Electrolysis of supercritical aqueous solutions at temperatures up to 800 K and pressures up to 400 MPa

https://doi.org/10.1016/S0021-9614(02)00236-7Get rights and content

Abstract

A study on the electrolysis of sub- and supercritical aqueous solutions of sodium hydroxide is presented here. Current density–potential curves are recorded at temperatures up to about 800 K and pressures up to 400 MPa. The molar concentration of the solutions, as referred to T=298 K, is varied from c=(0.01to1)mol·dm−3. The autoclave is described in detail. Several electrode materials are examined. The highest reproducibility is obtained with gold electrodes. Some results are also given for platinum, silver and nickel electrodes. Below T=473 K the current–potential curves show the familiar transition from a range of very low currents to an ohmic behaviour with steeply increasing currents above the decomposition potential. The decomposition potential decreases with increasing temperature. Above T=473 K the current–potential curves no longer reflect distinct transitions from low-current to high-current behaviour, and at T=773 K, almost linear current–potential curves are observed over the entire voltage range. Up to current densities of 50mA·cm−2, the pressure dependence of the current density at a given potential is marginal. At higher current densities, up to 35A·cm−2, a substantial pressure effect is observed.

Introduction

Dense and supercritical fluids have properties which differentiate them from the solid, liquid and gaseous states of matter and constitute almost a fourth state of aggregation. Thermodynamic and kinetic properties can be varied continuously over wide ranges from liquid-like to gas-like states. Transport coefficients are high, even at liquid-like densities. And, particularly interesting, there are many binary and multi-component systems which show complete miscibility of their highly polar with nonpolar constituents. Many such examples are binary aqueous systems. They are characterised by “critical curves”, which begin at the critical point of pure water at T=647.3 K and 22.1 MPa, and extend uninterrupted or interrupted to the critical point of the nonpolar component [1], [2].

A few experimental examples of binary aqueous systems are shown in figure 1 which depicts critical lines of aqueous solutions of carbon dioxide [3], [4], hydrogen [5], oxygen [6], and ethane [7]. To the left of these critical curves, the low-temperature side, is the two-phase region. To the right, at higher temperatures, exists the homogeneous one-phase supercritical area. The figure also shows the initial part of the critical curve of sodium chloride which, from the critical point of water, extends to higher temperatures and pressures [8], [9]. One may expect that the critical curve of sodium hydroxide shows a similar behaviour. Some knowledge of these critical curves is necessary in order to operate the electrolysis cells discussed below.

We study here the electrolysis of dilute aqueous solutions of NaOH at temperatures up to about 800 K and at high enough pressures to achieve liquid-like densities. Knowledge of the electrolytic conductance of such systems is an important prerequisite for the design of electrolysis cells. The isothermal pressure dependence of the electrolytic conductance has been studied in several instances, see for example the early work on the conductance of dilute solutions of KCl, HCl and KOH up to T=1023 K and nearly 300 MPa [10], [11], [12]. At room temperature and a density near 1g·cm−3, a pressure increase to 400 MPa causes a decrease of the molar conductance by about 8%. At these conditions the dielectric constant of water is high and the salts are highly dissociated. The conductance behaviour then reflects mobility effects, in accordance with the decrease of the diffusion coefficient (D) [13], [14] and the increase of the viscosity (η) of water [15], as approximately reflected by the Stokes-Einstein law [16]D=kT6πηr,where k stands for the Boltzmann constant and r for the hydrodynamic radius of the considered ion. For NaOH the extra mobility contribution of OH should be considered [17].

In contrast, at higher temperatures the pressure dependence of the molar conductance of aqueous solutions is generally positive. At high temperatures and low pressures the dielectric constant of water is low, and the salts are only partially dissociated. An increase in density causes an increase in the dielectric constant [18], [19] and hence, higher ionisation of the solute. For example, at T=673 K a pressure of 400 MPa leads to a water density of about 0.9g·cm−3 [20] and a dielectric constant of about 24 [18], [19]. At 100 MPa and T=673 K the density is only 0.69g·cm−3 and the dielectric constant is 12.5, which is, however, still sufficient for partial ionisation of salts with univalent ions.

The discussion of the performance of electrolysis in wide ranges of temperature and pressure requires the determination of current–potential curves and of decomposition potentials. Already in 1904, Th. Wulf reported on the pressure dependence of the decomposition potential for H2 and O2 of aqueous solutions with platinum electrodes up to T=338 K and 100 MPa [21]. He found that the decomposition potential varied only slightly and within the uncertainty ranges. In 1928 Tamman and Jenckel (1928) investigated the effect of pressure on current-voltage curves upto 300 MPa at room temperature [22]. They found that at constant potential the pressure caused the electric current density to grow. Lange showed in 1933 that at higher, but subcritical temperatures, higher pressures decreased the overpotential at an electrode [23]. The application of high pressures in these measurements required, however, comparatively low temperatures, because of the limited mechanical strength of the autoclave materials. Modern materials and techniques make possible the simultaneous application of high pressures and high supercritical temperatures.

Since the 1980s, there has been considerable interest in electrochemistry in supercritical fluids. However, this interest has been mainly limited to solvents such as CO2 with critical points at comparatively low temperatures and pressures [24]. A few voltametric measurements for supercritical aqueous solutions have been reported by Bard and coworkers [25], [26]. However, in these studies only moderate pressures, typically of about 30 MPa, were applied. No attempts were made to investigate the pressure dependence of the voltametric properties in detail, or to generate supercritical states of liquid-like density.

Besides an interest in the scientific character of current–potential curves and the contributions of the overpotential, there is a possible practical interest. Decomposition under pressure can produce pressurised hydrogen and oxygen gases directly without later compression. High-pressure electrolysis and fuel cells could perhaps be combined to achieve energy storage. At a later stage, electrolysis of binary water-salt systems could be extended to homogeneous ternary systems with hydrocarbons or other organic compounds. Hydrogenation and other chemical processes might be carried out using a supercritical electrolytic aqueous phase cell. In supercritical water oxidation processes [27] the generation of oxygen could be achieved by electrolysis [28].

Section snippets

Apparatus

All experiments were performed in a cylindrical autoclave consisting of a nickel-base, corrosion-resistant and high-strength alloy, designed for a maximum temperature of 873 K and a maximum pressure of 400 MPa. The solution was compressed in a first step by a rotating pump. The precompressed fluid mixture was subsequently brought to the desired final high pressure with a hand-operated spindle press. The pressure was measured with a precision of ±1 MPa by a pressure-calibrated Bourdon gauge. The

Gold electrodes at low current densities up to 50 mA·cm−2

Because gold electrodes showed the highest reproducibility, electrolysis behaviour is mainly illustrated here by experiments with gold electrodes. Some data given later for other electrode materials more or less confirm these results.

In a first series of experiments, current–potential curves were recorded with gold electrodes at current densities up to 50mA·cm−2. The concentrations of the NaOH solutions at T=298.15 K and at atmospheric pressure were c=(0.01,0.1and1)mol·dm−3, respectively. The

Discussion

At low temperatures all current–potential curves show a familiar behaviour with an indication of a decomposition potential which lies above the equilibrium decomposition potential, thus including an overpotential. At low voltages the products H2 and O2 at the electrodes give rise to a potential which practically compensates for the applied voltage. There is only a residual current due to the continuous diffusion of the gases away from the electrode surfaces. As long as the gas pressure is lower

Conclusions

In supercritical aqueous solutions electrolysis is very different from that near ambient conditions. Overpotential is not indicated in the isotherms of the current density against potential. Rather, these show an ohmic behaviour. The gaseous decomposition products, H2 and O2 are completely miscible with supercritical water, the high temperatures reduce the influence of activation barriers and the transport processes are largely enhanced due to high diffusivities even at liquid-like densities.

References (34)

  • W.L. Marshall et al.

    J. Inorg. Nucl. Chem.

    (1974)
  • B. Misch et al.

    Supercrit. Fluids

    (2000)
  • E.U. Franck

    Angew. Chemie

    (1961)
  • G.M. Schneider

    Ber. Bunsenges. Phys. Chem.

    (1966)
  • A.E. Mather et al.

    J. Phys. Chem.

    (1992)
  • K. Tödheide et al.

    Z. Phys. Chem. Neue Folge

    (1963)
  • T.M. Seward et al.

    Ber. Bunsenges. Phys. Chem.

    (1981)
  • M.L. Japas et al.

    Ber. Bunsenges. Phys. Chem.

    (1985)
  • A. Danneil et al.

    Chem. Ing. Tech.

    (1967)
  • J.L. Bischoff et al.

    Am. J. Sci.

    (1989)
  • E.U. Franck

    Z. Phys. Chem., N. F.

    (1956)
  • E.U. Franck

    Z. Phys. Chem., N. F.

    (1956)
  • E.U. Franck

    Z. Phys. Chem., N. F.

    (1956)
  • W. Lamb et al.

    J. Chem. Phys.

    (1981)
  • T. Buttenhoff et al.

    J. Phys. Chem.

    (1996)
  • K.H. Dudziak et al.

    Ber. Bunsenges. Phys. Chem.

    (1966)
  • E.U. Franck et al.

    Ber. Bunsenges. Phys. Chem.

    (1958)
  • Cited by (9)

    • A high temperature and pressure framework for supercritical water electrolysis

      2022, International Journal of Hydrogen Energy
      Citation Excerpt :

      The availability of oxygen gas at the cathode in this configuration is then likely to be the cause of the reduced temperature at which this phenomenon is observed, and at higher overpotentials. The presence of oxygen gas in solution, being reduced at the cathode, may also then be the cause of the shifting reversible potentials observed here and in the literature [18,35]. Fig. 6 highlights the supercritical region of the Arrhenius plots, with an observation made at low potentials.

    • Pressure intensified HO[rad] evolution from OER and electrolysis desulfurization

      2019, Electrochimica Acta
      Citation Excerpt :

      Water electrolysis under high temperature and pressure conditions (such as reaching a supercritical state) could greatly increase the water dissociation constant (Kw) and improve the degree of water dissociation into OH− [31]. And the non-polar gas (e.g. oxygen and hydrogen) under the supercritical state can be highly miscible with water [32] which would result in a very high coefficient of diffusion [33]. The high diffusion coefficient would enhance the ionic strength of electrolyte and meanwhile reduce the mass transfer resistance of electrolytic reactants during the electrolysis process.

    • Intensified hydrogen production and desulfurization at elevated temperature and pressure during coal electrolysis

      2018, Electrochimica Acta
      Citation Excerpt :

      With increasing temperature and pressure (especially, to achieve zero boundary state), the water dissociation constant (Kw) would increase obviously and produce more OH− and H+. Under this circumstance, water could be highly miscible with nonpolar gases such as hydrogen and oxygen [22], and the reactants exhibited very high diffusion coefficients [23]. These advantages applied in the electrolysis cell would certainly increase the ionic strength of the electrolyte and reduce the mass transfer resistance of surface chemical reactions.

    • High temperature and pressure alkaline electrolysis

      2009, International Journal of Hydrogen Energy
      Citation Excerpt :

      It is important to note that this observation of a greatly reduced applied potential in a high-temperature and pressure electrolysis system is not unprecedented. In a previous study by Boll and coworkers [25], the electrolysis of truly supercritical NaOH solutions, with temperatures of greater than 500 °C and pressures up to 400 MPa, was carried out using gold electrodes in a modified autoclave. As an example, the authors reported that the required electrical potential to provide a current density of 40 mA cm−2 was only 500 mV at a temperature of 530 °C and a pressure of 200 MPa.

    View all citing articles on Scopus
    View full text