mobile-main-menu-button
Menu

Category Archives: Uncategorized

  • WaveDriver Potentiostat: Cell Connections

    wavedriver_tinylogo.jpg
    This article is part of the Pine WaveDriver Potentiostat User's Guide

    This article describes how to make connections from a WaveDriver potentiostat to several different kinds of electrochemical cells. Users should be familiar with the cell cable color coding scheme used by Pine. Examples below include a traditional three-electrode cell, a rotating disk electrode (RDE), a rotating ring-disk electrode, and a simple two electrode configuration.

    Traditional Three-Electrode Cell

    In traditional three-electrode voltammetry, the electrochemical cell consists of three electrodes (counter, reference, and working) placed in an electrolyte solution. The primary current path through the cell is between the counter electrode and the working electrode, and potential is measured between the reference electrode and the working electrode. The photo below shows how to properly connect the potentiostat cell cable to this type of electrochemical cell.

    One of the most important connections is the signal ground (BLACK). Any metal object near the electrochemical cell (Faraday cage, clamp, ring stand, etc.) should be grounded to help prevent noise from interfering with the electrochemical measurement. In the photo below, the signal ground has been connected to the metal clamp holding the electrochemical cell.

    three_electrode_cell.jpg


    In the photo above, the counter electrode (drive) line is the coaxial cable which terminates at a GREEN banana plug, and the working electrode (drive) line is the cable which terminates at a RED banana plug. These are both low-impedance connections which the potentiostat uses to drive current through the cell.

    While driving current through the cell, the potentiostat also measures (senses) the potential between the working electrode and the reference electrode. In the photo above, the cable terminating at an ORANGE banana plug is the working electrode (sense) line, and the cable terminating at a WHITE banana plug is the reference electrode (sense) line. These are both high-impedance connections carrying a negligible amount of current.

    Note that two cables (sense and drive) are connected to the working electrode. Typically, the working electrode sense and drive signals are shorted together at a point very near the electrochemical cell. Failure to connect both of these cables to the working electrode will prevent the potentiostat from properly controlling the electrochemical cell. Shorting these two cables together is a simple matter of pushing the ORANGE banana plug (sense) into the back of the RED banana plug (drive) as shown below.

    wk1_sense_drive_short.jpg

    Working Electrode Sense & Drive Shorted Together near Cell

    When working with a bipotentiostat, the cell cable will provide two pairs of working electrode connections, one pair for the primary working electrode (K1) and another pair for the secondary working electrode (K2). When using a bipotentiostat with a traditional three-electrode cell, always use the primary working electrode connections (ORANGE and RED) to connect to the working electrode. The (unused) secondary working electrode connections (VIOLET and BLUE) should be shorted together but should not be connected to any electrode. Simply lay them to the side on the lab bench as shown in the photo below.

    k2_shorted_sense_and_drive.jpg

    Unused Secondary Working Electrode (K2) Connections
    (sense and drive are shorted together but not connected to any electrode)

    Rotating Disk (RDE) and Rotating Cylinder (RCE) Voltammetry

    In a rotating disk electrode (or a rotating cylinder electrode) experiment, the counter and reference electrode connections are made in the manner described above for traditional three-electrode voltammetry. Connections to the rotating electrode, however, are made via spring-loaded brush contacts which push against the rotating shaft.

    As an example, on the very popular Pine MSR Rotator system, there are two pairs of opposing brushes on either side of the rotating shaft. The upper pair of brush contacts (red) is used to make electrical contact with a rotating disk electrode (RDE) or a rotating cylinder electrode (RCE).

    f4-8.jpg

    To make good contact on opposite sides of the rotating shaft, both of the red brushes (left and right sides) should be used. Use a short banana jumper cable to connect the opposing brushes together (see photo above), and then connect the working electrode sense and drive cables (RED and ORANGE) to the short jumper cable.

    More information about rotating disk voltammetry may be found at the following links:

    Rotation Rate Control

    It is usually desirable to configure the potentiostat so that it can control the rotation rate during RDE and RCE experiments. This is accomplished by connecting the rotator to the potentiostat using a special rotation rate control cable (Pine part number AKCABLE4). One end of this cable is connected to Control Port B on the back panel of the WaveDriver (see photos below).

    wavedriver_rate_control_1.jpg   wavedriver_rate_control_2.jpg   wavedriver_rate_control_3.jpg


    The other end of the control cable is connected to the rotator motor controller. The photos above show the proper connections when using the WaveDriver to control a Pine MSR rotator. The cable has a single banana cable which plugs into the back panel of the motor controller (into the blue MOTOR STOP jack). The cable also has a coaxial cable (with a dual banana adapter) which plugs into the front panel (into the INPUT jacks).

    When connecting the WaveDriver to rotators other than the Pine MSR rotator, it is important to consider the proportionality between the WaveDriver rate control signal (1 RPM/mV) and the proportionality expected by the rotator. This topic is discussed in more detail at the link below:

    https://pineresearch.com/shop/knowledgebase/pine-msr-user-operation/#rotation_rate_control

    More information about rotating disk voltammetry may be found at the following links:

    Rotating Ring-Disk Voltammetry (RRDE)

    In a rotating ring-disk electrode (RRDE) experiment, the counter and reference electrode connections are made in the manner described above for traditional three-electrode voltammetry, and the connection to the rotating disk electrode is made in manner described above for a rotating disk electrode (RDE). The rotator usually has one or more brushes which contact the disk electrode and additional brushes which contact the ring electrode.

    Using the Pine MSR Rotator system as an example, there are two pairs of opposing brushes on either side of the rotating shaft. The upper pair of brush contacts (RED) is used to make electrical contact with the disk electrode while the lower pair of brushes (BLUE) makes contact with the ring electrode.

    f4-9.jpg

    To make good contact on opposite sides of the rotating shaft, both of the red brushes (left and right sides) should be used to contact the disk, and both of the blue brushes should be used to contact the ring. Use short banana jumper cables to connect the opposing brushes together (see photo above). Connect the primary working electrode (K1) cables to the disk electrode (RED and ORANGE cables) connect to the upper pair of brushes). Connect the secondary working electrode (K2) cables to the ring electrode (BLUE and VIOLET cables) to the lower pair of brushes).

    More information about rotating ring-disk voltammetry may be found at the following links:

    Two Electrode Cells

    In some electrochemical experiments there may only two electrodes in the cell. Examples include experiments with ion-selective electrodes (where the open circuit potential is measured between the ISE and a reference electrode) or solid-state experiments where the electrochemical behavior across a single interface is being probed.

    For such two electrode experiments, the counter (GREEN) and reference (WHITE) banana jacks should be shorted together and connected to one of the electrodes (i.e., the reference electrode). In addition the working electrode drive (RED) and sense (ORANGE) banana jacks should be shorted together and connected to the other electrode. An example is shown below where the “cell” is just a simple resistor.

    two_electrode_cell_connection.jpg

    Read More
  • WaveDriver Potentiostat: Cell Cables

    wavedriver_tinylogo.jpg
    This article is part of the Pine WaveDriver Potentiostat User's Guide

    Cell Cable Description

    The front panel of the WaveDriver has a large cell connection port. This port presents several signal, shield, and grounding lines for the working, counter, and reference electrode connections. It is important to understand that some of the signal lines are low impedance DRIVE lines and others are high impedance SENSE lines. In general, the DRIVE lines are used to drive current through the electrochemical cell while the SENSE lines are used to measure potential at various electrodes. Very little charge flows through the high impedance SENSE lines.

    The cell cable breaks out the various cell port connections to shielded coaxial cables which terminate at stackable banana plugs. Alligator clipsthat slide on to the banana plugs are included. The banana plugs are color coded according to the standard PRI electrode color scheme shown below.

    pri_color_code.jpg

    Color Code Key for Electrode Connections

    acp2e01_cable_kit.jpg acp2e05_cable_kit.jpg

    WaveDriver 20 Bipotentiostat Cell Cable (left) and WaveDriver 10 Potentiostat Cell Cable (right)

    Front Panel Cell Port

    This cell port on the front panel is a female “combination D-sub connector” containing seven (7) coaxial connectors and seventeen (17) pin sockets. The coaxial connectors each have driven shields to protect the inner signal pin, and only six of the seven coaxial connectors are used. The pinout diagram for the coaxial connectors is shown below. Note that one of the pin sockets provides a connection to the analog ground.

    cell_port_pinout.jpg


    The combination D-sub female socket (female) on the front panel is a “24W7” type connector. This means it has 24 total connections with seven of these twenty-four connections being coaxial connections. It is designed to mate with the plug (male) on the cell cable. Pine offers cell cables that fit into this socket, but customers wishing to make their own cable can consult major electronics suppliers to obtain the necessary parts and tools.

    Ordering Information

    Part numbers for various items in this cell cable kit are provided below.

    Part Number – Description
    ACP2E05 – WaveDriver 10 Cell Cable (counter, reference, working sense & drive, signal ground)
    ACP2E01 – WaveDriver 20 Cell Cable (same as ACP2E05 but also includes sense & drive lines for secondary working electrode)
    THCLIP – This is the alligator clip that slides on to a banana plug

    Customers wishing to purchase parts and tools for making a custom cell cable may find the following link useful:

    http://www.google.com/?q=3024W7PCM99A10X (a general search for related parts)

    Read More
  • WaveDriver Potentiostat: Cell Connections

    wavedriver_tinylogo.jpg
    This article is part of the Pine WaveDriver Potentiostat User's Guide

    This article describes how to make connections from a WaveDriver potentiostat to several different kinds of electrochemical cells. Users should be familiar with the cell cable color coding scheme used by Pine. Examples below include a traditional three-electrode cell, a rotating disk electrode (RDE), a rotating ring-disk electrode, and a simple two electrode configuration.

    Traditional Three-Electrode Cell

    In traditional three-electrode voltammetry, the electrochemical cell consists of three electrodes (counter, reference, and working) placed in an electrolyte solution. The primary current path through the cell is between the counter electrode and the working electrode, and potential is measured between the reference electrode and the working electrode. The photo below shows how to properly connect the potentiostat cell cable to this type of electrochemical cell.

    One of the most important connections is the signal ground (BLACK). Any metal object near the electrochemical cell (Faraday cage, clamp, ring stand, etc.) should be grounded to help prevent noise from interfering with the electrochemical measurement. In the photo below, the signal ground has been connected to the metal clamp holding the electrochemical cell.

    three_electrode_cell.jpg


    In the photo above, the counter electrode (drive) line is the coaxial cable which terminates at a GREEN banana plug, and the working electrode (drive) line is the cable which terminates at a RED banana plug. These are both low-impedance connections which the potentiostat uses to drive current through the cell.

    While driving current through the cell, the potentiostat also measures (senses) the potential between the working electrode and the reference electrode. In the photo above, the cable terminating at an ORANGE banana plug is the working electrode (sense) line, and the cable terminating at a WHITE banana plug is the reference electrode (sense) line. These are both high-impedance connections carrying a negligible amount of current.

    Note that two cables (sense and drive) are connected to the working electrode. Typically, the working electrode sense and drive signals are shorted together at a point very near the electrochemical cell. Failure to connect both of these cables to the working electrode will prevent the potentiostat from properly controlling the electrochemical cell. Shorting these two cables together is a simple matter of pushing the ORANGE banana plug (sense) into the back of the RED banana plug (drive) as shown below.

    wk1_sense_drive_short.jpg

    Working Electrode Sense & Drive Shorted Together near Cell

    When working with a bipotentiostat, the cell cable will provide two pairs of working electrode connections, one pair for the primary working electrode (K1) and another pair for the secondary working electrode (K2). When using a bipotentiostat with a traditional three-electrode cell, always use the primary working electrode connections (ORANGE and RED) to connect to the working electrode. The (unused) secondary working electrode connections (VIOLET and BLUE) should be shorted together but should not be connected to any electrode. Simply lay them to the side on the lab bench as shown in the photo below.

    k2_shorted_sense_and_drive.jpg

    Unused Secondary Working Electrode (K2) Connections
    (sense and drive are shorted together but not connected to any electrode)

    Rotating Disk (RDE) and Rotating Cylinder (RCE) Voltammetry

    In a rotating disk electrode (or a rotating cylinder electrode) experiment, the counter and reference electrode connections are made in the manner described above for traditional three-electrode voltammetry. Connections to the rotating electrode, however, are made via spring-loaded brush contacts which push against the rotating shaft.

    As an example, on the very popular Pine MSR Rotator system, there are two pairs of opposing brushes on either side of the rotating shaft. The upper pair of brush contacts (red) is used to make electrical contact with a rotating disk electrode (RDE) or a rotating cylinder electrode (RCE).

    f4-8.jpg

    To make good contact on opposite sides of the rotating shaft, both of the red brushes (left and right sides) should be used. Use a short banana jumper cable to connect the opposing brushes together (see photo above), and then connect the working electrode sense and drive cables (RED and ORANGE) to the short jumper cable.

    More information about rotating disk voltammetry may be found at the following links:

    Rotation Rate Control

    It is usually desirable to configure the potentiostat so that it can control the rotation rate during RDE and RCE experiments. This is accomplished by connecting the rotator to the potentiostat using a special rotation rate control cable (Pine part number AKCABLE4). One end of this cable is connected to Control Port B on the back panel of the WaveDriver (see photos below).

    wavedriver_rate_control_1.jpg   wavedriver_rate_control_2.jpg   wavedriver_rate_control_3.jpg


    The other end of the control cable is connected to the rotator motor controller. The photos above show the proper connections when using the WaveDriver to control a Pine MSR rotator. The cable has a single banana cable which plugs into the back panel of the motor controller (into the blue MOTOR STOP jack). The cable also has a coaxial cable (with a dual banana adapter) which plugs into the front panel (into the INPUT jacks).

    When connecting the WaveDriver to rotators other than the Pine MSR rotator, it is important to consider the proportionality between the WaveDriver rate control signal (1 RPM/mV) and the proportionality expected by the rotator. This topic is discussed in more detail at the link below:

    https://pineresearch.com/shop/knowledgebase/pine-msr-user-operation/#rotation_rate_control

    More information about rotating disk voltammetry may be found at the following links:

    Rotating Ring-Disk Voltammetry (RRDE)

    In a rotating ring-disk electrode (RRDE) experiment, the counter and reference electrode connections are made in the manner described above for traditional three-electrode voltammetry, and the connection to the rotating disk electrode is made in manner described above for a rotating disk electrode (RDE). The rotator usually has one or more brushes which contact the disk electrode and additional brushes which contact the ring electrode.

    Using the Pine MSR Rotator system as an example, there are two pairs of opposing brushes on either side of the rotating shaft. The upper pair of brush contacts (RED) is used to make electrical contact with the disk electrode while the lower pair of brushes (BLUE) makes contact with the ring electrode.

    f4-9.jpg

    To make good contact on opposite sides of the rotating shaft, both of the red brushes (left and right sides) should be used to contact the disk, and both of the blue brushes should be used to contact the ring. Use short banana jumper cables to connect the opposing brushes together (see photo above). Connect the primary working electrode (K1) cables to the disk electrode (RED and ORANGE cables) connect to the upper pair of brushes). Connect the secondary working electrode (K2) cables to the ring electrode (BLUE and VIOLET cables) to the lower pair of brushes).

    More information about rotating ring-disk voltammetry may be found at the following links:

    Two Electrode Cells

    In some electrochemical experiments there may only two electrodes in the cell. Examples include experiments with ion-selective electrodes (where the open circuit potential is measured between the ISE and a reference electrode) or solid-state experiments where the electrochemical behavior across a single interface is being probed.

    For such two electrode experiments, the counter (GREEN) and reference (WHITE) banana jacks should be shorted together and connected to one of the electrodes (i.e., the reference electrode). In addition the working electrode drive (RED) and sense (ORANGE) banana jacks should be shorted together and connected to the other electrode. An example is shown below where the “cell” is just a simple resistor.

    two_electrode_cell_connection.jpg

    Read More
  • Rotating Electrode Theory


    msr_small.jpg This article is adapted from Section 10 of the User Manual for the Pine Modulated Speed Rotator (MSR).

    Related Links: MSR User Manual, References Cited in This Article

    Hydrodynamic Voltammetry Theory


    10.1 – Forced Convection

    The current signal recorded during an electrochemical experiment is easily influenced or disturbed by the convection of various molecules and ions due to bulk movement of the solution. Proper interpretation of the current signal must accurately account for any contributions (desired or undesired) from solution convection. Thus, the control of solution movement is a critical part of any electrochemical experiment design, and the issue of convection cannot be ignored. Two opposing approaches are typically used to address the convection issue. At one extreme, an experiment can be conducted in a quiescent solution, so that convection makes little or no contribution to the observed current. The opposite extreme involves forced convection, where the solution is actively stirred or pumped in a controlled manner.

    At first glance, it may seem that the simplest and most obvious way to account for convection is to try to eliminate it entirely by using a quiescent (non-moving) solution. This is the approach used in many popular electroanalytical techniques(1) (including cyclic voltammetry, chronoamperometry, square wave voltammetry, and differential pulse voltammetry). The timescale for these methods is generally less than 30 seconds, and on such short timescales, the influence of convection in an unstirred solution is generally negligible. On longer timescales, however, even unstirred solutions are prone to convective interference from thermal gradients and subtle environmental vibrations.

    For long duration (steady-state) experiments, convection is unavoidable, so actively forcing(2) the solution to move in a well-defined and controlled manner is the preferred approach. An entire family of electroanalytical methods (broadly categorized as hydrodynamic voltammetry) couples precise control of solution flow with rigorous mathematical models defining the flow. Some of the many examples of hydrodynamic voltammetry include placing an electrode in a flow cell,(3) firing a jet of solution at an electrode target,(4-5) embedding an electrode in a microfluidic channel,(6) vibrating a wire-shaped electrode,(7) subjecting the solution to ultrasonication,(8) and rotating the electrode.(2,9-14)

    By far the most popular and widely used hydrodynamic methods are those that involve a rotating electrode. The rotating electrode geometries most amenable to mathematical modeling are the rotating disk electrode (RDE),(9-14) the rotating ring-disk electrode (RRDE),(15-26) and the rotating cylinder electrode (RCE).(27-32) Researchers take advantage of the stable, steady-state laminar flow conditions adjacent to an RDE or RRDE to carefully gather information about electrode reaction kinetics.(13,14,21,26,33-43) In contrast, the relatively chaotic and turbulent conditions adjacent to an RCE are exploited by corrosion scientists(44-69) wishing to mimic flow-induced pipeline corrosion conditions in the laboratory. Development of the RDE and RRDE as routine analytical tools has largely been carried out by the community of academic electroanalytical chemists, while the RCE has primarily been a tool used by the corrosion and electroplating industries.

    10.2 – Half Reactions

    Regardless of the rotating electrode geometry being used, the common theme is that an ion or molecule is being conveyed to the electrode surface, and upon arrival, it is either oxidized or reduced depending upon the potential applied to the rotating electrode. If a sufficiently positive potential is applied to the electrode, then the molecules (or ions) tend to be oxidized, and conversely, if a sufficiently negative potential is applied to the rotating electrode, the molecules (or ions) tend to be reduced.

    Reduction at a rotating electrode implies that electrons are being added to the ion or molecule by flowing out of the electrode and into the solution. A current travelling in this direction is said to be a cathodic current. The general form of a reduction half-reaction occurring at an electrode may be written as follows:

    O + ne^- \rightarrow R

    where  R represents the reduced form of the molecule (or ion),  O represents the oxidized form of the molecule (or ion), and  n is the total number of electrons added to the molecule (or ion) when it is converted from the oxidized form ( O ) to the reduced form ( R ).

    Oxidation at a rotating electrode implies that electrons are being removed from an ion or molecule and are travelling out of the solution and into the electrode. A current travelling in this direction is said to be an anodic current, and the oxidation occurring at the electrode can be represented by the following redox half reaction,

    R \rightarrow O+ne^-

    Given that electrochemical half reactions can occur in either direction, they are often written using chemical equilibrium notation* as follows:

    O + ne^- {\rightleftharpoons} R

    Each half reaction has an associated standard electrode potential ( E^{\circ} ) which is a thermodynamic quantity related to the free energy associated with the equilibrium. Like many other standard thermodynamic quantities, the standard electrode potential corresponds to a given standard state. The standard state corresponds to a thermodynamic system where the chemical activities of  O and  R are unity (i.e., when all solution concentrations are  1 \; mol/L , all gases are present at  1 \; atm partial pressure, and other materials are present as pure phases with unity activity).

    To account for the (likely) possibility of non-unity activities, the Nernst equation (see below) can be used to express the equilibrium electrode potential ( E_{NERNSTIAN} ) in terms of the actual activities.

    E_{NERNSTIAN} = E^{\circ} + \frac{RT}{nF}ln\left[{\frac{{\alpha}_O}{{\alpha}_R}}\right]

    where  F is the Faraday constant ( 96485 \; C/mol ),  R is the ideal gas constant ( 8.3145 \; J \; mol^{-1}\; K^{-1} ), and  T is the absolute temperature ( K ). Usually, the activities of molecules or ions dissolved in solution are assumed to be the same as their molar concentrations, so the Nernst Equation is often written as follows

    E_{NERNSTIAN} = E^{\circ} + \frac{RT}{nF}ln\left[{\frac{C_O}{C_R}}\right]

    where  C_O and  C_R are the concentrations of the dissolved molecules or ions in the oxidized and reduced forms, respectively, at the surface of the electrode. Note that any liquid or solid phase materials at the electrode surface (such as the solvent or the electrode itself) have unity activity and thus do not appear in the Nernst equation.

    This half reaction at an electrode can be driven in the cathodic (reducing) direction by applying a potential to the electrode ( E_{APPLIED} ) which is more negative than the equilibrium electrode potential ( E_{APPLIED}<E_{NERNSTIAN} ). Conversely, the half reaction can be driven in the oxidizing (anodic) direction by applying a potential more positive than the equilibrium electrode potential (E_{APPLIED}>E_{NERNSTIAN}).

    —–

    * By convention, redox half reactions are generally tabulated in textbooks and other reference works as reduction reactions (with the oxidized form on the left side and the reduced form on the right side, as shown above), but it is understood that the reaction may occur in either direction depending upon the potential applied to the electrode.

    10.3 – Voltammetry

    The term voltammetry refers broadly to any method where the electrode potential is varied while the current is measured.(1-2) The terminology associated with voltammetry varies across different industries and academic disciplines, but the underlying principles of all voltammetric techniques are very similar.

    The most common form of voltammetry involves sweeping the electrode potential from an initial value to a final value at a constant rate. When working in the context of electroanalytical chemistry with a non-rotating electrode, this technique is called linear sweep voltammetry (LSV). In the context of corrosion science, this kind of technique is usually called linear polarization resistance (LPR) or a Tafel analysis. The term cyclic voltammetry (CV) refers to a method where the electrode potential is swept repeatedly back-and-forth between two extremes.

    When working with a rotating electrode, it is common to further specify the kind of electrode being used as part of the technique name, such as rotating disk voltammetry, rotating ring-disk voltammetry, or rotating cylinder voltammetry. In each of these techniques, the rotation rate is held constant as the electrode is swept from one potential to another potential at a constant sweep rate. In electroanalytical chemistry, the potential sweep usually spans at least  200 \; mV on either side of the standard electrode potential, and rotation rates are usually between  100 \; RPM and  2400 \; RPM . However, in the context of a corrosion study, the potential sweep may span a much narrower range ( 50 \; mV ) using a slower sweep rate (less than  5 \; mV/sec ) with an emphasis on higher rotation rates.

    As an example, consider a solution that initially contains only the oxidized form of a molecule or ion. A rotating electrode is placed in this solution and is initially poised at a potential that is  200 \; mV more positive than the standard potential. At this potential, there is little or no current because there is nothing to oxidize (the molecule or ion is already oxidized), and the potential is not (yet) negative enough to cause any appreciable reduction of the molecule or ion.

    Next, the electrode potential is slowly ( 20 \; mV/sec ) swept in the negative (cathodic) direction (see Figure 10.1, left). As the applied potential approaches the standard electrode potential, a cathodic current is observed (see Figure 10.1, right). The cathodic current continues to increase as the potential moves past the standard electrode potential towards more negative potentials.

    The current eventually reaches a maximum value (limiting current) once the applied potential is sufficiently negative relative to the standard electrode potential. At such a negative potential, any oxidized form of the molecule or ion ( O ) that reaches the surface of the electrode is immediately converted to the reduced form ( R ) as shown below.

    O + ne^- \rightarrow R

    The observed cathodic current is the result of electrons flowing out of the electrode and into the solution. The rate of electron flow is limited only by how fast the oxidized form ( O ) can arrive at the electrode surface. The maximum current observed in this circumstance is called the cathodic limiting current ( i_{LC} ).

    Whenever an observed current is limited only by the rate at which material arrives at the electrode surface, the current is said to be mass transport limited. When working with a rotating electrode, the rate of mass transport is related to the rotation rate of the electrode. Rotating the electrode at a faster rate increases the rate at which material arrives at the electrode surface. Thus, the limiting current increases with increasing rotation rate. Experiments involving a rotating electrode are designed to purposefully exploit this fundamental relationship between the rotation rate and the limiting current.

    f10-1a.jpg   f10-1b.jpg

    Figure 10.1: Response to a Potential Sweep (Cathodic) from a Solution Initially Containing only the Oxidized Form (\textbf{O}) with no Reduced Form (\textbf{R})

    The cathodic sweep experiment described above (see Figure 10.1) applies to the case where the solution initially contains only the oxidized form ( O ) of the molecule or ion being studied. The opposite case yields similar results. Consider a solution that initially contains only the reduced form ( R ) of the molecule or ion being studied. The rotating electrode is initially poised at a potential that is about  200 \; mV more negative than the standard potential. At this potential, there is little or no current because there is nothing to reduce (the molecule or ion is already reduced), and the potential is not (yet) positive enough to cause any appreciable oxidation of the molecule or ion.

    Next, the electrode potential is slowly swept in the positive (anodic) direction (see Figure 10.2, left) and an anodic current is observed (see Figure 10.2, right). The anodic current eventually reaches a maximum value when the potential is sufficiently positive relative to the standard electrode potential. At this point, any of the reduced form ( R ) that reaches the electrode surface is immediately converted to the oxidized form ( O ).

    R \rightarrow O + ne^-

    The observed current is the result of electrons flowing into the electrode. The maximum current observed is called the anodic limiting current ( i_{LA} ).

    f10-2a.jpg   f10-2b.jpg

    Figure 10.2: Response to a Potential Sweep (Anodic) from a Solution Initially Containing only the Reduced Form (\textbf{R}) with no Oxidized Form (\textbf{O})

    10.3.1 – Voltammogram Plotting Conventions

    The two streams of data recorded during a voltammetry experiment are the potential vs. time and the current vs. time. Rather than plot these two streams separately (as shown in Figure 10.3, top), it is more common to plot current vs. potential (as shown in Figure 10.3, bottom). Such a plot is called a voltammogram.

    f10-2a.jpg  
    f10-2b.jpg
    f10-3c.jpg

    Figure 10.3: A Voltammogram is a Plot of Current versus Potential

    Although most electroanalytical researchers agree that current should be plotted along the vertical axis and potential should be plotted along the horizontal axis, there is not widespread agreement as to the orientation (direction) for each axis. Some researchers plot positive (anodic, oxidizing) potentials toward the right while others plot negative (cathodic, reducing) potential toward the right (as per classical polarography tradition). Furthermore, some researchers plot anodic (oxidizing) current upward along the vertical axis, while others plot cathodic (reducing) current in the upward direction.

    This means there are four possible conventions for plotting a voltammogram, and one should always take a moment to ascertain the orientation of the axes before interpreting a voltammogram. Fortunately, of the four possible ways to plot a voltammogram, only two are commonly used. The older tradition (based on classical polarography) plots cathodic current upwards along the vertical axis and negative (cathodic, reducing) potentials toward the right along the horizontal axis. A complex voltammogram involving four different limiting currents (see Figure 10.4, left) illustrates this convention, which is sometimes called the “North American” convention.

    North American Convention

    f10-4a.jpg

    European Convention

    f10-4b.jpg

    Figure 10.4: Two Popular Voltammogram Plotting Conventions

    The same data may be plotted using the “European” convention (see Figure 10.4, right). This convention plots anodic currents upward along the vertical axis and more positive (anodic, oxidizing) potentials to the right along the horizontal axis. The European convention is more readily understood by those outside the electroanalytical research community (because positive values are plotted to the right along the horizontal axis).

    The European convention is used throughout the remainder of this article. Note that this choice also implies a mathematical sign convention for the current. Specifically, positive current values are considered anodic, and negative current values are considered cathodic in this document. This sign convention is somewhat arbitrary, and electrochemical data processing software available from various manufacturers may or may not use this sign convention.

    10.3.2 – Measuring Limiting Currents

    The theoretical voltammetric response from a rotating electrode is a symmetric sigmoid-shaped wave (like the ideal voltammograms shown in Figure 10.3 and Figure 10.4). A perfect sigmoid has a flat baseline current before the wave and a flat limiting current plateau after the wave. The height of the wave (as measured from the baseline current to the limiting current plateau) is the mass-transport limited current.

    In actual “real world” experiments, the wave may be observed on top of a background current, and furthermore, the background current may be slightly sloped (see Figure 10.5). This (undesired) background current may be due to interference from oxidation or reduction of impurities or of the solvent itself. The background current may also be due to capacitive charging and discharging of the ionic double-layer that forms next to the polarized electrode surface.

    f10-5a.jpg   f10-5b.jpg

    Figure 10.5: Sloping Backgrounds in Voltammograms

    When attempting to measure the (desired) Faradaic mass-transport limited current at a rotating electrode, it is often necessary to account for the (undesired, possibly sloping) background current. If the background current has a constant slope across the entire voltammogram, then it is fairly easy to extrapolate the sloping baseline to a point underneath the limiting current plateau (see Figure 10.5, left). The limiting current is measured as the (vertical) distance between the plateau and the extrapolated baseline. In voltammograms where there is more than one wave, the plateau for the first wave is used as the baseline for the second wave (see  i_{LA2} in Figure 10.5, left).

    In some cases, the slope of the background current is not constant across the entire voltammogram. That is, the slope of the baseline leading up to the wave can be different than the slope of the plateau after the wave. It can be very difficult to discern exactly where to measure the limiting current along such a voltammogram. One approach is to extrapolate the baseline forward through the wave and also extrapolate the plateau backward through the wave. Then, the limiting current is measured as the vertical distance between the baseline and plateau at a point corresponding to the center of the voltammogram (see  i_{LA} in Figure 10.5, right).

    Finally, it should be noted that when the oxidized form ( O ) and the reduced form ( R ) of a molecule or ion are both present in a solution at the same time, the voltammogram is likely to exhibit both a cathodic and an anodic limiting current (see Figure 10.6). It can be very difficult to measure the limiting current properly in this case, especially if there is also a sloping background current. For this reason, most experiments with rotating electrodes are conducted in solutions where only one form of the molecule or ion is initially present.

    f10-6.jpg

    Figure 10.6: Voltammogram for a Solution Containing Both  \textbf{O} and  \textbf{R}

    10.4 – Rotating Disk Electrode (RDE) Theory

    The general theory describing mass transport at a rotating disk electrode (RDE) was developed by Benjamin Levich at the Institute of Electrochemistry at the Academy of Sciences of the USSR. Levich described the theory in his landmark book, Physicochemical Hydrodynamics, originally published in Russian in 1952. Ten years later, Levich’s book was translated(9) from Russian to English, and the RDE became more widely known(11) to western researchers. In the early 1960’s, Stanley Bruckenstein(10) at the University of Minnesota (and his students Dennis Johnson and Duane Napp) and Ronnie Bell(12) at Oxford University (and his student John Albery) began working with rotating electrodes. Subsequent generations of researchers expanded on this initial work, and the rotating disk electrode has since grown into a mature tool for probing electrochemical reaction kinetics.(13)

    The laminar flow at a rotating disk electrode conveys a steady stream of material from the bulk solution to the electrode surface. While the bulk solution far away from the electrode remains well-stirred by the convection induced by rotation, the portion of the solution nearer to the electrode surface tends to rotate with the electrode. Thus, if the solution is viewed from the frame of reference of the rotating electrode surface, then the solution appears relatively stagnant. This relatively stagnant layer is known as the hydrodynamic boundary layer, and its thickness ( {\delta}_H ) can be approximated,

    {\delta}_H = 3.6\left({\frac{v}{\omega}}\right)^{1/2}

    in terms of the kinematic viscosity of the solution ( {\nu} ) and the angular rotation rate ( {\omega} = 2{\pi}f/60 ), where f is the rotation rate in revolutions per minute). In an aqueous solution at a moderate rotation rate (~1000 \; RPM ), the stagnant layer is approximately  300 to  400 \; {\mu}m thick.

    Net movement of material to the electrode surface can be described mathematically by applying general convection-diffusion concepts from fluid dynamics. Mass transport of material from the bulk solution into the stagnant layer occurs by convection (due to the stirring action of the rotating electrode). But after the material enters the stagnant layer and moves closer to the electrode surface, convection becomes less important and diffusion becomes more important. Indeed, the final movement of an ion or molecule to the electrode surface is dominated by diffusion across a very thin layer of solution immediately adjacent to the electrode known as the diffusion layer.

    The diffusion layer is much thinner than the hydrodynamic layer. The diffusion layer thickness ( {\delta}_f ) can be approximated as follows,

    {\delta}_f = 1.61 (D_f)^{1/3}v^{1/6}{\omega}^{-1/2}

    in terms of the diffusion coefficient ( D_f ) of the molecule or ion. For a molecule or ion with a typical diffusion coefficient ( D_f \; {\cong} \; 10^{-5} \; cm^2/s ) in an aqueous solution, the diffusion layer is about twenty times thinner than the stagnant layer ( {\delta}_f \; {\cong} \; 0.05{\delta}_H ).

    The first mathematical treatment of convection and diffusion towards a rotating disk electrode was given by Levich. Considering the case where only the oxidized form of a molecule (or ion) of interest is initially present in the electrochemical cell, the cathodic limiting current ( i_{LC} ) observed at a rotating disk electrode is given by the Levich equation,(2,9)

    i_{LC} = 0.0620nFAD^{2/3}v^{-1/6}C_O{\omega}^{1/2}

    in terms of the concentration ( C_O ) of the oxidized form in the solution, the Faraday constant ( F = 96485 \; C/mol ), the electrode area ( A ), the kinematic viscosity of the solution ( {\nu} ), the diffusion coefficient ( D ) of the oxidized form, and the angular rotation rate ( {\omega} ). Alternatively, when the solution initially contains only the reduced form, the Levich equation for the anodic limiting current ( i_{LA} ) can be written as

    i_{LA} = 0.620nFAD^{2/3}{\nu}^{-1/6}C_R{\omega}^{1/2}

    where the concentration term ( C_R ) is for the reduced form rather than the oxidized form.

    10.4.1 – Levich Study

    A Levich Study is a common experiment performed using a rotating disk electrode in which a series of voltammograms is acquired over a range of different rotation rates. For a simple electrochemical system where the rate of the half reaction is governed only by mass transport to the electrode surface, the overall magnitude of the voltammogram should increase with the square root of the rotation rate (see Figure 10.7, left).

    f10-7l.jpg   f10-7r.jpg

    Figure 10.7: Levich Study – Voltammograms at Various Rotation Rates

    The currents measured during a Levich study are usually plotted against the square root of the rotation rate on a graph called a Levich plot. As predicted by the Levich equation, the limiting current (see red circles on Figure 10.7, right) increases linearly with the square root of the rotation rate (with a slope of  0.620 n F A D^{2/3}{\nu}^{-1/6}C ) and the line intercepts the vertical axis at zero. It is common to choose a set of rotation rates that are multiples of perfect squares (such as  100 ,  400 ,  900 ,  1600 \; RPM, etc.) to facilitate construction of this plot.

    If the electrochemical half-reaction observed during a Levich study is a simple and reversible half reaction (with no complications due to sluggish kinetics or coupled chemical reactions), then the shapes of the mass-transport controlled voltammograms will be sigmoidal regardless of the rotation rate. This means that the current observed at any given potential along the voltammogram will vary linearly with the square root of the rotation rate (see Figure 10.7, right). But, it is important to remember that the Levich equation only applies to the limiting current, not to the currents along the rising portion of the sigmoid.

    Limiting Current

    f10-8l.jpg

    Levich Plot

    f10-8m.jpg

    Koutecky-Levich Plot

    f10-8r.jpg

    Figure 10.8: Levich Study – Limiting Current versus Rotation Rate

    Because the Levich equation only applies to the limiting current, the results from a Levich experiment are typically presented as a simple plot of the limiting current versus the square root of the rotation rate (see Figure 10.8, center). An alternate method of presenting the data from a Levich study is based on a rearrangement of the Levich equation in terms of the reciprocal current.

    \frac{1}{i_L} = \left({\frac{1}{0.620nFAD^{2/3}{\nu}^{-1/6}C}}\right){\omega}^{-1/2}

    A plot of reciprocal current versus the reciprocal square root of the angular rotation rate (see Figure 10.8, right) is called a Koutecky-Levich(2,14) plot. Again, for a simple and reversible half reaction with no complications the data fall along a straight line that intercepts the vertical axis at zero. If the line intercepts the vertical axis above zero, however, this is a strong indication that the half-reaction is limited by sluggish kinetics rather than by mass transport.

    10.4.2 – Koutecky-Levich Analysis

    When the rate of a half reaction occurring at an electrode surface is limited by a combination of mass transport and sluggish kinetics, it is often possible to use a rotating disk electrode to elucidate both the mass transport parameters (such as the diffusion coefficient) and the kinetic parameters (such as the standard rate constant,  k^{\circ} ) from a properly designed Levich study. A full treatment of this kind of analysis(14) is beyond the scope of this document, but the following is a general description of how to extract kinetic information from a set of rotating disk voltammograms.

    When the electron transfer process at an electrode surface exhibits sluggish kinetics, the voltammogram appears stretched out along the potential axis and the shape of the sigmoidal wave is slightly distorted. Comparing a set of voltammograms with facile kinetics (see Figure 10.7) with a set of voltammograms with sluggish kinetics (see Figure 10.9), the mass transport limited current plateau (marked by red circles in each figure) is shifted further away from the standard electrode potential ( E^{\circ} ) when there are slow kinetics. Stated another way, when a sluggish redox half reaction is studied with a rotating disk electrode, a larger overpotential must be applied to the electrode to overcome the sluggish kinetics and reach the mass transport limited current.

    f10-9a.jpg   f10-9b.jpg

    Figure 10.9: Koutecky Levich Study – Voltammograms Revealing Sluggish Kinetics

    This distortion of the ideal sigmoidal shape of the voltammogram can be exploited as a way to measure the standard rate constant ( k^{\circ} ). The general approach is to acquire a set of voltammograms at different rotation rates (i.e., perform a Levich study) and then plot the reciprocal current (sampled at particular locations along the rising portion of each voltammogram) on a Koutecky-Levich Plot. In the example provided (see Figure 10.9, left), the current was sampled at two locations along the rising portion of the voltammograms (at  0 and  50 \; mV \; vs \; E^{\circ} , marked with blue triangles and purple squares) and at one location on the limiting current plateau (at  350 \; mV \; vs \; E^{\circ} , marked with red circles). A linear relationship is evident (see Figure 10.9, right) when these sampled currents are plotted on a Koutecky-Levich plot.

    For the set of currents sampled on the limiting current plateau (red circles), an extrapolation back to the vertical axis (i.e., to infinite rotation rate) yields a zero intercept. This is the identical result obtained for a facile half-reaction (see Figure 10.8, right) because these currents are sampled at a high enough overpotential that there are no kinetic limitations. Only mass transport limits the current, and the usual Levich behavior applies.

    However, for the two sets of currents sampled on the rising portion of the voltammogram (see Figure 10.9, blue triangles and purple squares), the extrapolation back to the vertical axis yields non-zero intercepts. This non-zero intercept indicates a kinetic limitation, meaning that even if mass transport were infinite (i.e., infinite rotation rate), the rate of the half-reaction would still be limited by the slow kinetics at the electrode surface.

    The linear portion of the data on a Koutecky-Levich plot is described by the Koutecky-Levich equation (below).

    \frac{1}{i} = \frac{1}{i_k} + \left({\frac{1}{0.620nFAD^{2/3}{\nu}^{-1/6}C}}\right){\omega}^{-1/2}

    Plotting the reciprocal current ( 1/i ) against the reciprocal square root of the angular rotation rate ( {\omega}^{-1/2} ) yields a straight line with an intercept equal to the reciprocal kinetic current ( 1/i_k ). The kinetic current is the current that would be observed in the absence of any mass transport limitations. By measuring the kinetic current at a variety of different overpotentials along the voltammogram, it is possible to determine the standard rate constant for the electrochemical half reaction.

    Further details regarding Koutecky-Levich theory, including various forms of the Koutecky-Levich equation which pertain to different electrochemical processes, can be found in the literature.(14)

    10.5 – Rotating Ring-Disk Electrode (RRDE) Theory

    In 1958, Russian electrochemist Alexander Frumkin suggested the idea of placing a concentric ring electrode around the rotating disk electrode.(15) His colleague, Lev Nekrasov, supervised construction of the world’s first rotating ring-disk electrode (RRDE) apparatus.(16-19) At the same time, Benjamin Levich and Yuri Ivanov began working on a theoretical description of solution flow at the RRDE. The four Russian researchers published their initial findings in 1959, and their work caught the attention of both Stanley Bruckenstein at the University of Minnesota and John Albery at Oxford University. Bruckenstein travelled to Moscow to learn more about the RRDE,(20) and after he returned home in 1965, Albery joined Bruckenstein’s research group (along with Dennis Johnson and Duane Napp). The experimental and mathematical work performed by these four researchers at Minnesota generated a significant series of papers about the RRDE(21-26) and placed the new technique on a firm theoretical foundation. Albery returned to Oxford and (working with his student Michael Hitchman) drew these theoretical papers together in a seminal volume titled Ring-Disc Electrodes.(21)

    The overall flow pattern at the RRDE initially brings molecules and ions to the disk electrode. After encountering the disk electrode, the subsequent outward radial flow carries a fraction of these molecules or ions past the surface of the ring electrode. This flow pattern allows products generated (upstream) by the half reaction at the disk electrode to be detected as they are swept (downstream) past the ring electrode.

    Two of the key parameters which characterize a given ring-disk geometry are the collection efficiency(23) and the transit time. The collection efficiency is the fraction of the material from the disk which subsequently flows past the ring electrode. It can be expressed as a fraction between  \text{0.0} and  \text{1.0} or as a percentage. Typical ring-disk geometries have collection efficiencies between  20\% and  30\% . The transit time is a more general concept indicating the average time required for material at the disk electrode to travel across the gap between the disk and the ring electrode. Obviously, the transit time is a function of both the gap distance and the rotation rate.

    10.5.1 – Theoretical Computation of the Collection Efficiency

    The theoretical collection efficiency can be computed(2,23) from the three principle diameters describing the RRDE geometry: the disk outer diameter ( d_1 ), the ring inner diameter ( d_2 ), and the ring outer diameter ( d_3 ). This somewhat tedious computation is made easier by normalizing the ring diameters with respect to the disk diameter as follows:

    {\sigma}_{OD} = \frac{d_3}{d_1}

    {\sigma}_{ID} = \frac{d_2}{d_1}

    Three additional quantities are defined in terms of the normalized diameters as follows:

    {\sigma}_{A} = {\sigma}_{ID}^3 - 1

    {\sigma}_{B} = {\sigma}_{OD}^3 - {\sigma}_{ID}^3

    {\sigma}_{C} = \frac{{\sigma}_{A}}{{\sigma}_{B}}

    If a complex function,  G(x) , is defined as follows,

    G(x) = \frac{1}{4} + \left(\frac{\sqrt{3}}{4{\pi}}\right) ln \left[{\frac{(x^{1/3} + 1)^3}{x+1}}\right] + \frac{3}{2{\pi}} {\arctan}\left[{\frac{2x^{1/3}-1}{\sqrt{3}}}\right]

    then the theoretical collection efficiency ( N_{theoretical} ) for a rotating ring disk electrode is given by the following equation:

    N_{theoretical} = 1-{\sigma}_{OD}^2+{\sigma}_B^{2/3}-G({\sigma}_C)-{\sigma}_B^{2/3}G({\sigma}_A)+{{\sigma}_{OD}^2}G({\sigma}_C{\sigma}_{OD}^3)

    10.5.2 – Empirical Measurement of the Collection Efficiency

    Direct computation of the theoretical collection efficiency is possible using the above relationships if the actual machined dimensions of the disk and ring are known for a particular RRDE. In practice, the actual RRDE dimensions may not be known due to uncertainties in the machining process and changes in the dimensions induced by electrode polishing or temperature cycling. For this reason, it is common practice to empirically measure the collection efficiency using a well-behaved redox system rather than to rely upon a computed value.

    The ferrocyanide/ferricyanide half reaction is a simple, single-electron, reversible half reaction that is often used as the basis for measuring collection efficiency.(36) The RRDE is placed in a solution containing a small concentration ( ~10 \; mM ) of potassium ferricyanide,  K_3Fe(CN)_6 , in a suitable aqueous electrolyte solution (such as  1.0 \; M potassium nitrate,  KNO_3 ) and is operated at rotation rates between  500 and  2000 \; RPM . Initially, both the ring and the disk electrodes are held at a sufficiently positive potential that no reaction occurs. Then, the potential of the disk electrode is slowly swept ( ~50 \; mV/sec ) towards more negative potentials, and a cathodic current is observed which corresponds to the reduction of ferricyanide to ferrocyanide at the disk.

    Fe(CN)_6^{3-} + e^- \rightarrow Fe(CN)_6^{4-}

    (reduction of ferricyanide to ferrocyanide at disk)

    As ferricyanide is reduced at the disk electrode, the ferrocyanide generated by this process is swept outward (radially) away from the disk electrode and toward the ring electrode. The ring electrode is held constant at a positive (oxidizing) potential throughout the experiment. Some (but not all) of the ferrocyanide generated at the disk travels close enough to the ring electrode that it is oxidized back to ferricyanide. Thus, an anodic current is observed at the ring electrode due to the oxidation of ferrocyanide to ferricyanide at the ring.

    Fe(CN)_6^{4-} \rightarrow Fe(CN)_6^{3-} + e^-

    (oxidation of ferrocyanide to ferricyanide at ring)

    The measured ratio of the ring (anodic) limiting current to the disk (cathodic) limiting current is the empirical collection efficiency. As the rotation rate increases, both the disk and the ring currents increase (see Figure 10.10). Because both the anodic and cathodic limiting currents are proportional to the square root of the rotation rate, the empirical collection efficiency is expected to be independent of the rotation rate.

    f10-10.jpg

    Figure 10.10: Rotating Ring-Disk Voltammograms at Various Rotation Rates

    Once the collection efficiency value has been established empirically for a particular RRDE, it can be treated as a property of that particular RRDE, even if the RRDE is used to study a different half reaction in a different solution on a different day. Although the empirically measured collection efficiency ( N_{empirical} ) is a ratio of two currents with opposite mathematical signs (anodic and cathodic), the collection efficiency is always expressed as a positive number.

    N_{empirical} = -\frac{i_{Limiting, \; ring}}{i_{Limiting, \; disk}}

    10.5.3 – Generator/Collector Experiments

    When a molecule or ion is oxidized or reduced at an electrode, it is often transformed into an unstable intermediate chemical species which, in turn, is likely to undergo additional chemical changes. The intermediate may have a long enough lifetime that it is capable of moving to the ring electrode and being detected. Or, the intermediate may be so unstable that it decays away before it can be detected at the ring. Consider the following reaction scheme at a rotating ring-disk electrode:

     A+ne^-{\rightarrow}X

     X \xrightarrow{k} Z

     X \rightarrow A + ne^-

    (reduction of A to unstable intermediate X at disk electrode)

    (chemical decay of X to electrochemically inactive Z)

    (oxidation of X back to A at ring electrode)

    In the above scheme, the disk electrode is poised at a potential where  A is reduced to  X , and the cathodic limiting current observed at the disk ( i_{DISK} ) is a measure of how much  X is being “generated” at the disk electrode. At the same time, the ring electrode is poised at a more positive potential where  X is oxidized back to  A , and the anodic limiting current observed at the ring ( i_{RING} ) is a measure of much  X is being “collected” at the ring. There is also a competing chemical reaction which is capable of eliminating  X before it has a chance to travel from the disk to the ring.

    The ratio of the ring current to the disk current under these conditions is called the apparent collection efficiency ( N_{apparent} ).

    N_{apparent} = -\frac{i_{RING}}{i_{DISK}}

    By comparing the apparent collection efficiency ( N_{apparent} ) to the previously measured empirical collection efficiency ( N_{empirical} ) for the same RRDE, it is possible to deduce the rate at which the competing chemical pathway is converting  X to  Z . That is, it is possible to use an RRDE “generator/collector” experiment to measure the kinetic behavior of unstable electrochemical intermediates.

    Whenever  N_{apparent} is equal to  N_{empirical} , it is an indication that the decay rate of the intermediate (via the  X \rightarrow Z pathway) is small with respect to the transit time required for  X to travel from the disk to the ring. One way to shorten the transit time is to spin the RRDE at a faster rate. At high rotation rates, the apparent collection efficiency should approach the empirical collection efficiency. Conversely, at slower rotation rates, the apparent collection efficiency may be smaller ( N_{apparent} <  N_{empirical} ) because some of the intermediate is consumed by the competing chemical pathway before  X can travel to the ring.

    By recording a series of rotating ring-disk voltammograms at different rotation rates and analyzing the results, it is possible to estimate the rate constant (k) associated with the intermediate chemical decay pathway. Various relationships have been proposed for this kind of analysis,(2) and one of the simplest is shown below.

    \frac{N_{empirical}}{N_{apparent}} = 1 + 1.28\left(\frac{\nu}{D}\right)^{1/3}\left(\frac{k}{\omega}\right)

    A plot of the ratio of the empirical to the apparent collection efficiency versus the reciprocal angular rotation rate should be linear. The slope of such a plot can yield the rate constant if the kinematic viscosity ( v ) and the diffusion coefficient ( D ) are known.

    10.5.4 – Comparing Two Competing Pathways

    Sometimes the intermediate generated by an electrochemical process can decay via two different pathways. As long as one of these pathways leads to an electrochemically active chemical species that can be detected at the ring, it is possible to determine which decay pathway is favored. Consider the following scheme:

     A + n_1e^- \rightarrow X

     X \xrightarrow{k_1} Z

     X \xrightarrow{k_2} Y

     Y \rightarrow B + n_2e^-

    (reduction of  A to unstable intermediate  X at disk electrode)

    (fast chemical decay of  X to electrochemically inactive  Z )

    (fast chemical decay of  X to electrochemically active  Y )

    (detection of  Y at ring electrode via oxidation of  Y to  B )

    In the above scheme, the disk electrode is poised at a potential where  A is reduced to  X , and the cathodic limiting current observed at the disk ( i_{DISK} ) is a measure of how much  X is being “generated” at the disk electrode. The intermediate  X is unstable, and as it is swept away from the disk and toward the ring, it rapidly decays to either  Y or  Z . By the time these species reach the ring, all of the  X has decayed away, and the solution in contact with the ring contains both  Y and  Z . The species  Z is electrochemically inactive and cannot be detected by the ring, but the species  Y is active. By carefully poising the ring electrode at a potential appropriate for detecting  Y (in this case, by oxidizing  Y to  B ), it is possible for the ring to “collect” any  Y which arrives at the surface of the ring.

    The ratio of the ring current (due to  Y being detected at the ring) to the disk current (due to  X being generated at the disk) reveals the extent to which the  X \rightarrow Y pathway is favored in comparison to the  X \rightarrow Z pathway. The fraction of the decay by the  X \rightarrow Y pathway ( {\theta}_{XY} ) can be computed as follows.

    {\theta}_{XY} = \left(\frac{1}{N_{empirical}}\right)\left(\frac{n_1}{n_2}\right)\left|\frac{i_{RING}}{i_{DISK}}\right|

    Note in the above equation that the fraction ( n_1/n_2 ) carefully accounts for any difference in the number of electrons involved in the disk half reaction and the number of electrons involved when detecting  Y at the ring electrode. Schemes involving more complex stoichiometry may require additional correction factors.

    The most commonly studied reaction at the RRDE is undoubtedly the oxygen reduction reaction (ORR).(33-43) When oxygen ( O_2 ) is dissolved in acidic media and reduced at a platinum electrode, one pathway leads to water as the ultimate reduction product while the other pathway leads to the formation of peroxide anions. In the context of hydrogen fuel cell research, the pathway which leads to water is preferred, and it is commonly called the four-electron pathway. The path to peroxide formation is called the two-electron pathway, and it is undesirable for a number of reasons, including the fact that peroxide can damage various polymer membrane materials found in a fuel cell. Further details on how to use an RRDE “generator/collector” experiment to distinguish between the two-electron and four-electron ORR pathways can be found in the electrochemical literature.(33,36)

    10.6 – Rotating Cylinder Electrode (RCE) Theory

    The rotating disk and ring-disk electrodes were developed primarily as a result of academic electroanalytical chemistry research. In contrast, the theory for the rotating cylinder electrode (RCE) was developed by industrial researchers(44-46) in the corrosion and electroplating communities. While the flow of solution at a rotating disk (or ring-disk) is laminar over a wide range of rotation rates, the flow at the surface of a rotating cylinder is turbulent(31) at all but the slowest rotation rates. Thus, the RCE is an excellent tool for creating and controlling turbulent flow conditions in the laboratory, and it is most commonly used to mimic turbulent corrosion conditions found in large scale industrial settings such as oilfield pipeline corrosion.(56-69)

    The turbulent flow at a rotating cylinder electrode conveys material from the bulk solution towards the electrode surface. While the bulk solution remains well stirred by the main vortex induced by the rotating electrode, the layer of solution adjacent to the cylinder surface tends to rotate with the electrode. Thus, a high shear condition is set up at the surface of the rotating cylinder, spinning off smaller Taylor vortices adjacent to the rotating electrode.

    Net movement of material to the surface of a rotating cylinder was first characterized by Morris Eisenberg(27,28) in 1954 (about the same time that Levich was describing the rotating disk electrode). Eisenberg’s work eventually led to the Eisenberg equation which gives the limiting current at a rotating cylinder electrode

    i_L = 0.0487nFAC(d_{cyl})^{+0.4}(D)^{+0.644}(\nu)^{-0.344}(\omega)^{+0.7}

    in terms of the concentration ( C ) and diffusion coefficient ( D ) of the molecule or ion being studied, the Faraday constant ( F = 96485 \; C/mol ), the electrode area ( A ), the diameter of the cylinder ( d_{cyl} ), the kinematic viscosity of the solution ( \nu ), and the angular rotation rate ( {\omega} = \frac{2{\pi}f}{60} , where  f is the rotation rate in revolutions per minute). In the years since Eisenberg’s initial work with the rotating cylinder, additional work by Gabe, Kear, Walsh, and Silverman has described industrial applications of the RCE.(29-32, 44-69)

     

    10.7 References

    The references for this section can be found here.

    Read More
  • PineChem Download

    afcbp1.jpg Related Hardware Links: CBP Bipotentiostat, Interface Boards

    Related Software Links: AfterMath, PineChem, NIDAQ Device Driver

    Existing PineChem users are encouraged to upgrade to our more powerful AfterMath software package.

      For an overview of our vintage software PineChem, review the User Guide

    Windows 7 / Windows Vista

    The use of PineChem with Windows 8 and Windows 7 and Windows Vista is not supported. Windows 8/7Vista users should upgrade to the more powerful AfterMath software instead.

    PineChem for Windows 2000/NT/XP Users

    PineChem 2.8.0H Download. Customers who are using Windows 2000/NT/XP may download a full version of PineChem 2.8.0H using the link below. This version of our PineChem software is the final version which will install under the Windows 2000/NT/XP operating systems.

     

    After downloading the file, you will need a password to “unzip” the encrypted WinZip archive file. Contact Pine Research with the serial number of your bipotentiostat and we will provide you with the password. All of the files in the WinZip file should be placed in a temporary folder on your hard drive. Then, click on the “setup.exe” application to install PineChem on your system.

    NIDAQ Version 7.4.1 Device Driver Download. The required NIDAQ device driver for use with Windows 2000/NT/XP is version 7.4.1 of the driver. This driver may be downloaded directly from the National Instruments web site. Note that it is necessary to use this “Legacy NIDAQ” device driver. Do not attempt to use a newer device driver.

    PineChem for Windows 98 Users

    PineChem 2.8.0H Download. Customers who are still using Windows 98 may download a full version of PineChem 2.8.0H using the link below. Pine no longer officially supports Windows 98, and this version of our PineChem software is the final version which will install under the Windows 98 operating system.

    After downloading the file, you will need a password to “unzip” the encrypted WinZip archive file. Contact Pine Research with the serial number of your bipotentiostat and we will provide you with the password. All of the files in the WinZip file should be placed in a temporary folder on your hard drive. Then, click on the “setup.exe” application to install PineChem on your system.

    NIDAQ Version 6.9.1 Device Driver Download. The required NIDAQ device driver for use with Windows 98 is an old version of the driver. This driver may be downloaded directly from the National Instruments web site. Note that it is necessary to use this “Legacy NIDAQ” device driver. Do not attempt to use a newer device driver with Windows 98.

    PineChem for Windows 95 Users

    PineChem 2.7.9A Download. Customers who are still using Windows 95 may download a full version of PineChem 2.7.9A using the link below. Pine no longer officially supports Windows 95, and this version of our PineChem software is the final version which will install under the Windows 95 operating system.

    After downloading the file, you will need a password to “unzip” the encrypted WinZip archive file. Contact Pine Research with the serial number of your bipotentiostat and we will provide you with the password. All of the files in the WinZip file should be placed in a temporary folder on your hard drive. Then, click on the “setup.exe” application to install PineChem on your system.

    NIDAQ Version 6.9.1 Device Driver Download. The required NIDAQ device driver for use with Windows 95 is an old version of the driver. This driver may be downloaded directly from the National Instruments web site. Note that it is necessary to use this “Legacy NIDAQ” device driver. Do not attempt to use a newer device driver with Windows 95.

    Device Driver Installation Instructions

    In general, installation of the driver is a simple matter of unzipping the device driver package to a temporary folder on your hard drive and then clicking on the “setup.exe” file.

    IMPORTANT SECURITY NOTE: The NIDAQ driver should always be installed from a Windows user account which has administrative privileges. If the user who will routinely be using the AFCBP1 Bipotentiostat does not have an account with administrative privileges, then special steps must be taken to assure that the user is granted access to certain installation folders. Click here for more details about this issue.

    Activating the Device Driver

    After you have installed the device driver on your computer (and after you have physically installed the interface board inside your computer), you should run the Measurement & Automation Explorer (MAX) utility. The MAX utility is installed on your desktop when you install the NIDAQ device driver.

    Using the MAX utility, look inside the “Legacy NIDAQ Devices” folder and make certain that your interface board appears. The interface board usually appears as a “PCI-MIO-16E-4” board in the list of legacy devices. It is usually assigned a device number equal to one (1).

    NIDAQ Device Driver Version History

    For an extremely detailed version of previous NIDAQ versions, consult the link below.

    Read More
  • Rotating Electrode References


    10.7 – References

    Basic Information about Electroanalytical Chemistry

    1.    PT Kissinger and WR Heineman, Laboratory techniques in electroanalytical chemistry, Marcel Dekker, New York (1996).

    2.    AJ Bard and LR Faulkner, Electrochemical Methods-Fundamentals and Applications, 2nd Edition, John Wiley & Sons, New York (2000) Chapter 9.

    Examples of Various Types of Hydrodynamic Voltammetry

    3.   DC Johnson, SG Weber, AM Bond, RM Wightman, RE Shoup and IS Krull, Electroanalytical voltammetry in flowing solutions, Analytica Chimica Acta 180 (1986) 187-250.

    4.   H Gunasingham and B Fleet, Wall-jet electrode in continuous monitoring voltammetry, Analytical Chemistry 55 (1983) 1409-1414.

    5.   JV Macpherson and PR Unwin, Hydrodynamic Modulation Voltammetry with an Oscillating Microjet Electrode, Analytical Chemistry 71 (1999) 4642.

    6.   IE Henley, K Yunus and AC Fisher, Voltammetry under Microfluidic Control: Computer-Aided Design Development and Application of Novel Microelectrochemical Reactors, J. of Physical Chemistry B 107 (2003) 3878-3884.

    7.   KW Pratt and DC Johnson, Vibrating wire electrodes—I. Literature review, design and evaluation, Electrochemica Acta 27 (1982) 1013-1021.

    8.   C Hagan and LA Coury, Comparison of hydrodynamic voltammetry implemented by sonication to a rotating disk electrode, Analytical Chemistry 66 (1994) 399-405.

    Seminal Publications Describing Rotating Electrodes

    9.   VG Levich, Physicochemical Hydrodynamics, Prentice-Hall, Upper Saddle River NJ (1962).

    10.   S. Bruckenstein and T. Nagai, The Rotated, Mercuy-Coated Platinum Electrode, Anal. Chem. 33 (1961) 1201.

    11.    Z Galus, C Olson, HY Lee and RN Adams Rotating Disk Electrodes, Anal. Chem. 34 (1962) 164.

    12.    WJ Albery and RP Bell, Kinetics of Dissociation of Weak Acids Measured by a Rotating Platinum Disc Electrode , Proc. Chem. Soc. (1963) 169.

    13.   S Bruckenstein and B Miller, Unraveling Reactions with Rotating Electrodes, Acc. Chem. Res. 10 (1977) 54-61.

    14.   S Treimer, A Tanga and DC Johnson, Consideration of the Application of Koutecky-Levich Plots in the Diagnoses of Charge-Transfer Mechanisms at Rotated Disk Electrodes, Electroanalysis 14 (2002) 165-171.

    15.    EF Dalton, Historical Origins of the Rotating Ring-Disk Electrode, Electrochem. Soc. Interface 23 (2016) Issue 3, 50-59.

    16.    AN Frumkin, LN Nekrasov, VG Levich, and YB Ivanov, The Use of a Rotating Ring-Disk Electrode for Studying Intermediate Products of Electrochemical Reactions, J. Electroanal. Chem. 1 (1959) 84.

    17.    LN Nekrasov and NP Berezina, The Electrolytic Reduction of Copper on a Disc-Ring Electrode, Dokl. Akad. Nauk SSSR 142 (1961) 885.

    18.    LN Nekrasov and L Müller, Study of the Cathodic Reduction of Oxygen on Platinum in Alkaline Solutions Using a Rotating Disk and Ring Electrode, Dokl. Akad. Nauk SSSR 149 (1963) 1107.

    19.    L Müller and LN Nekrasov, Study of the Electroreduction of Oxygen on Smooth Platinum in Acid Solutions by the Method of a Revolving Disc Electrode with a Ring, Dokl. Akad. Nauk SSSR 154 (1964) 437.

    20.    S. Bruckenstein, The Relations Between the Limiting Diffusion Currents at Rotating Disk, Ring, and Ring-Disk Electrodes, Elektrokhimiya 2 (1966) 1085.

    21.    WJ Albery and ML Hitchman, Ring-Disc Electrodes, Clarendon Press, Oxford (1971).

    22.   WJ Albery, Ring-disc electrodes. Part 1.— A new approach to the theory, Trans. Faraday Soc. 62 (1966) 1915-1919.

    23.   WJ Albery and S Bruckenstein, Ring-disc electrodes. Part 2.— Theoretical and experimental collection efficiencies, Trans. Faraday Soc. 62 (1966) 1920-1931.

    24.   WJ Albery, S Bruckenstein and DT Napp, Ring-disc electrodes. Part 3.— Current-voltage curves at the ring electrode with simultaneous currents at the disc electrode, Trans. Faraday Soc. 62 (1966) 1932-1937.

    25.   WJ Albery, S Bruckenstein and DC Johnson, Ring-disc electrodes. Part 4.— Diffusion layer titration curves, Trans. Faraday Soc. 62 (1966) 1938-1945.

    26.   WJ Albery, Ring-disc electrodes. Part 5.— First-order kinetic collection efficiencies at the ring electrode, Trans. Faraday Soc. 62 (1966) 1946-1954.

    27.   M Eisenberg, CW Tobias and CR Wilke, Ionic Mass Transfer and Concentration Polarization at Rotating Electrodes, Journal of the Electrochemical Society 101 (1954) 306.

    28.   M Eisenberg, CW Tobias and CR Wilke, Chem. Eng. Progr. Symp. Ser. 51 (1955) 1.

    29.   DR Gabe, Rotating Cylinder Electrode, J. Appl. Electrochem. 4 (1974) 91.

    30.   DR Gabe and DJ Robinson, Mass Transfer in a Rotating Cylinder Cell–I. Laminar Flow, Electrochemica Acta 17 (1972) 1121.

    31.   DR Gabe and DJ Robinson, Mass Transfer in a Rotating Cylinder Cell–II. Turbulent Flow, Electrochemica Acta 17 (1972) 1129.

    32.   DR Gabe and FC Walsh, The Rotating Cylinder Electrode: A Review of Development, J. Appl. Electrochem. 13 (1983) 3.

    Studies of Oxygen Reduction Reaction (ORR) Kinetics using Rotating Disk and Ring-Disk Electrodes (RDE and RRDE)

    33.   Y Garsany, OA Baturina, KE Swider-Lyons and SS Kocha, Experimental Methods for Quantifying the Activity of Platinum Electrocatalysts for the Oxygen Reduction Reaction, Analytical Chemistry 82 (2010) 6321-6328.

    34.   HA Gasteiger, SS Kocha, B Sompalli and FT Wagner, Activity benchmarks and requirements for Pt, Pt-alloy, and non-Pt oxygen reduction catalysts for PEMFCs, Applied Catalysis B: Environmental 56 (2005) 9-35.

    35.   UA Paulus, A Wokauna, GG Scherera, TJ Schmidt, V Stamenkovic, NM Markovic and PN Ross, Oxygen reduction on high surface area Pt-based alloy catalysts in comparison to well defined smooth bulk alloy electrodes, Electrochimica Acta 47 (2002) 3787-3798.

    36.   UA Paulus, TJ Schmidt, HA Gasteiger and RJ Behm, Oxygen reduction on a high-surface area Pt/Vulcan carbon catalyst: a thin-film rotating ring-disk electrode study, J. of Electroanalytical Chem. 495 (2001) 134-145.

    37.   TJ Schmidt, UA Paulus, HA Gasteiger and RJ Behm, The oxygen reduction reaction on a Pt/carbon fuel cell catalyst in the presence of chloride anions, J. of Electroanalytical Chem. 508 (2001) 41-47.

    38.   G Brisard, N Bertranda, PN Ross and NM Markovic, Oxygen reduction and hydrogen evolution–oxidation reactions on Cu(hkl) surfaces, J. of Electroanalytical Chem. 480 (2000) 219-224.

    39.   L Geniès, R Faure and R Durand, Electrochemical reduction of oxygen on platinum nanoparticles in alkaline media, Electrochimica Acta 44 (1998) 1317-1327.

    40.   E Higuchia, H Uchidab and M Watanabe, Effect of loading level in platinum-dispersed carbon black electrocatalysts on oxygen reduction activity evaluated by rotating disk electrode, J. of Electroanalytical Chem. 583 (2005) 69-76.

    41.   ZD Weia, SH Chanb, LL Lia, HF Caia, ZT Xiab and CX Sunc, Electrodepositing Pt on a Nafion-bonded carbon electrode as a catalyzed electrode for oxygen reduction reaction, Electrochimica Acta 50 (2005) 2279-2287.

    42.   S Marcotte, D Villers, N Guillet, L Roué and JP Dodelet, Electroreduction of oxygen on Co-based catalysts: determination of the parameters affecting the two-electron transfer reaction in an acid medium, Electrochimica Acta 50 (2004) 179-188.

    43.   S Durón, R Rivera-Noriega, P Nkeng, G Poillerat and O Solorza-Feria, Kinetic study of oxygen reduction on nanoparticles of ruthenium synthesized by pyrolysis of Ru3(CO)12, J. of Electroanalytical Chem. 566 (2004) 281-289.

    Applications of the Rotating Cylinder Electrode (RCE)

    44.   DR Gabe and FC Walsh, Enhanced Mass Transfer at the Rotating Cylinder Electrode–I. Characterization of a Smooth Cylinder and Roughness Development in Solutions of Constant Concentration, J. Appl. Electrochem. 14 (1984) 555.

    45.   DR Gabe and FC Walsh, Enhanced Mass Transfer at the Rotating Cylinder Electrode–II. Development of Roughness for Solutions of Decreasing Concentration, J. Appl. Electrochem. 14 (1984) 565.

    46.   DR Gabe and FC Walsh, Enhanced Mass Transfer at the Rotating Cylinder Electrode–III. Pilot and Production Plant Experience, J. Appl. Electrochem. 15 (1985) 807.

    47.   DR Gabe and PA Makanjuola, Enhanced Mass Transfer Using Roughened Rotating Cylinder Electrodes in Turbulent Flow, J. Appl. Electrochem. 17 (1987) 370.

    48.   DR Gabe, GD Wilcox, J Gonzalez-Garcia and FC Walsh, The Rotating Cylinder Electrode: Its Continued Development and Application, J. Appl. Electrochem. 28 (1998) 759.

    49.   G Kear, BD Barker, K Stokes and FC Walsh, Flow Influenced Electrochemical Corrosion of Nickel Aluminum Bronze – Part I. Cathodic Polarization, J. Appl. Electrochem. 34 (2004) 1235.

    50.   G Kear, BD Barker, K Stokes and FC Walsh, Flow Influenced Electrochemical Corrosion of Nickel Aluminum Bronze – Part II. Anodic Polarization and Derivation of the Mixed Potential, J. Appl. Electrochem. 34 (2004) 1241.

    51.   Q Lu, MM Stack and CR Wiseman, AC Impedance Spectroscopy as a Technique for Investigating Corrosion of Iron in Hot Flowing Bayer Liquors, J. Appl. Electrochem. 31 (2001) 1373.

    52.   JM Maciel and SML Agostinho, Use of a Rotating Cylinder Electrode in Corrosion Studies of a 90/10 Cu–Ni Alloy in 0.5M H2SO4 Media, J. Appl. Electrochem. 30 (2000) 981.

    53.   JM Grau and JM Bisang, Mass Transfer Studies at Rotating Cylinder Electrodes of Expanded Metal, J. Appl. Electrochem. 35 (2005) 285.

    54.   A Eklund and D Simonsson, Enhanced Mass Transfer to a Rotating Cylinder Electrode with Axial Flow, J. Appl. Electrochem. 18 (1988) 710.

    55.   KD Efird, EJ Wright, JA Boros and TG Hailey, Correlation of Steel Corrosion in Pipe Flow with Jet Impingement and Rotating Cylinder Tests, Corrosion 49 (1993) 992.

    56.   DC Silverman, Rotating Cylinder Electrode for Velocity Sensitivity Testing, Corrosion 40 (1984) 220.

    57.   DC Silverman and ME Zerr, Application of the Rotating Cylinder Electrode – E-Brite® 26-1 in Concentrated Sulfuric Acid, Corrosion 42 (1986) 633.

    58.   DC Silverman, Rotating Cylinder Electrode – Geometry Relationships for Prediction of Velocity-Sensitive Corrosion, Corrosion 44 (1988) 42.

    59.   DC Silverman, Corrosion Prediction in Complex Environments using Electrochemical Impedance Spectroscopy, Electrochimica Acta 38 (1993) 2075.

    60.   DC Silverman, Technical Note: On Estimating Conditions for Simulating Velocity-Sensitive Corrosion in the Rotating Cylinder Electrode , Corrosion 55 (1999) 1115.

    61.   DC Silverman, Technical Note: Simplified Equation for Simulating Velocity-Sensitive Corrosion in the Rotating Cylinder Electrode at Higher Reynolds Numbers, Corrosion 59 (2003) 207.

    62.   DC Silverman, The Rotating Cylinder Electrode for Examining Velocity-Sensitive Corrosion – A Review, Corrosion 60 (2004) 1003.

    63.   DC Silverman, Technical Note: Conditions for Similarity of Mass-Transfer Coefficients and Fluid Shear Stresses between the Rotating Cylinder Electrode and Pipe, Corrosion 61 (2005) 515.

    64.   G Wranglen, J Berendson and G Karlberg, Apparatus for Electrochemical Studies of Corrosion Processes in Flowing Systems, in Physico-Chemical Hydrodynamics, edited by B Spalding (London: Adv. Publications, 1977) 461.

    65.   RA Holser, G Prentice, RB Pond and R Guanti, Use of Rotating Cylinder Electrodes to Simulate Turbulent Flow Conditions in Corrosion Systems, Corrosion 46 (1990) 764.

    66.   TY Chen, AA Moccari and DD Macdonald, Development of Controlled Hydrodynamic Techniques for Corrosion Testing, Corrosion 48 (1992) 239.

    67.   S Nesic, GT Solvi and S Skjerve, Comparison of Rotating Cylinder and Loop Methods for Testing CO2 Corrosion Inhibitors, British Corrosion Journal 32 (1997) 269.

    Standard Methods for using the Rotating Cylinder Electrode (RCE) in Oilfield Corrosion Studies

    68.   ASTM G 170, Standard Guide for Evaluating and Qualifying Oilfield and Refinery Corrosion Inhibitors in the Laboratory (2001).

    69.   ASTM G 185, Standard Practice for Evaluating and Qualifying Oil Field and Refinery Corrosion Inhibitors Using the Rotating Cylinder Electrode (2006).

     

    Read More
  • Modulated Speed Rotator (MSR) User Manual

    Overview

    Welcome to the User Manual for the Modulated Speed Rotator (MSR) system offered by Pine Research Instrumentation. The MSR system is a general purpose electrode rotator which may be used with Rotating Disk Electrodes (RDE), Rotating Ring-Disk Electrodes (RRDE), and Rotating Cylinder Electrodes (RCE).

    The most recent complete version of this manual can be downloaded as a PDF file (click here to download – 5.0 MB PDF file). Portions of the manual have also been transcribed to wiki format and can be reached by clicking on the active links in the table of contents below.

    Related Links: MSR Main Support Page, Pine Research Instrumentation
    (keywords: rotating desk electrode, rotating ring-disk electrode, rotating cylinder electrode)

    Table of Contents

    Read More
  • MSR Rotator: Motor Cable

    msr_small.jpg

    The MSR rotator has a motor controller which is connected to the motor unit via a motor cable.

    Related Links: MSR Rotator

    HD-15 Connector

    The motor cable for the AFMSRCE rotator utilizes a 15-pin cable with HD-15 connectors on each end. The pinout for this connector is explained in the table below:

    wavenow_db15_pinout.jpg
    pins description
    1-4 Motor Supply (-)
    6-9 Motor Supply (+)
    11-12 Tachometer (+)
    13-14 Tachometer (-)

    Motor Control Cable Wiring

    Earlier Cable Designs

    Earlier versions of the MSR rotator (such as the AFMSRX and AFMSRXE) connected to a round black connector on top of the motor unit. There were a total of four pins in this connector. One pair of pins provided power to the motor, and the other pair of pins provided the rotation rate signal from the tachometer.

    Round Connector Wiring Scheme
    msrx_cowling_top_view.jpg
    pin description wire color
    1 Motor Supply (+) white
    3 Motor Supply (-) green
    9 Tachometer Signal (+) red
    7 Tachometer Signal (-) black

    Internal Wiring

    If you remove the cowling from the motor unit and inspect the cables connected to the motor and tachometer, you will see four cables with different colors. The signals in each cable are described in the table below.

    wire color description
    green Motor Supply (-)
    white Motor Supply (+)
    red Tachometer (+)
    black Tachometer (-)

    Internal Motor Cable Wiring

    Read More
  • MSR Rotator: Operation

    Table of Contents

    Section 4 – Operation


    This section of the manual discusses information pertaining to routine operation of the rotator. Users of the rotator should be familiar with all of the information in this section prior to operating the rotator.

    4.1 – The Rotating Shaft

    The electrode shaft normally rotates in a clockwise direction as viewed from the top of the rotator. The upper end of a standard shaft has a  1/4" ( 6.35 \; mm ) outer diameter. When properly mounted in the rotator, the upper  2.7" ( 68 \; mm ) of the shaft is inside the motor unit, while the remaining length of the shaft extends down below the motor unit.

    The rotator accepts shafts for use with Rotating Disk Electrodes (RDE), Rotating Cylinder Electrodes (RCE) or Rotating Ring-Disk Electrodes (RRDE). Electrical connection is accomplished using one or more silver-carbon brushes to contact metal surfaces on the upper portion of the rotating shaft. Each shaft is specially designed to provide one or two current paths down to the electrode tip. These current paths are electrically isolated from the mounting area near the top of the shaft.

    f4-1.jpg

    Figure 4.1: Contact Areas at Top of Rotating Electrode Shafts

    The uppermost portion of the shaft is used to mount the shaft into the rotator (see Figure 4.1). This mounting area is electrically isolated from the remainder of the shaft so that the electrode connections remain isolated from the rotator chassis. An insulating spacer just below the mounting area isolates the mounting area from the electrode contact area.

    For an RDE or RCE shaft (see Figure 4.1, left), the entire metal exterior of the shaft below the insulating spacer is in electrical contact with the disk (or cylinder) electrode. For an RRDE shaft (see Figure 4.1, right), there are two insulating spacers. The portion of the shaft between the two insulating spacers provides electrical contact with the disk electrode. The lower portion of the shaft (below the lower insulating spacer) provides electrical contact with the ring electrode.

    f4-2.jpg

    Figure 4.2: The Brush Chamber (side view)

    The shaft is connected to the rotator motor via a brass motor coupling located inside the brush chamber (see Figure 4.2). Two clamshell doors surround the brush chamber. These doors are securely latched during rotator operation and push two pairs of contact brushes against the rotating shaft. The upper (red) pair of brushes makes contact with the disk (or cylinder) while the lower (blue) pair makes contact with the ring on a rotating ring-disk electrode.

    4.1.1 – Installing a Shaft

    icon_no_power.jpg WARNING:

    Rotating shaft. Entanglement hazard.
    Turn off the power to the rotator and disconnect the power cord from the power source before installing or removing the electrode shaft or before installing or removing an electrode tip on the end of the shaft.

    AVERTISSEMENT:

    Arbre en rotation. Danger d’enchevêtrement.
    Éteignez le rotateur et débranchez le cordon d’alimentation de la source d’alimentation avant d’installer ou d’enlever l’arbre de l’électrode ou avant d’installer ou d’enlever un embout d’électrode à l’extrémité de l’arbre.

    icon_caution.jpg

    WARNING:

    Do not use or attempt to rotate an electrode shaft that has been dropped, bent or otherwise physically damaged.
    Inspect the shaft to be certain that it is not damaged.

    AVERTISSEMENT:

    N’utilisez pas et ne tentez pas de mettre en rotation un arbre d’électrode qui est tombé, a été tordu ou a été endommagé physiquement d’une autre manière ou d’une autre.
    Inspectez l’arbre pour vous assurer qu’il n’a pas été endommagé.

    icon_bulb.jpg TIP:

    It is often easier to remove or install a shaft by disconnecting the motor control cable and inverting the entire motor unit on the center post. Several of the photos in this section of the manual show the rotator motor in such an inverted position.

    icon_bulb.jpg TIP:

    Do not lose the white plastic washer on the door latch.

    1. Invert the orientation of the motor unit so that it is upside-down as shown. Loosen the latch on the clamshell doors. Open the doors to provide access to the brush chamber.
    2. s411a.jpg
    3. If there is a shaft already installed, use the hex driver tool ( 5/64" , provided) to loosen the two screws on the motor coupling. Do not remove these screws entirely; just loosen them by one or two turns of the hex driver. Usually it is necessary to hold the motor coupling in place with one hand while loosening the screws with the other hand.
    4. s411b.jpg

      icon_info.jpgNOTE:

      A new rotator has tape around the motor coupling to protect the hex screws. Remove this tape and loosen the hex screws if needed to allow the shaft to enter the coupling.

      icon_bulb.jpg TIP:

      Apply a small amount of a silicon-based grease to the top of the shaft before installing the shaft into the motor coupling. This helps to prevent the shaft from sticking in the coupling.

      f4-3.jpg

      Figure 4.3: Proper (left) and Improper (right) Shaft Insertion Positions

    5. Install the shaft by sliding it through the hole in the lower bearing assembly and into the brush chamber.
      • The shaft should be pushed as far as possible into the motor coupling so that the contact brushes are properly aligned with the electrical contact areas on the rotating electrode shaft (see Figure 4.3).
      • If the shaft is properly installed, the brushes will contact metal surfaces on the shaft.
      • If the shaft is improperly installed, the brushes may contact an insulating gap on the shaft, and the connection to the rotating electrode will fail.
      s411c.jpg
    6. Use the hex driver tool ( 5/64" ) to securely tighten both hex screws on the motor coupling. Gently tug on the shaft to make sure it is securely mounted in the motor coupling.
    7. s411d.jpg
    8. Close the clamshell doors and tighten the latch.
    9. Remount the motor unit on the center post (in the non-inverted position).
    10. icon_caution.jpg

      CAUTION:

      Before reconnecting the rotator power cable or the motor control cable to the control unit, be sure the control unit power switch is off and the rotation rate knob is turned to the fully counterclockwise position.

      ATTENTION:

      Avant de reconnecter le câble d’alimentation du rotateur ou le câble de commande du moteur à l’unité de commande, assurez-vous que l’interrupteur de l’unité de commande est en position éteinte et que le bouton de commande de la vitesse de rotation est complètement tourné dans la position inverse des aiguilles d’une montre.

    11. Reconnect the motor control cable from the control unit to the motor unit.
    12. Reconnect the power cable from the power source to the control unit.
    13. icon_caution.jpg WARNING:

      Do not turn on the rotator or rotate the electrode shaft if the shaft is not securely mounted in the motor coupling.
      Inspect the shaft to be certain that it is securely mounted.

      AVERTISSEMENT:

      Ne mettez pas le rotateur en marche ni l’arbre de l’électrode en rotation si l’arbre n’est pas correctement raccordé au moteur.
      Inspectez l’arbre pour vous assurer qu’il est bien fixé.

    14. With the rotation rate knob in the fully counterclockwise position, turn on the control unit.
    15. Slowly turn the rotation rate knob clockwise until the shaft rotates between  100 and  200 \; RPM .
    16. While the shaft is slowly rotating ( 100 to  200 \; RPM ), inspect the rotating shaft to assure that it is rotating properly about the axis of rotation. If the shaft is wobbling, vibrating, or tilting away from the axis of rotation, then turn off the rotator and remove the shaft from the rotator.
    17. icon_info.jpgNOTE:

      A “Precision Shaft Alignment Kit” is available separately. This kit includes a dial indicator used to measure the “runout” at the end of the shaft (see Section 7.1 for kit part number).

      icon_caution.jpg WARNING:

      Do not use an electrode shaft which appears to wobble, vibrate, or tilt away from the axis of rotation while rotating. Such a shaft is either improperly installed or physically damaged. Turn off the rotator, disconnect electrical power, and remove the shaft immediately.

      AVERTISSEMENT:

      N’utilisez pas un arbre d’électrode qui semble osciller, vibrer ou dévier de l’axe de rotation pendant la rotation. Cet arbre est soit installé de manière incorrecte soit endommagé physiquement. Éteignez le rotateur,déconnectez l’alimentation électrique et retirez l’arbre immédiatement.

    18. If the shaft is rotating properly along the axis of rotation, then it is ready for use. Some shafts are actually single-piece electrodes where the electrode tip is permanently attached to the shaft. But most shafts are designed to accept a variety of different tips. For these “shaft and tip” designs, the shaft may remain mounted in the rotator, and changing the tip is a simple matter of unscrewing one tip and then threading a new tip on to the shaft.
    f4-4.jpg

    Figure 4.4: Installing a Tip on to a Shaft

    4.1.2 – Changing the Tip on a Shaft

    icon_no_power.jpg WARNING:

    Rotating shaft. Entanglement hazard.
    Turn off the power to the rotator and disconnect the power cord from the power source before installing or removing the electrode shaft or before installing or removing an electrode tip on the end of the shaft.

    AVERTISSEMENT:

    Arbre en rotation. Danger d’enchevêtrement.
    Éteignez le rotateur et débranchez le cordon d’alimentation de la source d’alimentation avant d’installer ou d’enlever l’arbre de l’électrode ou avant d’installer ou d’enlever un embout d’électrode à l’extrémité de l’arbre.

    icon_caution.jpg

    WARNING:

    Do not use or attempt to rotate an electrode shaft that has been dropped, bent or otherwise physically damaged.
    Inspect the shaft to be certain that it is not damaged.

    AVERTISSEMENT:

    N’utilisez pas et ne tentez pas de mettre en rotation un arbre d’électrode qui est tombé, a été tordu ou a été endommagé physiquement d’une autre manière ou d’une autre.
    Inspectez l’arbre pour vous assurer qu’il n’a pas été endommagé.

    icon_caution.jpg

    WARNING:

    Do not use an electrode shaft which appears to wobble, vibrate, or tilt away from the axis of rotation while rotating. Such a shaft is either improperly installed or physically damaged. Turn off the rotator, disconnect electrical power, and remove the shaft immediately.

    AVERTISSEMENT:

    N’utilisez pas un arbre d’électrode qui semble osciller, vibrer ou dévier de l’axe de rotation pendant la rotation. Cet arbre est soit installé de manière incorrecte soit endommagé physiquement. Éteignez le rotateur, déconnectez l’alimentation électrique et retirez l’arbre immédiatement.

    icon_caution.jpg

    WARNING:

    Do not use or attempt to rotate an electrode tip that has been dropped or otherwise physically damaged.
    Inspect the electrode tip to be certain that it is not damaged.

    AVERTISSEMENT:

    N’utilisez pas et ne tentez pas de mettre en rotation un embout d’électrode qui est tombée ou a été endommagée physiquement d’une autre manière ou d’une autre.
    Inspectez l’embout d’électrode pour vous assurer qu’elle n’a pas été endommagée.

    1. When removing a tip from a shaft or installing a new tip on a shaft, use one hand to prevent the shaft from rotating while using the other hand to gently turn the tip.
    2. Remove the old tip from the shaft by gently unscrewing the tip by hand.
      No tools are required to remove a tip from a shaft.
    3. icon_caution.jpg CAUTION:

      Do not use tools on the shaft or electrode tip.
      Never use a tool to unscrew a tip from a shaft.
      If a tip cannot be removed from a shaft by hand, then contact the factory for further instructions.

      ATTENTION:

      N’utilisez pas d’outils sur l’arbre ou sur l’embout d’électrode.
      N’utilisez jamais d’outil pour dévisser un embout d’électrode d’un arbre.
      Si un embout d’électrode ne peut être retirée d’un arbre manuellement, communiquez avec l’usine pour obtenir des instructions supplémentaires.


    4. Thread the new tip on to the shaft (see Figure 4.4) and gently tighten it by hand. Never use a tool to tighten the tip on to the shaft.
    5. icon_caution.jpg CAUTION:

      Before reconnecting the rotator power cable or the motor control cable to the control unit, be sure the control unit power switch is off and the rotation rate knob is turned to the fully counterclockwise position.

      ATTENTION:

      Avant de reconnecter le câble d’alimentation du rotateur ou le câble de commande du moteur à l’unité de commande, assurez-vous que l’interrupteur de l’unité de commande est en position éteinte et que le bouton de commande de la vitesse de rotation est complètement tourné dans la position inverse des aiguilles d’une montre.


    6. Reconnect the motor control cable from the control unit to the motor unit.
    7. Reconnect the power cable from the power source to the control unit.
    8. icon_caution.jpg WARNING:

      Do not turn on the rotator or rotate the electrode shaft if the shaft is not securely mounted in the motor coupling.
      Inspect the shaft to be certain that it is securely mounted.

      AVERTISSEMENT:

      Ne mettez pas le rotateur en marche ni l’arbre de l’électrode en rotation si l’arbre n’est pas correctement raccordé au moteur.
      Inspectez l’arbre pour vous assurer qu’il est bien fixé.

    9. With the rotation rate knob in the fully counterclockwise position, turn on the control unit.
    10. Slowly turn the rotation rate knob clockwise until the shaft is rotating between  100 and  200 \; RPM .
    11. While the shaft is slowly rotating ( 100 to  200 \; RPM ), inspect the rotating shaft and tip to assure that both are rotating properly about the axis of rotation. If the shaft or tip is wobbling, vibrating, or tilting away from the axis of rotation, then turn off the rotator and remove the shaft from the rotator.
    12. icon_caution.jpg WARNING:

      Do not use an electrode tip which appears to wobble, vibrate, or tilt away from the axis of rotation while rotating. Such an electrode tip is either improperly installed or physically damaged. Turn off the rotator, disconnect electrical power, and remove the electrode tip immediately.

      AVERTISSEMENT:

      N’utilisez pas un embout d’électrode qui semble osciller, vibrer ou dévier de l’axe de rotation pendant la rotation. Ce embout d’électrode est soit installée de manière incorrecte soit endommagée physiquement. Éteignez le rotateur, déconnectez l’alimentation électrique et retirez l’embout d’électrode immédiatement.


      icon_info.jpgNOTE:

      A “Precision Shaft Alignment Kit” is available separately. This kit includes a dial indicator used to measure the “runout” at the end of the shaft (see Section 7.1 for kit part number).

    13. If the shaft and tip are rotating properly along the axis of rotation, then the next step is to mount the electrochemical cell that holds the test solution (see Section 4.2).

    4.2 – Mounting the Cell

    All cells should be clamped to the side post and also supported from below using the cell platform. For a cell with multiple side ports, carefully orient the cell so that any accessories mounted in the side ports have enough clearance. Smaller cells may be clamped using a traditional laboratory clamp secured to the center port (see Figure 4.5, left). Larger cells may be clamped using a large diameter column clamp (see Figure 4.5, right).

    f4-5a.jpg f4-5b.jpg

    Figure 4.5: Properly Supported and Clamped Electrochemical Cells

    The cell platform and clamp positions allow adjustment of the vertical position of the cell with respect to the motor unit. In addition, the vertical position of the motor unit is easily adjusted. Usually, it is easier to mount and clamp the cell in a fixed vertical position. Then, the rotating electrode can be moved vertically down into the cell or up out of the cell as needed.


    icon_caution.jpg CAUTION:

    When raising and lowering the motor unit along the main support rod, be sure to hold the motor unit carefully so that it does not unexpectedly fall and break the glass cell located below the motor unit.

    ATTENTION:

    Lorsque vous montez ou descendez le bloc moteur le long de la barre principale, veillez à bien le tenir pour éviter qu’il ne chute brutalement et ne casse la cellule de verre située sous le bloc moteur.


    icon_caution.jpg CAUTION:

    Position the motor unit with respect to the glass cell so that the electrode tip is immersed approximately   \textbf{~1.0 cm} into the test solution.
    Excessive immersion may corrode the shaft or tip by allowing liquids to seep into the joint between the shaft and tip.

    ATTENTION:

    Positionnez le bloc moteur en fonction de la position de la cellule de verre, de telle sorte que l’embout d’électrode soit immergée sur environ 1 cm dans la solution d’essai.
    Une immersion excessive peut entraîner la corrosion de l’arbre ou d’embout d’électrode en provoquant l’infiltration de liquides dans le joint situé entre l’arbre et l’embout d’électrode.


    icon_caution.jpg CAUTION:

    Center the rotating electrode within the opening on the cell so that it does not rub against the walls of the opening. Damage will occur if the rotating shaft or tip abrades against these walls.

    ATTENTION:

    Centrez l’électrode rotative dans l’ouverture de la cellule pour qu’elle ne frotte pas les bords de l’ouverture.
    Le frottement des bords de la cellule par l’arbre ou par l’embout d’électrode entraînera des dommages.

    4.3 – The Enclosure


    icon_caution.jpg WARNING:

    Rotating shaft.
    Do not turn on the rotator or rotate the electrode shaft unless the enclosure window is secured to all four pins as shown below.
    Use extreme caution when operating the rotator at rotation rates above  \textbf{2000 RPM}

    AVERTISSEMENT:

    Arbre en rotation.
    Ne mettez pas le rotateur en marche et ne marche ni l’arbre de l’électrode en rotation si la fenêtre du boîtier n’est pas fermée à l’aide des quatre broches tel qu’indiqué ci-dessous.
    Soyez extrêmement prudent lorsque vous utilisez le rotateur à des vitesses de rotation supérieures à  \textbf{2000 tr/min} .

    Figure-4.6-The-Enclosure.jpg

    Figure 4.6: Enclosure Properly Mounted on All Four Pins

    After the cell has been mounted and the electrode has been lowered into the cell, securely mount the enclosure by hooking the enclosure to the four pins on the enclosure base (see Figure 4.6).

    Note that the enclosure has small openings near the bottom which permit cell connections, purge gas tubing, and coolant to be carefully routed to the electrochemical cell from locations outside the enclosure.

    4.4 – Cell Connections

    The counter electrode and the reference electrode are usually mounted in appropriate side ports on the electrochemical cell (see Figure 4.7). The counter electrode is often a simple platinum wire or carbon rod to which an alligator clip is easily affixed.

    Figure-4.7-PINE-NEW-Counter-and-Reference.jpg

    Figure 4.7: Connection of Counter and Reference Electrodes

    Always consult the manual for the potentiostat system to determine which cell cable leads should be connected to the counter and reference electrodes. For newer Pine potentiostats, the reference electrode cable is color coded as white, and the counter electrode cable is color coded as green. Many commercially available reference electrodes have a sturdy pin connector on the top end which can accept an alligator clip. The cable which connects the reference electrode to the potentiostat should be of the shielded (coaxial) type, and care should be taken to route this cable well away from noise sources such as power cords, networking cables, or video monitors.


    icon_info.jpg NOTE:

    Cell cables on newer model Pine potentiostats use GREEN to mark the counter electrode connection and WHITE to mark the reference electrode connection (see Figure 4.7).
    Older Pine bipotentiostats use RED to mark the counter electrode and use a BNC connector for the reference electrode.

    icon_bulb.jpg TIP:

    There is no universally accepted color coding scheme for marking potentiostat cell cable connections. If you are using the rotator with a third-party potentiostat, consult the potentiostat documentation for information about the cell cable markings.

    4.4.1- RDE and RCE Wiring

    There are two pairs of brushes which provide electrical contact with the rotating shaft (see Figure 4.8). The upper pair of brush contacts (red) is used to make electrical contact with a rotating disk electrode (RDE) or a rotating cylinder electrode (RCE).

    To make good contact on opposite sides of the rotating shaft, both of the red brushes (left and right sides) should be used. Use a short banana jumper cable to connect the opposing brushes together (see Figure 4.8), and then connect the working electrode cable(s) from the potentiostat to the jumper cable.

    f4-8.jpg

    Figure 4.8: Brush Connections for a Rotating Disk Electrode (RDE) or a Rotating Cylinder Electrode (RCE)

    icon_bulb.jpg TIP:

    Most modern potentiostats provide separate cable connections for the working electrode “drive” line and for the working electrode “sense” line. The drive line carries current while the sense line measures the potential. Both of these lines must be connected to the rotating electrode brushes. (Note that many older potentiostats use only one cable to carry both the drive and sense signals for the working electrode.)

    icon_info.jpg NOTE:

    Cell cables on newer model Pine potentiostats use RED to mark the working electrode “drive” line and ORANGE to mark the working electrode “sense” line. Both of these should be connected to the rotator brushes (see Figure 4.8).

    icon_info.jpg NOTE:

    Older Pine bipotentiostats use only one cable connection for both the “drive” and the “sense” signals. This connection is marked with a YELLOW banana jack on the bipotentiostat front panel.

    icon_bulb.jpg TIP:

    There is no universally accepted color coding scheme for marking potentiostat cell cable connections. If you are using the rotator with a third-party potentiostat, consult the potentiostat documentation for information about the cell cable markings.

    4.4.2 – RRDE Wiring

    The lower pair of brush contacts are only used with a rotating ring-disk electrode (see Figure 4.9). The lower pair of brushes (blue) contacts the ring electrode while the upper pair (red) contacts the disk electrode. Banana jumper cables are used to short together the opposing brushes in each pair to assure good contact with both sides of the rotating shaft.

    icon_bulb.jpg TIP:

    It is possible that the brush assemblies on a rotator that has been in use for some time may have been replaced or swapped, and thus, the colors of the brushes may not be as described in the previous paragraph or as shown in Figure 4.9. The important concept to remember is that the UPPER pair of brushes contacts the disk electrode, and the LOWER pair of brushes contacts the ring electrode.

    Figure-4.9-RRDE-Wiring-Scheme-Pine-Version.jpg

    Figure 4.9: Brush Connections for a Rotating Ring-Disk Electrode (RRDE)

    icon_bulb.jpg TIP:

    A bipotentiostat is required when working with a rotating ring-disk electrode. A bipotentiostat provides independent control of two different working electrodes in the same electrochemical cell.

    icon_info.jpg NOTE:

    Cell cables on newer model Pine bipotentiostats use RED and ORANGE to mark the first working electrode drive and sense lines, respectively. Both of these cables must be connected to the upper pair of electrode brushes (red) to contact the disk. Cell cables on newer model Pine bipotentiostats use BLUE and VIOLET to mark the second working electrode drive and sense lines, respectively. Both of these cables must be connected to the lower pair of electrode brushes (blue) to contact the ring.

    icon_info.jpg NOTE:

    Older Pine bipotentiostats use only one cable for both the “drive” and the “sense” signals. The YELLOW connection corresponds to the disk, and the BLUE connection corresponds to the ring.

    f4-10.jpg

    Figure 4.10: Stackable Banana Connector with Optional Stud Connector

    The jumper cables used to short the opposing brushes feature stackable banana plugs. If the cell cables from the potentiostat also terminate with banana plugs, then these plugs can simply be inserted directly into either end of the jumper cable. If the cell cables from the potentiostat terminate with alligator clips, then the easiest way to connect such alligator clips is to first insert a banana stud connector into the jumper cable (see Figure 4.10). The small tab on the banana stud provides a good place to attach the alligator clip.

    f4-11a.jpg   Figure-4.11B-routing-cell-cable.jpg

    Figure 4.11: Routing Cables out of the Enclosure

    4.4.3 – Routing Cables and Tubing

    The motor control cable may be routed out of the top of the enclosure to connect the motor unit to the control unit (see Figure 4.11). The enclosure has slots along the bottom of the window that provide clearance for routing cell cables and any tubing out of the enclosure. If required, cables and tubing may be routed through the back panel by drilling small holes in the panel. Any such drilled holes should have a diameter no greater than  13.0 \; mm ( 0.5 \; in ).

    4.5 – Proper Grounding

    To avoid issues with signal noise when making electrochemical measurements, it is important to properly ground all metal objects near an electrochemical cell to the earth ground. This generally includes the metal chassis of the instrumentation (potentiostat and rotator), the clamps and supports used to physically secure the electrochemical cell, and any peripheral equipment (heaters, stirrers, etc.) used in conjunction with the measurement.

    4.5.1 – Terminology

    icon_info.jpg NOTE:

    When working with electrochemical equipment, it is important to understand the meanings of terms such as “earth ground”, “chassis terminal”, and “DC common”.

    An earth ground connection is available in most modern laboratories via the “third prong” on the power receptacle for the local power system. The power system infrastructure for a laboratory building usually has a long metal probe buried in the earth, and the “third prong” in the building wiring is connected to this earth connection.

    A chassis terminal is a connection to the metal chassis surrounding an instrument. Depending upon how the instrument is connected to other experimental apparatus, a chassis terminal may or may not be connected to the earth ground.

    In the context of an electrochemical experiment involving a rotating electrode, the DC Common (also known as the signal ground or signal common) is the zero voltage reference point used by the signal measurement (or waveform generation) circuitry in the potentiostat and the rotation control circuitry in the rotator. The DC common may or may not be connected to the earth ground depending upon how the experimental apparatus is arranged and depending upon the internal circuitry of the potentiostat.

    4.5.2 – The Earth Ground Connection

    On the front panel of the MSR rotator control unit, there is an earth ground connection (see Figure 4.12). This connection point is in contact with the earth ground via the “third prong” of the power cord. As long as the power system in the laboratory (via the “third prong”) offers a robust connection to the earth, then this front panel connection can be used as an earth grounding point.

    It is also very important to note that the chassis of the rotator control unit is in direct contact with the earth ground connector. Thus, it is not possible to isolate the control unit chassis from the “third prong” earth ground.

    The chassis of the motor unit is also normally connected to the chassis of the control unit, and thus, to earth ground. This connection is usually made in an indirect fashion. Because the motor control cable (which connects the motor unit to the control unit) is a shielded cable, the shield assures that the chassis of the motor unit and the chassis of the control unit are electrically connected. And because the control unit chassis is connected to earth ground, the motor unit chassis is also in contact with earth ground.

    icon_info.jpg NOTE:

    Newer rotators have a shielded motor control cable which has HD-15 connectors on each end of the cable. The shield line in this cable assures that the chassis of the motor unit is in contact with the chassis of the control unit. The chassis of the control unit is, in turn, in contact with earth ground via the “third prong” on the power cord; thus, the motor unit chassis is also earth grounded.

    icon_info.jpg NOTE:

    Some older rotator models did not have a shielded motor control cable. These older, unshielded motor control cables are easily recognized because they are permanently connected to the control unit. If working with one of these older rotators, it is necessary to purposefully make a connection from the motor unit chassis to the earth ground connection on the front panel of the rotator control box.

    4.5.3 – A Typical Grounding Strategy

    While the details of proper grounding for any given electrochemical experiment may differ, a common approach when working with a rotating electrode cell is to connect the chassis of all instruments involved in the experiment to the earth ground at a common point. The earth ground connection on the front panel of the rotator control box serves as a convenient common point for such a strategy. There are two important grounding connections that should be considered in any rotating electrode experiment.

    The first important connection is to ground any metal clamps or cell supports near the electrochemical cell (see Figure 4.12). If left ungrounded, these metal objects may act as a source of environmental noise in the measured signals from the electrochemical cell. A simple banana cable with an alligator clip may be used to connect these metal objects to the earth ground.

    Figure-4.12-Grounded-Lab-Clamp.jpg

    Figure 4.12: Connect Metal Objects to Earth Ground on Control Box Front Panel

    The second important connection that is frequently (but not always) necessary is to connect the potentiostat to the earth ground (see Figure 4.13). If such a connection is to be made, it is important to understand what type of grounding connections are offered by the particular potentiostat being used (consult the potentiostat documentation). Most potentiostats offer either a chassis terminal connection or a signal ground (DC Common) connection. Some potentiostats offer both options, and if this is the case, only one of the two options should be used at a time, preferably the chassis terminal.

    Modern potentiostats are usually designed so that electrode connections (working, counter, and reference), the chassis, and the DC Common are all able to “float” with respect to the earth ground. This floating configuration is considered ideal because it gives the researcher maximum flexibility when working with electrochemical cells that may contain an earth grounded component. Compromising the floating configuration (by earth grounding either the chassis or the DC Common) should be avoided when possible, but there are cases where this is the only way to reduce noise in the measured electrochemical signals.

    Figure 4.13: Connect Potentiostat to Earth Ground on Control Box Front Panel

    If the potentiostat offers a chassis terminal connection (often located on the back panel of the potentiostat), then connecting this chassis terminal to the earth ground on the rotator control box may (or may not) reduce or eliminate noise in the measured electrochemical signals. For a potentiostat with a chassis that normally “floats” with respect to earth ground, the chassis will no longer be floating after making this direct connection to earth ground.

    icon_info.jpg NOTE:

    The chassis terminal on the back panel of the Pine WaveDriver 10 and WaveDriver 20 potentiostats normally “floats” with respect to the earth ground. If the potentiostat chassis is connected to the earth ground on the front panel of the rotator control unit, the potentiostat no longer “floats” with respect to earth ground.

    If the potentiostat does not offer a chassis terminal connection (or if earth grounding the chassis terminal is not an option), then a less desirable, but alternative approach is to attempt to make use of the potentiostat’s signal ground (DC Common). If the DC Common (often provided as an alligator clip connection on the cell cable) is connected to the earth ground on the rotator control box, this may (or may not) reduce or eliminate signal noise.

    In general, in an electrochemical experiment, it is always ideal to maintain as much isolation between the DC Common, the earth ground, and the instrument chassis as possible. Such isolation increases flexibility when working with electrochemical cells that may contain earth grounded components. But in almost all cases, a cell containing a rotating electrode does not have any earth grounded electrodes, allowing some trade-off between flexibility and the use of earth grounding to reduce signal noise.

    4.6 – Using the Rotator in a Glove Box

    The rotator may be placed in a glove box when working with air or moisture sensitive compounds. A smaller base (sold separately) is available for use in a glove box (see Figure 4.14).

    It is important to understand that the low humidity environment found in most glove boxes increases the rate of wear on both the brush contacts and the internal brushes within the motor itself.

    To mitigate the wear rate of the brush contacts, it is recommended that four special low-humidity brushes (sold separately) be installed prior to placing the rotator in the glove box. Contact Pine Research for more details.

    f4-13.jpg

    Figure 4.14: Glove Box Configuration

    icon_caution.jpg CAUTION:

    Using the rotator in a dry environment such as a low humidity glove box will increase the wear rate of the internal motor brushes.
    See Section 6.5 for more information about how to replace a worn motor.

    ATTENTION:

    L’utilisation du rotateur dans un environnement sec, tel qu’une boîte à gants à faible taux d’humidité augmente la vitesse d’usure des balais internes du moteur.
    Consultez la section 6.5 pour en savoir plus sur la manière de remplacer un moteur usé.

    4.7 – Rotation Rate Control

    icon_caution.jpg CAUTION:

    Always turn the rotation rate control knob completely counterclockwise (towards the zero rotation rate position) before turning on the rotator.

    ATTENTION:

    Tournez toujours le bouton de commande de la vitesse de rotation complètement dans le sens inverse des aiguilles d’une montre (vers la position vitesse de rotation égale à zéro) avant de mettre le rotateur en marche.


    icon_info.jpg NOTE:

    The fully counterclockwise position corresponds (nominally) to a rotation rate of zero. Even with the knob in this position, there may be some residual rotation (typically less than  \textbf{10 RPM} ) in either the clockwise or counter-clockwise direction.

    Always begin each session using the rotator with the power turned off and the rotation rate control knob in the fully counterclockwise position. The fully counterclockwise position corresponds to the slowest rotation rate, and it is always safest to turn on the rotator with the knob in this position.

    4.7.1 Manual Control of Rotation

    To rotate the electrode under manual control, turn on the control unit power and slowly turn the rotation rate control knob clockwise. As the knob is turned clockwise, the rotation rate increases and the display on the control unit shows the rotation rate.

    4.7.2 Monitoring the Rotation Rate

    The rotation rate is always displayed on the front panel, but it can also be monitored at the output jacks on the front panel of the control unit. The signal presented at the output jacks is a voltage which is proportional to the rotation rate. The proportionality ratio is  1.0 \; mV/RPM .

    icon_info.jpg NOTE:

    The rotation rate is controlled to within  \textbf{1.0\%} of the display value selected using the rotation rate control knob. It is normal for the last one or two digits on the display to flicker.

    4.7.3 – External Control of the Rotation Rate

    It is often convenient for the rotation rate to be controlled via an externally supplied signal. Many potentiostats are capable of providing such a signal to control the rotation rate while simultaneously performing electrochemical measurements. An externally supplied signal is also required when performing hydrodynamically modulated voltammetry, where the rotation rate is varied sinusoidally as electrochemical measurements are made with the potentiostat.

    Figure-4.15A-Rate-Control-Cable-Pine-Version.jpg
     
    Figure-4.15B-Rate-Control-Cable-Pine-WaveNow.jpg   Figure-4.15C-Rate-Control-Cable-Pine-WaveDriver.jpg

    Figure 4.15: Connecting the Rotation Rate Control Cable

    A special cable is available to connect Pine potentiostats to the MSR rotator (see Figure 4.15). One line of this cable carries an analog rate control signal from the potentiostat to the INPUT jacks on the front panel of the control box. A second line carrying a digital control signal is connected to the MOTOR STOP jack on the back panel of the control box. The other end of the special cable is connected to the potentiostat.

    The analog rate control signal from the potentiostat is a voltage that is proportional to the desired rotation rate. The MSR rotator is factory configured to use a  \text{1.0 RPM/mV} ratio, which is the ratio compatible with Pine potentiostats. Other ratios are available for use with other potentiostats (see Section 6.7).

    External control of the rotation rate may involve a signal connection between a potentiostat from one manufacturer being connected to a rotator from another manufacturer. The signals on these various instruments may have been calibrated to different tolerances by each manufacturer. Small signal level differences within these tolerances can add up, causing the actual rotation rate (as displayed on the control unit) to differ slightly from the specified rotation rate (as entered by the user of the potentiostat software).

    4.7.4 – External Motor Stop Control

    An external digital signal can be applied across the MOTOR STOP banana jacks on the back panel to bring the rotator to a complete stop (see Figure 2.4). This digital signal can be used by a potentiostat or other external instrument to assure that the rotation rate is actually zero. The logic for this digital signal may be either “active HIGH” or “active LOW”.

    For the Pine MSR rotator, the MOTOR STOP is configured at the factory to use “active HIGH” logic. If desired, a jumper setting inside the control box can be configured to use the opposite logic (see Section 6.8 for details).

    If the MOTOR STOP logic is configured to be “active HIGH”, then the motor is allowed to rotate if a signal greater than  \text{2.0 volts} is applied across the MOTOR STOP banana jacks. If the two banana jacks are shorted together (i.e., if the MOTOR STOP stop signal is driven to ground), then the motor stops rotating.

    If the MOTOR STOP logic is configured to be “active LOW”, then the motor will stop if a signal greater than 2.0 volts is applied across the MOTOR STOP banana jacks. If the two banana jacks are shorted together (i.e., if the MOTOR STOP signal is driven to ground), then the motor is allowed to rotate.

    icon_info.jpgNOTE:

    When the control unit is configured for “active HIGH” logic and when no connections are made to the MOTOR STOP banana jacks, the motor is allowed to rotate. An internal “pull up” circuit assures that the motor stop signal remains “high” in this case.

    4.8 – Circuit Protection

    The power switch on the back panel also acts as a circuit breaker to help protect the control unit circuitry. If the circuit breaker trips, then it can be reset by turning the power switch to the full “off” position and then turning the switch back “on” again.

    A secondary circuit breaker on the front panel protects the windings in the motor. If this circuit breaker trips, then the circuit breaker can be reset by pressing the “RESET” button on the front panel.

    Read More
  • MSR Rotator: Power Supply


    The rotator control unit has two internal power supplies. One of these power supplies is  \text{30 VDC (100 W)} and the other is  \text{24 VDC (65 W)} . These power supplies are located within the high voltage section of the control unit (see below).

    Figure-7-2-MSR-power-supplies.jpg


    electric_shock.jpg

    WARNING:

    Risk of electric shock.
    Disconnect power before servicing the control unit.
    The shaded region in the figure above shows the location of the high voltage power entry and the two internal power supplies.

    AVERTISSEMENT:

    Risque de décharge électrique.
    Déconnectez l’alimentation avant de procéder à l’entretien de l’unité de commande.
    La région ombragée de la figure (ci-dessus) présente l’emplacement de l’entrée de puissance à haute tension et des deux alimentations internes.

    In the (rare) event that it is necessary to replace one of the internal power supply units, it is important to carefully identify which type of power supply is to be replaced. First, determine which supply ( \text{30 V} or  \text{24 V} ) is to be replaced (see above).

    Next, carefully examine the power supply to determine which brand of power supply is to be replaced. Early rotators (prior to October 2014) use power supplies manufactured by Astrodyne, FranMar International or TDK Lambda. Later rotators (after October 2014) use power supplies manufactured by XP Power. Depending upon when the rotator was manufactured, choose the appropriate replacement part shown below.

    Figure7_3Website.jpg


    electric_shock.jpg

    WARNING:

    Risk of electric shock.
    Disconnect all power before servicing the rotator.

    AVERTISSEMENT:

    Risque de décharge électrique.
    Déconnectez toutes les sources d’alimentation avant de procéder à l’entretien du rotateur.


    rotating-hazard.jpg

    WARNING:

    Rotating shaft. Entanglement hazard.
    Turn off the power to the rotator and disconnect the power cord from the power source before continuing with this procedure.

    AVERTISSEMENT:

    Arbre en rotation. Danger d’enchevêtrement.
    Éteignez le rotateur et débranchez le cordon d’alimentation de la source d’alimentation avant de poursuivre cette procédure.

    Read More
  • previous 10
    Next 10