain visible when the scan rate is increased. In contrast, when the redox analyte is in solution and diffuses to/from the electrode, the peak current is proportional to the square root of the scan rate (see: Randles–Sevcik equation).

Peak area

Irrespective of scan rate, the area under the peak (in units of AV) is equal to , where is the number of electrons exchanged in the oxidation/reduction of the center, is the electrode surface and is the electroactive coverage (in units of mol/cm2). [3] The latter can therefore be deduced from the area under the peak after subtraction of the capacitive current.

Peak shape

Slow scan rate

At slow scan rates there should be no separation between the oxidative and reductive peaks.

  • A one-electron site (e.g. a heme or FeS cluster) gives a broad peak (fig 1A). The equation that gives the shape and intensity of the peak is:
Ideally, the peak position is in both directions. The peak current is (it is proportional to scan rate, , and to the amount of redox sites on the electrode, ). The ideal half width at half height (HWHH) equates mV at 20 °C. Non-ideal behaviour may result in the peak being broader than the ideal limit.
  • The peak shape for a two-electron redox site (e.g. a flavin) depends on the stability of the half-reduced state (fig 1B). If the half-reduced state is stable over a large range of electrode potential, the signal is the sum of two one-electron peaks (purple line in fig 1B). If the half reduced state is unstable, the signal is a single peak (red line in fig 1B), which may have up to four times the height and half the width of a one-electron peak. [6] , [8]
  • A protein that contains multiple redox centers should give multiple peaks which all have the same area (scaled by ).
Fast scan rates

If the reaction is a simple electron transfer reaction, the peaks should remain symmetrical at fast scan rates. A peak separation is observed when the scan rate , where is the exchange electron transfer rate constant in Butler Volmer theory. Laviron equation [4] , [8] , [9] predicts that at fast scan rates, the peaks separate in proportion to . The larger or the smaller , the larger the peak separation. The peak potentials are , [4] as shown by lines in fig 2B ( is the charge transfer coefficient). Examining the experimental change in peak position against scan rate therefore informs on the rate of interfacial electron transfer .

Effect of coupled chemical reactions

Coupled reactions are reactions whose rate or equilibrium constant is not the same for the oxidized and reduced forms of the species that is being investigated. For example, reduction should favour protonation (): the protonation reaction is coupled to the reduction at . The binding of a small molecule (other than the proton) may also be coupled to a redox reaction.

Two cases must be considered depending on whether the coupled reaction is slow or fast (meaning that the time scale of the coupled reaction is larger or smaller than the voltammetric time scale [10] ).

  • Fast chemical reactions that are coupled to electron transfer (such as protonation) only affect the apparent values of and , [7] but the peaks remain symmetrical. The dependence of on ligand concentration (e.g. the dependence of on pH plotted in a Pourbaix diagram) can be interpreted to obtain the dissociation constants (e.g. acidity constants) from the oxidized or reduced forms of the redox species.
  • Asymmetry may result from slow chemical reactions that are coupled to (and gate) the electron transfer. From fast scan voltammetry, information can be gained about the rates of the reactions that are coupled to electron transfer. The case of and reversible surface electrochemical reactions followed by irreversible chemical reactions was addressed by Laviron in refs [6] , [11] but the data are usually interpreted using the numerical solution of the appropriate differential equations. [9]

Experiments with redox enzymes

In studies of enzymes, the current results from the catalytic oxydation or reduction of the enzyme's substrate.

The electroactive coverage of large redox enzymes (such as laccase, hydrogenase etc.) is often too low to detect any signal in the absence of substrate, but the electrochemical signal is amplified by catalysis: indeed, the catalytic current is proportional to turnover rate times electroactive coverage. The effect of varying the electrode potential, the pH or the concentration of substrates and inhibitors etc. can be examined to learn about various steps in the catalytic mechanism. [8]

Interpretation of the value of the catalytic current

For an enzyme immobilised on an electrode, the value of the current at a certain potential equates , where is the number of electrons exchanged in the catalytic reaction, is the electrode surface, is the electroactive coverage, and TOF is the turnover frequency (or "turnover number"), that is, the number of substrate molecules transformed per second and per molecule of adsorbed enzyme).The latter can be deduced from the absolute value of the current only on condition that is known, which is rarely the case. However, information is obtained by analysing the relative change in current that results from changing the experimental conditions.

The factors that may influence the TOF are (i) the mass transport of substrate towards the electrode where the enzyme is immobilised (diffusion and convection), (ii) the rate of electron transfer between the electrode and the enzyme (interfacial electron transfer), and (iii) the "intrinsic" activity of the enzyme, all of which may depend on electrode potential.

The enzyme is often immobilized on a rotating disk working electrode (RDE) that is spun quickly to prevent the depletion of the substrate near the electrode. In that case, mass transport of substrate towards the electrode where the enzyme is adsorbed may not be influential.

Steady-state voltammetric response

Under very oxidising or very reducing conditions, the steady-state catalytic current sometimes tends to a limiting value (a plateau) which (still provided there is no mass transport limitation) relates to the activity of the fully oxidised or fully reduced enzyme, respectively. If interfacial electron transfer is slow and if there is a distribution of electron transfer rates (resulting from a distribution of orientations of the enzymes molecules on the electrode), the current keeps increasing linearly with potential instead of reaching a plateau; in that case the limiting slope is proportional to the turnover rate of the fully oxidised or fully reduced enzyme. [8]

The change in steady-state current against potential is often complex (e.g. not merely sigmoidal). [12]

Departure from steady-state

Another level of complexity comes from the existence of slow redox-driven reactions that may change the activity of the enzyme and make the response depart from steady-state. [13] Here, slow means that the time scale of the (in)activation is similar to the voltammetric time scale [10] . If a RDE is used, these slow (in)activations are detected by a hysteresis in the catalytic voltammogram that is not due to mass-transport. The hysteresis may disappear at very fast scan rates (if the inactivation has no time to proceed) or at very slow scan rates (if the (in)activation reaction reaches a steady-state). [14]

Combining protein film voltammetry and spectroscopy

Conventional voltammetry offers a limited picture of the enzyme-electrode interface and on the structure of the species involved in the reaction. Complementing standard electrochemistry with other methods can provide a more complete picture of catalysis. [15] [16] [17]

Related Research Articles

<span class="mw-page-title-main">Electrochemistry</span> Branch of chemistry

Electrochemistry is the branch of physical chemistry concerned with the relationship between electrical potential difference, as a measurable and quantitative phenomenon, and identifiable chemical change, with the potential difference as an outcome of a particular chemical change, or vice versa. These reactions involve electrons moving via an electronically-conducting phase between electrodes separated by an ionically conducting and electronically insulating electrolyte.

<span class="mw-page-title-main">Scanning tunneling microscope</span> Instrument able to image surfaces at the atomic level by exploiting quantum tunneling effects

A scanning tunneling microscope (STM) is a type of microscope used for imaging surfaces at the atomic level. Its development in 1981 earned its inventors, Gerd Binnig and Heinrich Rohrer, then at IBM Zürich, the Nobel Prize in Physics in 1986. STM senses the surface by using an extremely sharp conducting tip that can distinguish features smaller than 0.1 nm with a 0.01 nm depth resolution. This means that individual atoms can routinely be imaged and manipulated. Most microscopes are built for use in ultra-high vacuum at temperatures approaching zero kelvin, but variants exist for studies in air, water and other environments, and for temperatures over 1000 °C.

In electrochemistry, the Nernst equation is a chemical thermodynamical relationship that permits the calculation of the reduction potential of a reaction from the standard electrode potential, absolute temperature, the number of electrons involved in the oxydo-reduction reaction, and activities of the chemical species undergoing reduction and oxidation respectively. It was named after Walther Nernst, a German physical chemist who formulated the equation.

<span class="mw-page-title-main">Galvanic cell</span> Electrochemical device

A galvanic cell or voltaic cell, named after the scientists Luigi Galvani and Alessandro Volta, respectively, is an electrochemical cell in which an electric current is generated from spontaneous Oxidation-Reduction reactions. A common apparatus generally consists of two different metals, each immersed in separate beakers containing their respective metal ions in solution that are connected by a salt bridge or separated by a porous membrane.

<span class="mw-page-title-main">Cyclic voltammetry</span>

Cyclic voltammetry (CV) is a type of potentiodynamic electrochemical measurement. In a cyclic voltammetry experiment, the working electrode potential is ramped linearly versus time. Unlike in linear sweep voltammetry, after the set potential is reached in a CV experiment, the working electrode's potential is ramped in the opposite direction to return to the initial potential. These cycles of ramps in potential may be repeated as many times as needed. The current at the working electrode is plotted versus the applied voltage to give the cyclic voltammogram trace. Cyclic voltammetry is generally used to study the electrochemical properties of an analyte in solution or of a molecule that is adsorbed onto the electrode.

<span class="mw-page-title-main">Chronoamperometry</span>

Chronoamperometry is an electrochemical technique in which the potential of the working electrode is stepped and the resulting current from faradaic processes occurring at the electrode is monitored as a function of time. The functional relationship between current response and time is measured after applying single or double potential step to the working electrode of the electrochemical system. Limited information about the identity of the electrolyzed species can be obtained from the ratio of the peak oxidation current versus the peak reduction current. However, as with all pulsed techniques, chronoamperometry generates high charging currents, which decay exponentially with time as any RC circuit. The Faradaic current - which is due to electron transfer events and is most often the current component of interest - decays as described in the Cottrell equation. In most electrochemical cells this decay is much slower than the charging decay-cells with no supporting electrolyte are notable exceptions. Most commonly a three electrode system is used. Since the current is integrated over relatively longer time intervals, chronoamperometry gives a better signal to noise ratio in comparison to other amperometric techniques.

In electrochemistry, overpotential is the potential difference (voltage) between a half-reaction's thermodynamically determined reduction potential and the potential at which the redox event is experimentally observed. The term is directly related to a cell's voltage efficiency. In an electrolytic cell the existence of overpotential implies that the cell requires more energy than thermodynamically expected to drive a reaction. In a galvanic cell the existence of overpotential means less energy is recovered than thermodynamics predicts. In each case the extra/missing energy is lost as heat. The quantity of overpotential is specific to each cell design and varies across cells and operational conditions, even for the same reaction. Overpotential is experimentally determined by measuring the potential at which a given current density is achieved.

<span class="mw-page-title-main">Tafel equation</span> Equation relating the rate of an electrochemical reaction to the overpotential

The Tafel equation is an equation in electrochemical kinetics relating the rate of an electrochemical reaction to the overpotential. The Tafel equation was first deduced experimentally and was later shown to have a theoretical justification. The equation is named after Swiss chemist Julius Tafel.

" It describes how the electrical current through an electrode depends on the voltage difference between the electrode and the bulk electrolyte for a simple, unimolecular redox reaction ".

Electroanalytical methods are a class of techniques in analytical chemistry which study an analyte by measuring the potential (volts) and/or current (amperes) in an electrochemical cell containing the analyte. These methods can be broken down into several categories depending on which aspects of the cell are controlled and which are measured. The three main categories are potentiometry, coulometry, and voltammetry.

<span class="mw-page-title-main">Double layer (surface science)</span> Condensed matter physics

A double layer is a structure that appears on the surface of an object when it is exposed to a fluid. The object might be a solid particle, a gas bubble, a liquid droplet, or a porous body. The DL refers to two parallel layers of charge surrounding the object. The first layer, the surface charge, consists of ions adsorbed onto the object due to chemical interactions. The second layer is composed of ions attracted to the surface charge via the Coulomb force, electrically screening the first layer. This second layer is loosely associated with the object. It is made of free ions that move in the fluid under the influence of electric attraction and thermal motion rather than being firmly anchored. It is thus called the "diffuse layer".

In chemistry, an electrochemical reaction mechanism is the step by step sequence of elementary steps, involving at least one outer sphere electron transfer, by which an overall chemical change occurs.

A rotating disk electrode (RDE) is a working electrode used in three electrode systems for hydrodynamic voltammetry. The electrode rotates during experiments inducing a flux of analyte to the electrode. These working electrodes are used in electrochemical studies when investigating reaction mechanisms related to redox chemistry, among other chemical phenomena. The more complex rotating ring-disk electrode can be used as a rotating disk electrode if the ring is left inactive during the experiment.

An ultramicroelectrode (UME) is a working electrode used in a voltammetry. The small size of UME give them large diffusion layers and small overall currents. These features allow UME to achieve useful steady-state conditions and very high scan rates (V/s) with limited distortion. UME were developed independently by Wightman and Fleischmann around 1980. Small current at UME enables electrochemical measurements in low conductive media, where voltage drop associated with high solution resistance makes these experiments difficult for conventional electrodes. Furthermore, small voltage drop at UME leads to a very small voltage distortion at the electrode-solution interface which allows using two-electrode setup in voltammetric experiment instead of conventional three-electrode setup.

In electrochemistry, ITIES is an electrochemical interface that is either polarisable or polarised. An ITIES is polarisable if one can change the Galvani potential difference, or in other words the difference of inner potentials between the two adjacent phases, without noticeably changing the chemical composition of the respective phases. An ITIES system is polarised if the distribution of the different charges and redox species between the two phases determines the Galvani potential difference.

In electrochemistry, exchange current density is a parameter used in the Tafel equation, Butler–Volmer equation and other electrochemical kinetics expressions. The Tafel equation describes the dependence of current for an electrolytic process to overpotential.

Charge transfer coefficient, and symmetry factor are two related parameters used in description of the kinetics of electrochemical reactions. They appear in the Butler–Volmer equation and related expressions.

In cyclic voltammetry, the Randles–Ševčík equation describes the effect of scan rate on the peak current ip. For simple redox events such as the ferrocene/ferrocenium couple, ip depends not only on the concentration and diffusional properties of the electroactive species but also on scan rate.

Scanning electrochemical microscopy (SECM) is a technique within the broader class of scanning probe microscopy (SPM) that is used to measure the local electrochemical behavior of liquid/solid, liquid/gas and liquid/liquid interfaces. Initial characterization of the technique was credited to University of Texas electrochemist, Allen J. Bard, in 1989. Since then, the theoretical underpinnings have matured to allow widespread use of the technique in chemistry, biology and materials science. Spatially resolved electrochemical signals can be acquired by measuring the current at an ultramicroelectrode (UME) tip as a function of precise tip position over a substrate region of interest. Interpretation of the SECM signal is based on the concept of diffusion-limited current. Two-dimensional raster scan information can be compiled to generate images of surface reactivity and chemical kinetics.

<span class="mw-page-title-main">Fast-scan cyclic voltammetry</span>

Fast-scan cyclic voltammetry (FSCV) is cyclic voltammetry with a very high scan rate (up to 1×106 V·s−1). Application of high scan rate allows rapid acquisition of a voltammogram within several milliseconds and ensures high temporal resolution of this electroanalytical technique. An acquisition rate of 10 Hz is routinely employed.

<span class="mw-page-title-main">Pseudocapacitance</span>

Pseudocapacitance is the electrochemical storage of electricity in an electrochemical capacitor (Pseudocapacitor). This faradaic charge transfer originates by a very fast sequence of reversible faradaic redox, electrosorption or intercalation processes on the surface of suitable electrodes. Pseudocapacitance is accompanied by an electron charge-transfer between electrolyte and electrode coming from a de-solvated and adsorbed ion. One electron per charge unit is involved. The adsorbed ion has no chemical reaction with the atoms of the electrode since only a charge-transfer takes place.

References

  1. Jeuken, Lars J. C. (2016). "Structure and Modification of Electrode Materials for Protein Electrochemistry". Biophotoelectrochemistry: From Bioelectrochemistry to Biophotovoltaics. Advances in Biochemical Engineering/Biotechnology. Vol. 158. Springer, Cham. pp. 43–73. doi:10.1007/10_2015_5011. ISBN   9783319506654. PMID   27506830.
  2. Jeuken, Lars J. C. (2009-09-23). "Electrodes for integral membrane enzymes". Natural Product Reports. 26 (10): 1234–40. doi:10.1039/b903252e. ISSN   1460-4752. PMID   19779638.
  3. 1 2 3 Bard, Allen J. (2001). Electrochemical methods : fundamentals and applications. Faulkner, Larry R., 1944- (2nd ed.). New York: Wiley. ISBN   9780471043720. OCLC   43859504.
  4. 1 2 3 4 Laviron, E. (1979). "General expression of the linear potential sweep voltammogram in the case of diffusionless electrochemical systems". Journal of Electroanalytical Chemistry and Interfacial Electrochemistry. 101 (1): 19–28. doi:10.1016/s0022-0728(79)80075-3.
  5. Chen, Kaisheng; Hirst, Judy; Camba, Raul; Bonagura, Christopher A.; Stout, C. David; Burgess, Barbara. K.; Armstrong, Fraser A. (2000-06-15). "Atomically defined mechanism for proton transfer to a buried redox centre in a protein". Nature. 405 (6788): 814–817. Bibcode:2000Natur.405..814C. doi:10.1038/35015610. ISSN   1476-4687. PMID   10866206. S2CID   4303347.
  6. 1 2 3 Plichon, V.; Laviron, E. (1976). "Theoretical study of a two-step reversible electrochemical reaction associated with irreversible chemical reactions in thin layer linear potential sweep voltammetry". Journal of Electroanalytical Chemistry and Interfacial Electrochemistry. 71 (2): 143–156. doi:10.1016/s0022-0728(76)80030-7.
  7. 1 2 Laviron, E. (1980). "Theoretical study of a 1e, 1H+ surface electrochemical reaction (four-member square scheme) when the protonation reactions are at equilibrium". Journal of Electroanalytical Chemistry and Interfacial Electrochemistry. 109 (1–3): 57–67. doi:10.1016/s0022-0728(80)80106-9.
  8. 1 2 3 4 Léger, Christophe; Bertrand, Patrick (2008-07-01). "Direct Electrochemistry of Redox Enzymes as a Tool for Mechanistic Studies" (PDF). Chemical Reviews. 108 (7): 2379–2438. doi: 10.1021/cr0680742 . ISSN   0009-2665. PMID   18620368.
  9. 1 2 Armstrong, Fraser A.; Camba, Raúl; Heering, Hendrik A.; Hirst, Judy; Jeuken, Lars J. C.; Jones, Anne K.; Le′ger, Christophe; McEvoy, James P. (2000-01-01). "Fast voltammetric studies of the kinetics and energetics of coupled electron-transfer reactions in proteins". Faraday Discussions. 116 (116): 191–203. Bibcode:2000FaDi..116..191A. doi:10.1039/b002290j. ISSN   1364-5498. PMID   11197478.
  10. 1 2 Savéant, Jean-Michel (2006), Elements of Molecular and Biomolecular Electrochemistry: An Electrochemical Approach to Electron Transfer Chemistry, John Wiley & Sons, p. 455, doi:10.1002/0471758078, ISBN   978-0-471-44573-9
  11. Laviron, E. (1972). "Influence of the adsorption of the depolarizer or of a product of the electrochemical reaction on polarographic currents". Journal of Electroanalytical Chemistry and Interfacial Electrochemistry. 35 (1): 333–342. doi:10.1016/s0022-0728(72)80318-8.
  12. Elliott, Sean J; Léger, Christophe; Pershad, Harsh R; Hirst, Judy; Heffron, Kerensa; Ginet, Nicolas; Blasco, Francis; Rothery, Richard A; Weiner, Joel H; Armstrong, Fraser (2002). "Detection and interpretation of redox potential optima in the catalytic activity of enzymes". Biochimica et Biophysica Acta (BBA) - Bioenergetics. 1555 (1–3): 54–59. doi: 10.1016/s0005-2728(02)00254-2 . PMID   12206891.
  13. Fourmond, Vincent; Léger, Christophe (2017). "Modelling the voltammetry of adsorbed enzymes and molecular catalysts" (PDF). Current Opinion in Electrochemistry. 1 (1): 110–120. doi:10.1016/j.coelec.2016.11.002.
  14. del Barrio, Melisa; Sensi, Matteo; Orain, Christophe; Baffert, Carole; Dementin, Sébastien; Fourmond, Vincent; Léger, Christophe (2018). "Electrochemical Investigations of Hydrogenases and Other Enzymes That Produce and Use Solar Fuels" (PDF). Accounts of Chemical Research. 51 (3): 769–777. doi:10.1021/acs.accounts.7b00622. ISSN   0001-4842. PMID   29517230. S2CID   3803863.
  15. Murgida, Daniel H.; Hildebrandt, Peter (2005). "Redox and redox-coupled processes of heme proteins and enzymes at electrochemical interfaces". Physical Chemistry Chemical Physics. 7 (22): 3773–84. Bibcode:2005PCCP....7.3773M. doi:10.1039/b507989f. ISSN   1463-9084. PMID   16358026.
  16. Ash, Philip A.; Vincent, Kylie A. (2016). Biophotoelectrochemistry: From Bioelectrochemistry to Biophotovoltaics. Advances in Biochemical Engineering/Biotechnology. Vol. 158. Springer, Cham. pp. 75–110. doi:10.1007/10_2016_3. ISBN   9783319506654. PMID   27475648.
  17. Kornienko, Nikolay; Ly, Khoa H.; Robinson, William E.; Heidary, Nina; Zhang, Jenny Z.; Reisner, Erwin (May 1, 2019). "Advancing Techniques for Investigating the Enzyme–Electrode Interface". Accounts of Chemical Research. 52 (5): 1439–1448. doi:10.1021/acs.accounts.9b00087. ISSN   0001-4842. PMC   6533600 . PMID   31042353.