Nernst equation (derivation)

In 1887, Walther Nernst, a German chemist, developed the Nernst equation, which describes the relationship between an electrochemical reaction’s potential and its standard electrode potential under standard and non-standard conditions.

The derivation of the Nernst equation involves the following steps:

    1. Derive the formula for the reversible open circuit potential , Eo, of an electrochemical cell at constant temperature and pressure under standard conditions.
    2. Derive the expression for Eo in terms of chemical potentials of the reaction species at constant temperature and pressure.
    3. Derive a ‘correction factor’ for Eo in step2. This final expression, the Nernst equation, describes how the resultant Eo varies with activities of chemical species.

 

Step 1

From the definition of enthalpy, H = U + pV, we have

dH=dU+d(pV)\; \; \; \; \; \; \; \; 1

Substitute the definition of the internal energy of a system undergoing reversible change, dU = dqrev + dwrev in eq1

dH=dq_{rev}+dw_{rev}+d(pV)\; \; \; \; \; \; \; \;2

Substitute the definition of the change in Gibbs energy of a system, dG = dHd(TS) in eq2

dG+TdS+SdT=dq_{rev}+dw_{rev}+pdV+Vdp

At constant temperature and pressure, dT = dp = 0

dG=dq_{rev}+dw_{rev}+pdV-TdS\; \; \; \; \; \; \; \; 3

Substitute the definition of the change in entropy of a system, dS = dqrev/T in eq3

dG=dw_{rev}+pdV\; \; \; \; \; \; \; \; 4

Substitute the definition of total reversible work, wrev = wex + wadd, where wex is reversible expansion work and wadd is reversible additional work other than reversible expansion work, in eq4

dG=dw_{ex}+dw_{add}+pdV\; \; \; \; \; \; \; \; 5

Substitute the definition of the change in reversible expansion work, dwex = –pdV in eq5

dG=dw_{add}\; \; \; \; \; \; \; \; 6

Substitute the definition of the reversible electrical work between two points in an electric circuit, dwadd = –nFEdξ in eq6

\Delta _rG=-nFE\; \; \; \; \; \; \; \; 7

where n is the number of moles of electrons, F is the Faraday constant, E is the open circuit potential of the electrochemical cell and ΔrG = dG/ is the Gibbs energy of the electrochemical cell reaction.

When the open circuit potential of an electrochemical cell is measured at standard conditions (1 bar, 298K, 1M), we write eq7 as:

\Delta _rG^{\, o}=-nFE^{\, o}\; \; \; \; \; \; \; \; 8

 

Step 2

The chemical potential of the j-th species, in a multi-component system is

\mu_j=\left (\frac{\partial G}{\partial n_j} \right )_{T,p,n'}\; \; \; \; \; \; \; 9

where n‘ is all n other than nj.

The total differential of the Gibbs energy of the system at constant temperature and pressure is

dG=\left (\frac{\partial G}{\partial n_a} \right )_{T,p,n'}dn_a+\left (\frac{\partial G}{\partial n_b} \right )_{T,p,n'}dn_b+...\; \; \; \; \; \; \; 10

Substituting eq9 in eq10

dG=\mu_adn_a+\mu_bdn_b+...=\sum_{j}\mu_jdn_j\; \; \; \; \; \; \; \; 11

Eq11 can be expressed in another way by considering the reversible reaction

v_aa+v_bb+...\rightleftharpoons v_{\alpha}\alpha+v_{\beta}\beta+...

with the change in Gibbs energy at constant temperature and pressure of the system given by eq11. Next, let’s define

dn_j=v_jd\xi\; \; \; \; \; \; \; \; 12

where vj is the stoichiometric number of the j-th species in the reversible reaction and is the amount of substance that is being changed in the reaction.

ξ is called the extent of a reaction and has units of amount in moles. Note that the stoichiometric numbers for reactants are negative by convention and those for products are positive. Substituting eq12 in eq11

\left ( \frac{\partial G}{\partial \xi} \right )_{T,p}= \sum_{j}\mu_jv_j\; \; \; \; \; \; \; \; 13

Substitute the definition of the Gibbs energy of an electrochemical cell reaction, \Delta _rG=\left ( \frac{\partial G}{\partial \xi} \right )_{T,p} , into eq13

\Delta _rG=\sum_{j}\mu_jv_j\; \; \; \; \; \; \; \; 14

Substitute eq7 from Step1 in eq14

E=-\frac{1}{nF}\sum_{j}\mu_jv_j\; \; \; \; \; \; \; \; 15

When the open circuit potential of an electrochemical cell is measured at standard conditions (1 bar, 298K, 1M), we write eq15 as:

E^{\, o}=-\frac{1}{nF}\sum_{j}\mu^o_{\, j}v_j\; \; \; \; \; \; \; \; 16

 

Step 3

Firstly, let’s consider a system with a single pure substance. Substitute the definition of the change in Gibbs energy of a system, dG = dHd(TS), in the change of enthalpy of the system dH = dU + d(pV)

dG+SdT+TdS=dU+pdV+Vdp\; \; \; \; \; \; \; \; 17

The reversible change in internal energy of a pure substance (a system with a single pure substance only does expansion work) can be expressed as

dU=dq_{rev}+dw_{ex}\; \; \; \; \; \; \; \; 18

Substitute the definitions of entropy, dS = dqrev/T, and expansion work, –pdV, in eq18

dU=TdS-pdV\; \; \; \; \; \; \; \; 19

Substitute eq19 in eq17

dG=Vdp-SdT\; \; \; \; \; \; \; \; 20

At constant temperature, dT = 0 and eq20 becomes dG = Vdp. Integrating this new expression on both sides from p_i to p_f,

G\left ( p_f \right )=G\left ( p_i \right )+nRTln\frac{p_f}{p_i}

Let pi = 1 bar = po (i.e. standard conditions) and pf = p

G=G^{\, o}+nRTln\frac{p}{p^{\, o}}

Secondly, we extend the above working to a multi-component system of ideal gases. Since G is an extensive property, we add the Gibbs energies of all components to give the total Gibbs energy of the system:

G_{total}=\left ( G_a^{\: o}+n_aRTln\frac{p_a}{p_a^{\, o}} \right )+\left ( G_b^{\: o}+n_bRTln\frac{p_b}{p_b^{\, o}} \right )+...

Let’s assume the reference pressures of the component gases are the same since they are all ideal gases, i.e. pao = pbo = p

G_{total}=\left ( G_a^{\, o}+G_b^{\, o}+...\right )+\left ( n_aRTln\frac{p_a}{p^o}+n_bRTln\frac{p_b}{p^o}+...\right )

G_{total}=G_{total}^{\, o}+\left ( n_aRTln\frac{p_a}{p^o}+n_bRTln\frac{p_b}{p^o}+...\right )

For simplicity, let GTotal and GoTotal be G and Go respectively.

G=G^{\, o}+\left ( n_aRTln\frac{p_a}{p^o}+n_bRTln\frac{p_b}{p^o}+...\right )

Taking the partial derivative of the above with respect to nj, at constant temperature, pressure and the amount of the other components, n‘, we have

\left ( \frac{\partial G}{\partial n_j} \right )_{T,p,n'}=\left ( \frac{\partial G^{\, o}}{\partial n_j} \right )_{T,p,n'}+\left [ \frac{\partial \left ( n_aRTln\frac{p_a}{p^{\, o}}+ n_bRTln\frac{p_b}{p^{\, o}}+...\right )}{\partial n_j} \right ]_{T,p,n'}

For the last partial derivative of the above equation, only the j-th term survives.

\left ( \frac{\partial G}{\partial n_j} \right )_{T,p,n'}=\left ( \frac{\partial G^{\, o}}{\partial n_j} \right )_{T,p,n'}+RTln\frac{p_j}{p^{\, o}}\; \; \; \; \; \; \; \; 21

Substitute eq9 from Step 2 in eq21

\mu_j=\mu_j^{\, o}+RTln\frac{p_j}{p^{\, o}}\; \; \; \; \; \; \; \; 22

Similarly, for a solute in a dilute solution satisfying Henry’s law, we can write

\mu_j=\mu_j^{\, o}+RTln\frac{c_j}{c^{\, o}}\; \; \; \; \; \; \; \; 23

where cj and co are the j-th solute’s concentration and molar concentration respectively.

We can fine-tune eq22 and eq23 by replacing pj/po and cj/co with γj(pj/po) and γj(cj/co) where γj is a factor called the activity coefficient that can account for both ideal and non-ideal fluids (γ = 1 and γ < 1 for ideal and non-ideal fluids respectively). Next, we combine the modified equations into a single equation by letting aj = γj(pj/po) and aj = γj(cj/co), where aj is the activity of species j.

\mu_j=\mu_j^{\, o}+RTlna_j\; \; \; \; \; \; \; \; 24

Substituting eq24 in eq15 from Step 2,

E=-\frac{1}{nF}\sum_{j}\left ( \mu_j^{\, o}+RTlna_j \right )v_j=-\frac{1}{nF}\sum_{j} \mu_j^{\, o}v_j-\frac{RT}{nF}\sum_{j}v_jlna_j \; \; \; \; \; \; \; \; 25

Substituting eq16 from Step 2 in eq25

E=E^{\, o}-\frac{RT}{nF}\sum_{j}ln\, a_j^{\, v_j}=E^{\, o}-\frac{RT}{nF}ln\prod_{j}a_j^{\, v_j}

E=E^{\, o}-\frac{RT}{nF}lnQ\; \; \; \; \; \; \; \; 26

where Q is the reaction quotient with Q\prod _{j}a_j^{\, v_j}. 

Hence, \frac{RT}{nF}lnQ is the ‘correction factor’ for Eo, as the activities of reaction species vary. Eq26 is the Nernst equation.

For an electrochemical equation of the form Ox+ne^-\rightleftharpoons Red, the Nernst equation can be written as:

E=E^{\, o}-\frac{RT}{nF}ln\frac{a_{ Red}}{a_{ Ox}}\; \; \; \; \; \; \; \; 27

Note that since Eo is by definition a reduction potential, the reference equation is always Ox+ne^-\rightleftharpoons Red and hence eq27.

 

Next article: pH glass electrode
Content page of advanced electrochemistry
Content page of Advanced chemistry
Main content page

pH measurement using a glass electrode

A typical experiment setup for pH measurement consists of a glass electrode (indicator electrode), a reference electrode (a regular AgCl electrode), the analyte and a potentiometer. The potentiometer ensures that current flow is negligible in the electrochemical circuit so that IR drop and overpotential are minimised, resulting in more accurate measurements. Such a method in electrochemistry where current is negligible is called potentiometry.

Comparing the setup with a circuit diagram (see above diagram), the overall potential of the electrochemical cell is:

E_{cell}=E_{glass\: elec}-E_{ext\: ref\: elec}+E_{liq\: jn}\; \; \; \; \; \; \; 31

The liquid junction potential arises due to the difference in mobility of K+ and Cl in the controlled flow of the saturated KCl solution out of the porous frit of the external reference AgCl electrode (to complete the circuit). Eanalyte = IR = 0 since current flow is negligible in potentiometric methods.

 

Question

How does the potentiometer control current in the circuit?

Answer

A potentiometer works by sliding a contact along a resistor and finding the relationship between VL and Ecell (see diagram below).

For a pH measurement experiment, the contact moves to the point where no current flows. Ecell is then calculated by:

E_{cell}=\frac{R_1R_L+R_2R_L+R_1R_2}{R_2R_L}V_L

In a modern pH meter, a computer controls the potentiometer.

 

Lastly, before a pH meter is put to use, it is calibrated using solutions of known pH. The reason for this is as follows:

Substituting eq30 from the previous article in eq31,

E_{cell}=L'-0.0592\, pH

where L’ = LEext ref elec + Eliq jn.

Since Eliq jn cannot be calculated or measured directly, the value of L’ must be determined by calibrating the pH meter with solutions of known pH.

 

Next article: IR drop
Previous article: pH glass electrode
Content page of advanced electrochemistry
Content page of Advanced chemistry
Main content page

IR drop (electrochemistry)

IR drop in electrochemistry is the voltage drop due to the resistance of the analyte (and the internal KCl solution if a calomel or AgCl electrode is used) in a close circuit.

In the previous article regarding pH measurement, which involves potentiometry, no net current flows in the circuit. When a current flows, e.g. in coulometry and electrogravimetry, we have to consider two effects:

We can represent IR drop as an internal resistance of a battery in the following circuit diagram:

The general equation for a galvanic cell is:

E_{galvanic}=IR=E_{cell,open}-Ir\; \; \; \; \; \; \; \; 32

where Ecell, open = EREL.

Eq32 when applied to a Zn electrochemical cell gives the following values:

Ecell, open and Egalvanic for the above example are recorded as negative values by the voltmeters because of the way the terminals of the voltmeters are connected to the respective circuits. From eq32, \left | E_{galvanic} \right | is always less than \left | E_{cell,open} \right |.

For an electrolytic cell,

E_{applied}=E_{cell,open}-Ir\; \; \; \; \; \; \; \; 33

If no current is flowing in the electrolytic cell, eq33 reduces to Eapplied = Ecell, open. For electrolysis to proceed, \left | E_{applied} \right |>\left | E_{cell,open} \right |. For example, the variation of Eapplied for the electrolytic reduction of Zn2+ is illustrated by the following diagram:

Once again, Eapplied for the above example is recorded as a negative value by the voltmeter because of the way the voltmeter is connected to the circuit.

 

Question

With reference to the above electrochemical cell that is composed of a saturated AgCl reference electrode, a Zn working electrode and a 0.5 M Zn2+ electrolyte at rtp, what external voltage must be applied to generate an electrolytic current of 3 mA if the internal resistance is 10 Ω. What is the IR drop?

Answer

Zn^{2+}(aq)+2e^-\rightleftharpoons Zn(s)\; \; \; \; \; \; \; E^{\, o}=-0.762\: V

AgCl(s)+e^-\rightleftharpoons Ag(s)+Cl^-(aq)\; \; \; \; \; \; \; E^{\, o}=0.199\: V

Using the Nernst equation,

 

E_{cell,open}=-0.762-\frac{0.0592}{2}log \frac{1}{0.5}-0.199=-0.970\: V

Using eq33,

E_{applied}=-0.970-(0.003)(10)=-1.000\: V

Therefore, the IR drop is

 

Next article: Polarisation
Previous article: pH measurement using a glass electrode
Content page of advanced electrochemistry
Content page of Advanced chemistry
Main content page

pH glass electrode

A pH glass electrode is an ion-selective electrode that is sensitive to H+, and therefore, used to measure pH of solutions. It is composed of a glass bulb that is H+-selective and an internal AgCl electrode.

The glass bulb is specially manufactured with a composition of 70+% SiO2 and 20+% Na2O. It has a silicate network within the glass matrix, with some Si atoms singly bonded to O atoms and other Si atoms coordinated with Na+ (see diagram above). On the outer and inner surfaces of the bulb, some of the SiO arms are also coordinated with Na+, while the rest of the SiO are uncoordinated.

Before the pH glass electrode can be used, it must be hydrated (soaked in water of pH = 7) to form two layers of gel. This is to allow H+ to diffuse into the gel layers to convert the SiONa+ groups on both surfaces to SiOH+. Since the inner surface of the bulb is already in contact with a KCl buffer solution at pH 7, both the inner and outer surfaces develop the same ion-exchange equilibrium when the electrode is dipped in water:

H^++SiO^-Na^+\rightleftharpoons Na^++SiO^-H^+

This ion-exchange occurs because H+ binds more strongly to SiO than Na+. Therefore, the position of the equilibrium is almost entirely to the right. On the inner surface, the displaced sodium ions enter the buffer solution, and on the outer surface, displaced sodium ions enter water. The hydrated surfaces now consist of a combination of SiOH+ and SiOarms, and is ready for use. The electrode is immersed in an analyte of unknown pH, establishing the following equilibrium on the outer surface:

\begin{matrix} H^+_{\; \; analyte}+SiO^-\\ outer\; activity \end{matrix} \rightleftharpoons\; \begin{matrix} SiO^-H^+_{\; \; analyte}\\ glass\; activity \end{matrix}

and the following equilibrium on the inner surface equilibrium:

\begin{matrix} H^+_{\; \; buffer}+SiO^-\\ inner\; activity \end{matrix} \rightleftharpoons\; \begin{matrix} SiO^-H^+_{\; \; buffer}\\ glass\; activity \end{matrix}

If the pH of the analyte is less than 7, the outer surface will have fewer uncoordinated SiO arms and hence, less negative charges as compared to the inner wall, whose equilibrium remains unchanged. The difference in charges on the surfaces produces a potential difference across the glass bulb. Furthermore, it confers enough activation energy for sodium ions within the glass matrix to dissociate from the silicate network. These mobile sodium ions then migrate interstitially across the bulb and along the potential gradient, to complete the circuit.

Even though the above equilibria are ion-exchanges and not redox reactions, the potential across the bulb Eb is found empirically to be Nernst-like, where:

E_{outer}=E^{\, o}_{\; glass}-\frac{RT}{nF}ln\frac{a_{glass}}{a_{\, outer}}

E_{inner}=E^{\, o}_{\; glass}-\frac{RT}{nF}ln\frac{a_{glass}}{a_{\, inner}}

E_b=E_{outer}-E_{inner}=-\frac{RT}{nF}ln\frac{a_{inner}}{a_{\, outer}}\; \; \; \; \; \; \; \; 28

 

Question

Why Eb = EouterEinner and not Eb = EinnerEouter?

Answer

We write Eb = Eouter  Einner to be consistent with the way we write Ecell ER – EL, which follows the IUPAC convention (consistency is best visualised by tracing the direction of electron flow in the circuit).

 

Comparing with a circuit diagram (see above diagram), the overall potential of the glass electrode is:

E_{glass\: electrode}=E_b-E_{int \, soln}+E_{int\: elec}

where Eint elec is the potential of the internal AgCl electrode and Eint soln is the potential drop across the saturated KCl solution.

Since no current flows in potentiometric methods,

E_{glass\: electrode}=E_b+E_{int\: elec}

Furthermore, manufacturing defect or damage of electrode over time results in the electrode recording a non-zero potential when the analyte is pH = 7. We call this an asymmetry potential, which must be included in Eglass:

E_{glass\: electrode}=E_b+E_{int\: elec}+E_{asym}\; \; \; \; \; \; \; \; 29

Substituting eq28 in eq29, where ainner, Easym and Eint elec are constants,  Eglass eletrode at rtp is:

E_{glass\, electrode}=L+0.0592\, log\, a_{outer}=L-0.0592\, pH\; \; \; \; \; \; \; \; 30

where L = Eint elec + Easym – 0.0592 log ainner.

Question

What is the function of the internal AgCl electrode?

Answer

For the pH glass electrode to work in a pH measurement experiment, there must be 2 electrodes, where redox reactions occur (to generate and receive electrons). The internal AgCl electrode serves as one of the electrodes, while the reference electrode (an external regular AgCl electrode) serves as the other electrode. The bulb, which is the pH-sensing element, can be perceived as an extension of the internal electrode.

Next article: pH measurement using a glass electrode
Previous article: Nernst equation
Content page of advanced electrochemistry
Content page of Advanced chemistry
Main content page

Polarisation (electrochemistry)

Polarisation is the deviation of the electrode potential from its equilibrium value, Eeqm , when a certain level of current flows. It is composed of two components: activation (or kinetic) polarisation and concentration polarisation.

Activation polarisation is the change in electrode potential from its equilibrium value when a certain level of current flows, resulting from the overcoming of the activation energy of the rate-determining step in an electrochemical reaction. Concentration polarisation is the change in electrode potential from its equilibrium value when a certain level of current flows, caused by inefficiencies in electrode reactions due to sluggish mass transfer between the electrode surface and the solution.

The extent of the polarisation of an electrode or an electrochemical cell is measured by a quantity called overpotential, η, where

\eta=\eta_{act}+\eta_{con}=E-E_{eqm}\; \; \; \; \; \; \; \; 34

It is difficult to directly measure the individual components of overpotential when an electrochemical cell is operating. However, by making certain assumptions, it is possible to derive formulas attributing to ηact and ηcon, and design experiments based on these assumptions to measure the two quantities.

 

Question

Why is IR drop not part of the overpotential equation?

Answer

Overpotential is a consequence of the reactions at an electrode while the IR drop is characteristic of the bulk solution.

 

Next article: Activation Polarisation
Previous article: IR drop
Content page of advanced electrochemistry
Content page of Advanced chemistry
Main content page

Activation polarisation (electrochemistry)

Activation polarisation is the change in electrode potential from its equilibrium value when a current flows, as a result of overcoming the activation energy of the rate-determining step in an electrochemical reaction.

To understand the concept of activation polarisation, let’s analyse an experiment with the following electrode reaction:

Ox+e^-\begin{matrix} r_{red}\\\rightleftharpoons \\ r_{ox} \end{matrix}Red\; \; \; \; \; \; \; \; 35

where rred is defined as the flow of the number of moles of Ox per unit area of the electrode per unit time, i.e. the rate of reduction of Ox, and rox is the flow of the number of moles of Red per unit area of the electrode per unit time, i.e. the rate of oxidation of Red. We represent rred and rox with the following rate equations:

r_{red}=k_c[Ox]\; \; \; \; \; \; \; \; 36

r_{ox}=k_a[Red]\; \; \; \; \; \; \; \; 37

where kc and ka are the cathodic rate constant and anodic rate constant respectively (each with units of length per time). From the combined Faraday’s laws of electrolysis, we have

m=\frac{Q}{F}\left ( \frac{M}{z} \right )\; \; \Rightarrow \; \; zF\frac{m}{M}=It\; \; \Rightarrow \; \; zF\frac{no.\: of\: moles}{t}=I\; \; \; \; \; \; \; \; 38

Divide eq38 throughout by the area of electrode,

\frac{I}{A}=zF\frac{no.\; of\; moles}{At}\; \; \Rightarrow \; \; j=zFr\; \; \; \; \; \; \; \; 39

where j = I/A, is the current density. With reference to eq35, z = 1, and so we can rewrite eq36 and eq37 as:

j_c=Fk_c[Ox]\; \; \; \; \; \; \; \; 40

j_a=Fk_a[Red]\; \; \; \; \; \; \; \; 41

where jc and ja are the cathodic current density and the anodic current density respectively.

To apply the above equations, consider a piece of metal M immersed in an M+ solution. Over time, an electric double layer is formed, with the development of an equilibrium electrode potential between the metal and its ionic solution, such that rred = rox or jc = ja (i.e. no net current flowing). If we connect the M+/M half-cell to an SHE, the circuit is closed.

Assuming that M lies below hydrogen in the electrochemical series, a net current flows towards the SHE, i.e. electrons flow towards M, the cathode. This results in a less positive electrode reduction potential and the equilibrium in eq35 shifting to the right, with rred > rox. Since the electrode potential deviates from its equilibrium value when a current flows, we say that the electrode is polarised.

In terms of current density, jc is no longer equal to ja, and because it is a reduction reaction, jc > ja, i.e. a net cathodic current flows. The net current density j is given by:

j=j_a-j_c=Fk_a[Red]-FK_c[Ox]\; \; \; \; \; \; \; \; 42

 

Question

Why is the net current density jajc and not jcja?

Answer

This is to be consistent with IUPAC’s definition that a net anodic current is positive, while a net cathodic current is negative (as mentioned above, the current is cathodic when jc > ja, which makes j < 0).

 

Substituting the Eyring equation k=Ae^{-\frac{\Delta ^{\ddagger'}G}{RT}} where \Delta^{\ddagger'}G is the activation Gibbs energy in eq42,

j=FA_a\left [ Red \right ]e^{-\frac{\Delta ^{\ddagger'}G_a}{RT}}-FA_c\left [ Ox \right ]e^{-\frac{\Delta ^{\ddagger'}G_c}{RT}}\; \; \; \; \; \; \; \; 43

with

j_a=FA_a\left [ Red \right ]e^{-\frac{\Delta ^{\ddagger'}G_a}{RT}}\; \; \; \; \; \; \; \; 44

j_c=FA_c\left [ Ox \right ]e^{-\frac{\Delta ^{\ddagger'}G_c}{RT}}\; \; \; \; \; \; \; \; 45

When the electrode is at equilibrium, ja = jc and we can rewrite eq44 and eq45 as:

j_0=FA_a[Red]e^{-\frac{\Delta ^\ddagger G_a}{RT}}\; \; \; \; \; \; \; \; 46

j_0=FA_c[Ox]e^{-\frac{\Delta ^\ddagger G_c}{RT}}\; \; \; \; \; \; \; \; 47

where j0 denotes the equal current densities at equilibrium and is called the exchange current density; and \Delta^{\ddagger}G_a and \Delta^{\ddagger}G_c are the anodic activation Gibbs energy at equilibrium and the cathodic activation Gibbs energy at equilibrium respectively.

Next, we shall investigate the energy profile of the reaction of eq35 at the cathode at equilibrium (see Fig I below). For simplicity, we assume that the transition state has equal resemblance to M+ and M and therefore the activation Gibbs energy required for the reduction reaction is the same as that of the oxidation reaction.

As mentioned, when the electrode is connected to an SHE, its potential becomes less positive, which in turn causes a decrease of the activation Gibbs energy for the reduction reaction (electrode has stronger attraction for M+) and an increase in the activation Gibbs energy for the oxidation reaction (more difficult for M+ to form). The resultant energy profile is shown in Fig II. Note that the peaks in Fig I and Fig II are aligned at the same Gibbs energy level for convenience.

 

Question

Use eq44 and eq45 to justify the change in activation Gibbs energies for the reduction and oxidation reaction at the cathode when a current flows

Answer

When a cathodic current flows, jc > ja, i.e. cathodic current density level increases versus that at equilibrium. Assuming that the electrolyte is well stirred such that [Ox] remains constant, \Delta ^\ddagger G_c in eq45 must decrease. Simultaneously, the reduction in anodic current density at the same electrode results in an increase in \Delta ^\ddagger G_a (see eq44 where [Red] is constant, as it is a solid).

 

By letting \left | \Delta G \right | be the magnitude of the total change in Gibbs energy between the state when a current flows and the state when the electrode is at equilibrium, we have:

\Delta ^{\ddagger}G_c=\Delta ^{\ddagger'}G_c+\alpha \left | \Delta G \right |\; \; \; \; \; \; \; \; 48

\Delta ^{\ddagger}G_a=\Delta ^{\ddagger'}G_a-\left ( 1- \alpha\right )\left | \Delta G \right |\; \; \; \; \; \; \; \; 49

where \alpha is a fraction of \left | \Delta G \right | and is called the transfer coefficient. If you recall, the formula \Delta _rG=-nFE is derived based on the definition of the reversible electrical work between two points in a circuit with a potential difference E. The change in Gibbs energy in our case is due to the change in electrode potential when a current flows, EEeqm and so, we have:

\left | \Delta G \right |=-nF\left ( E-E_{eqm} \right )\; \; \; \; \; \; \; \; 50

Substituting eq34 from the previous article in eq50, the magnitude of the change in Gibbs energy for the passage of one mole of electrons (n = 1) is:

\left | \Delta G \right |=-F\eta\; \; \; \; \; \; \; \; 51

Substituting eq51 in eq48 and eq49,

\Delta^{\ddagger}G_c=\Delta^{\ddagger'}G_c-\alpha F\eta\; \; \; \; \; \; \; \; 52

\Delta^{\ddagger}G_a=\Delta^{\ddagger'}G_a+\left ( 1- \alpha\right )F\eta\; \; \; \; \; \; \; \; 53

Substituting eq53 and eq52 in eq44 and eq45 respectively,

j_a=FA_a[Red]e^{-\frac{\Delta^\ddagger G_a}{RT}}e^{\frac{(1-\alpha)F\eta}{RT}}\; \; \; \; \; \; \; \; 54

j_c=FA_c[Ox]e^{-\frac{\Delta^\ddagger G_c}{RT}}e^{\frac{-\alpha F\eta}{RT}}\; \; \; \; \; \; \; \; 55

Substituting eq46 and eq47 in eq54 and eq55 respectively,

j_a=j_0e^{(1-\alpha)f\eta}\; \; \; \; \; \; \; \; 56

j_c=j_0e^{-\alpha f\eta}\; \; \; \; \; \; \; \; 57

where f = F/RT

The net current density is obtained by substituting eq56 and eq57 in eq42:

j=j_0\left [ e^{\left ( 1-\alpha \right )f\eta}- e^{-\alpha f\eta}\right ]\; \; \; \; \; \; \; \; 58

Eq58 is known as the Butler-Volmer equation. If the experiment that we have described so far consists of an electrolyte that is well stirred (so that mass transfer effects are minimised, i.e. no concentration polarisation), the polarisation of the electrode arising from the flow of current is attributed entirely to the change in activation energies of the reduction and oxidation reactions at the electrode, i.e. activation polarisation. Since η is a measure of activation polarisation, η in eq58 is ηact,cat, the activation overpotential at the cathode. Note that ηact,cat < 0 , because η = EEeqm where E decreases at the cathode upon polarisation.

If M is more electropositive than hydrogen, it becomes the anode when connected to the SHE. The flow of electrons away from M results in a less negative electrode reduction potential. Using the same logic, we arrive at eq58 with η being ηact,an, the activation overpotential at the anode, where ηact,an > 0.

In general, for the electrochemical cell M_a\left | M_a^{\; +}\left | \right |M_c^{\; +} \right |M_c (see this article for notation),

E_{galvanic}=\Delta\phi_R-\Delta\phi_L=\left ( E_R+\eta_R \right )-\left ( E_L+\eta_L \right )=E_{cell,open}+\eta_R-\eta_L\; \; \; \;\; 59

where Ecell, open is the open-circuit potential of the cell at equilibrium, which is usually the standard electrode potential of the cell, E0. Since ηR < 0 and ηL > 0, we can rewrite eq59 as:

E_{galvanic}=E_{cell,open}-\Pi \; \; \; \; \; \; \; \; 60

where \Pi=\left | \eta_R \right |+\eta_L. Eq60 shows that the relationship of Egalvanic < Ecell,open is always true when a current flows in an electrochemical cell.

For an electrolytic cell, Ecell,open is usually negative. If we also consider IR drop, the magnitude of the externally applied potential must be greater than the sum of the magnitudes of the cell open-circuit potential, potential due to IR drop and the overpotential, for electrolysis to proceed:

E_{applied}=E_{cell,open}-Ir-\Pi\; \; \; \; \; \; \; \; 61

 

Next article: Concentration Polarisation
Previous article: Polarisation
Content page of advanced electrochemistry
Content page of Advanced chemistry
Main content page

Concentration polarisation (electrochemistry)

Concentration polarisation is the change in electrode potential from its equilibrium value when a certain level of current flows, due to inefficiencies in electrode reactions caused by sluggish mass transfer between the electrode surface and the solution.

When a piece of metal is immersed in its ionic solution, an equilibrium electrode potential is established with no net current flowing. If the metal is connected to another half-cell, e.g. an SHE, and the metal is below hydrogen in the electrochemical series, a net cathodic current flows.

M^{z+}+ze^- \begin{matrix} r_{red}\\\rightleftharpoons \\ r_{ox} \end{matrix}M

Furthermore, if the metal electrode is consuming species faster than reactants reaching the electrode surface, a concentration gradient arises between the bulk solution and the electrode surface. Since diffusion is a slow process, the depletion of reactants at the electrode surface leads to a lower electrode potential (similarly, when the electrode is producing species faster than the products leaving the electrode surface, the electrode again becomes polarised). The difference between the modified potential and the equilibrium potential is called the concentration overpotential ηcon.

To derive a mathematical relationship between current (or current density) and concentration overpotential, we have to assume that activation overpotential of the cell is negligible. This is achieved by using of non-polarisable electrodes.

At equilibrium, the potential of the cathode is given by the Nernst equation:

E=E^{\, o}+\frac{RT}{zF}ln[M^{z+}]\; \; \; \; \; \; \; \; 62

When a current flows and polarisation of the electrode occurs, [Mz+] becomes [Mz+’] where [Mz+’] < [Mz+]. The corresponding open-circuit electrode potential is

E'=E^{\, o}+\frac{RT}{zF}ln[M^{z+'}]\; \; \; \; \; \; \; \; 63

The concentration overpotential of the polarised electrode is

\eta_{con}=E'-E=\frac{RT}{zF}ln\frac{\left [ M^{z+'} \right ]}{\left [ M^{z+} \right ]}\; \; \; \; 64

According to Fick’s first law of diffusion, the flux of ions, r, towards the cathode is

r=-D\frac{\left [ M^{z+'} \right ]-\left [ M^{z+} \right ]}{x}\; \; \; \; \; \; \; \; 65

where D is the coefficient of diffusion with SI units m2s-1 and x is the distance between the bulk solution and the electrode surface.

 

Question

How is Fick’s first law derived?

Answer

In diffusion, particles move from a region of higher concentration to one of lower concentration. The flux of particles (i.e. the movement of the number of particles N per unit area per unit time) across a distance x is therefore proportional to the difference of [N]x=x and [N]x=0 divided by the change in distance, i.e.

r=D\frac{d[N]}{dx}

Knowing that d[N] < 0, if we defined r as positive in the direction of x, the above equation becomes

r=-D\frac{d[N]}{dx}

 

Substituting eq65 in eq39 of the previous article, the cathodic current density is:

j=zFD\frac{\left [ M^{z+} \right ]-\left [ M^{z+'} \right ]}{x}\; \; \; \; \; \; \; \; 66

Note that current in a cell not only flows from one electrode to the another electrode along a wire, but also continues to flow back to the first electrode through the electrolyte, and eq66 represents this flow. Rearranging eq66,

\left [ M^{z+'} \right ]=\left [ M^{z+} \right ]-\frac{jx}{zFD}\; \; \; \; \; \; \; \; 67

Substituting eq67 in eq64 and rearranging,

j=\frac{zFD\left [ M^{z+} \right ]}{x}\left ( 1-e^{zf\eta_{\, con}} \right )\; \; \; \; \; \; \; \; 68

where f = F/RT.

Concentration overpotential, like activation overpotential, leads to a lower potential than the equilibrium potential of a galvanic cell when a current flows.

E_{galvanic}=E_{cell,open}-\Pi

where \Pi=\left | \eta_{con,R} \right |+\eta_{con,L}. In the case of an electrolytic cell, a higher applied potential is needed to maintain the desired current. If we include IR drop, the formula is:

E_{applied}=E_{cell,open}-Ir-\Pi

 

Next article: Three-electrode cell
Previous article: Activation Polarisation
Content page of advanced electrochemistry
Content page of Advanced chemistry
Main content page

Three-electrode cell

A three-electrode cell is used for electrochemical measurements involving the passage of current. It is a good set-up to evaluate electrochemical quantities like overpotential and IR drop. A typical three-electrode system consists of a working electrode, a reference electrode and a counter (or auxillary) electrode (see diagram below).

Generally, the potential of a reference electrode must remain constant in order to measure the potential of a working electrode. However, electrodes become polarised when a current flows. The solution is to introduce a third electrode called a counter electrode to the cell for dynamic measurements. A high impedance voltmeter displays the potential of the working electrode and is connected between the working electrode and the reference electrode so that very negligible current flows between them. Experiments involving the polarisation of the working electrode at various current levels can therefore be conducted with respect to the counter electrode. Furthermore, the reference electrode is fitted with a movable column of electrolyte called the Luggin capillary, which allows the measurement of IR drop by varying the distance d.

It is important to note that the counter electrode serves only to complete the circuit for current flow. Ideally, it should not to interfere with reaction at the working electrode and is designed to be non-polarisable. One way to minimise the formation of overpotential at the counter electrode is to increase its surface area, A. This in turn reduces its current density (a fixed amount of current from the potentiostat is spread over a larger area), which is described by the Butler-Volmer equation:

j=\frac{I}{A}=j_0\left [ e^{(1-\alpha)f\eta}-e^{-\alpha f\eta} \right ]

Since j0 is an intensive property, \left [ e^{(1-\alpha)f\eta}-e^{-\alpha f\eta} \right ] is small if A is large. This implies that

e^{(1-\alpha)f\eta}\approx -e^{-\alpha f\eta}

(1-\alpha)f\eta\approx -\alpha f\eta\; \; \Rightarrow \; \; f\eta\approx 0

Since f ≠ 0, η ≈ 0.

 

Next article: Measuring overpotential
Previous article: Concentration polarisation
Content page of advanced electrochemistry
Content page of Advanced chemistry
Main content page

Measuring overpotential

The three-electrode cell is used to measure the activation overpotential of a cell at various levels of current. Consider the following experiment setup:

Fe^{3+}(aq)+e^-\rightleftharpoons Fe^{2+}(aq)\; \; \; \; \; E_{vs\: AgCl(sat'd)}=+0.57\: V

The opening end of the Luggin capillary is brought very close to the working electrode (d≈0) and the potential of the working electrode Eeqm is measured with the potentiometer and counter electrode removed from the circuit. The two are then reconnected to the circuit and the externally applied potential is turned on to a magnitude greater than 0.57 V (e.g. 0.8V), resulting in Fe2+ being oxidised to Fe3+ at the anode and H+ reduced to H2 at the cathode. The potential of the working electrode Epol is again recorded. Assuming that the electrolyte is well-stirred, the difference between Epol and Eeqm is the activation overpotential of the anode with respect to the current flowing at 0.8 V. By increasing the applied potential, we can tabulate the increasing levels of current, I, flowing in the circuit (see table below) and consequently record the corresponding overpotentials ηact,anode using eq61.

If the overpotential at the anode is greater than +100 mV, the second exponential in the Butler-Volmer equation tends to zero, giving:

j=\frac{I}{A}=j_0\, e^{(1-\alpha)f\eta}\; \; \; \; \; \; \; \; 69

This means that the net current density at the anode mainly comprises of anodic current density. Taking natural logarithm on both side of eq69,

ln\left ( \frac{I}{A} \right )=ln j_0+(1-\alpha)f\eta\; \; \; \; \; \; \; \; 70

Eq70 is known as the Tafel equation, named after the Swiss chemist, Julius Tafel. The plot of ln(I/A) against η is called a Tafel plot and it allows us to find the value of the transfer coefficient \alpha from the gradient of the line and the exchange current density j0 from the intercept. Such an experiment that monitors current at various potentials is called voltammetry.

The three-electrode cell can also be modified slightly to measure the total overpotential of a cell, \Pi. Consider the following experiment with metallic copper for both the anode and cathode, and AgCl electrode as the reference electrode.

With the potentiometer set to 0 V,

    • Measure potential of the left electrode EL, eqm by connecting the reference electrode to junction X.
    • Measure potential of the right electrode ER, eqm by connecting the reference electrode to junction Y.

We’d expect EL, eqm = ER, eqm. Next, turn the potentiometer to 1.0 V and

    • Measure potential of the left electrode (anode) Ean, pol by connecting the reference electrode to junction X.
    • Measure potential of the right electrode (cathode) Ecat, pol by connecting the reference electrode to junction Y.

We’d expect Ean, pol > EL, eqm and Ecat, pol < ER, eqm.

\eta_{an}=E_{an,pol}-E_{L,eqm}>0\; \; \; and\; \; \; \eta_{cat}=E_{cat,pol}-E_{R,eqm}<0

\Pi=\eta_{an}+\left | \eta_{cat} \right |

We can also measure IR drop (in mV/cm) for the above setup by setting the external applied voltage at 1.0 V and measuring the anode potential at different Luggin capillary distances, d.

Next article: Pourbaix diagram
Previous article: Three-electrode cell
Content page of advanced electrochemistry
Content page of Advanced chemistry
Main content page

Pourbaix diagram

A Pourbaix diagram is a reduction potential versus pH graph. It was invented by Marcel Pourbaix, a Belgian chemist, in the 1930s. A typical Pourbaix diagram illustrates the equilibria of electrochemical and non-electrochemical reactions. It consists of plots of equations derived either from the Nernst equation (for electrochemical reaction equilibria) or the equilibrium constant relation (for non-electrochemical reaction equilibria), with all reactions related to a particular metal or non-metal. For example, the equilibria of acidified water are:

E1 : O_2(g)+4H^+(aq)+4e^-\rightleftharpoons 2H_2O(l)\; \; \; \; \; \; \; E^{\, o}=+1.23\, V

E2 : 2H^+(aq)+2e^-\rightleftharpoons H_2(g)\; \; \; \; \; \; \; E^{\, o}=0.00\, V

with the following Nernst equations at rtp, a_{H_2O}=1 , f_{O_2}=1\: atm and f_{H_2}=1\: atm:

E_1=1.23-\frac{(8.31)(298)}{(4)(96485)}ln\frac{1}{\left ( a_{H^+} \right )^4}=1.23-0.0592pH\; \; \; \; \; \; \; 71

E_2=-0.0592pH\; \; \; \; \; \; \; \; 72

The Pourbaix diagram for acidified water is the E/ESHE-pH graph of eq71 and eq72 (where E/ESHE is the electrode potential of a reaction relative to the Standard Hydrogen Electrode):

A Pourbaix diagram is therefore a map of the relative stability of species in a reaction with respect to E/ESHE and pH. Acidified water is unstable above the upper line with regard to the evolution of oxygen gas. Below the lower line, acidified water is again unstable, as it is converted to hydrogen gas. In other words, acidified water is relatively stable in the region between the two lines. Understanding the electrochemical stability of water is important in many industrial processes, e.g. the refining of copper, which has the following Pourbaix diagram:

The equilibria involved are:

1) Cu^{2+}(aq)+2e^-\rightleftharpoons Cu(s)\; \; \; \; \; E^{\, o}=+0.337\, V

2) CuO(s)+2H^+(aq)\rightleftharpoons Cu^{2+}(aq)+H_2O(l)

3) 2Cu^{2+}(aq)+H_2O(l)+2e^-\rightleftharpoons Cu_2O(s)+2H^+(aq)\; \; \; \; \; E^{\, o}=+0.203\, V

4) Cu_2O(s)+2H^+(aq)+2e^-\rightleftharpoons 2Cu(s)+H_2O(l)\; \; \; \; \; E^{\, o}=+0.471\, V

5) CuO(s)+2H^+(aq)+2e^-\rightleftharpoons Cu_2O(s)+H_2O(l)\; \; \; \; \; E^{\, o}=+0.669\, V

6) HCuO_2^{\; -}(aq)+H^+(aq)\rightleftharpoons CuO(s)+H_2O(l)

7) 2HCuO_2^{\; -}(aq)+4H^+(aq)+2e^-\rightleftharpoons Cu_2O(s)+3H_2O(l)\; E^{\, o}=+1.783\, V

and their corresponding equations are:

a) E=0.337+0.0296loga_{Cu^{2+}}

b) K=\frac{a_{Cu^{2+}}}{a_{H^+}^{\; \; \; \; \; 2}}\; \; \Rightarrow \; \; logK=loga_{Cu^{2+}}+2pH\; \; \; \left ( K=10^{7.89} \right )

c) E=0.203+0.0591loga_{Cu^{2+}}+0.0591pH

d) E=0.471-0.0591pH

e) E=0.669-0.0591pH

f) K=\frac{1}{a_{H^+}a_{HCuO_2^{\; -}}}\; \; \Rightarrow \; \; logK=-loga_{HCuO_2^{\; -}}+pH\; \; \; \left ( K=10^{18.83} \right )

g) E=1.783+0.0591loga_{HCuO_2^{\; -}}-0.1182pH

Equations (a), (c), (d), (e) and (g) are derived using the Nernst equation, while equations (b) and (f) originated from the equilibrium constant relation. We assume all activities of aqueous species are equivalent to their concentrations and that the diagram is plotted with all aqueous copper species concentrations of 10-6 M.

With reference to the copper Pourbaix diagram, the horizontal line corresponds to eq(a), which is independent of pH (i.e. E doesn’t change by varying pH). The two vertical lines correspond to eq(b) and eq(f), both of which are independent of E (i.e. pH doesn’t change by varying E). The slanted lines correspond to eq(c), eq(d), eq(e) and eq(g). Furthermore, the Pourbaix diagram for the electrolysis of acidified water is superimposed on the diagram for copper, as the electrolyte consists of water.

The Pourbaix diagram allows us to select the appropriate E and pH for copper refining, which is represented by equilibrium(1). The preferred pH is between 2 and 4 (for [Cu2+] = 10-6 M), to avoid the formation of oxides. Furthermore, the more negative the cathodic potential is (by application of an external potential:  see diagram below), the greater equilibrium(1) shifts to the right. However, if the potential is too negative, the deposition of copper has to compete with the evolution of hydrogen gas at the cathode, which reduces the yield. Hence, the thermodynamically preferred E value is roughly between 0.00 V and 0.16 V for 2 < pH < 4 and [Cu2+] = 10-6 M. In practice, the preferred E value is also influenced by kinetics, as a result of overpotential, IR drop, etc.

Finally, the copper Pourbaix diagram is also useful for corrosion prevention. According to the diagram below, a layer of oxide is formed to prevent further corrosion of Cu (passivation) when the solution has a pH of between 8 to 12. In acidic solutions, the corrosion of Cu is prevented by bringing the potential of the metal to less than 0.1 V.

 

Previous article: Measuring overpotential
Content page of advanced electrochemistry
Content page of Advanced chemistry
Main content page
Mono Quiz