Vous êtes sur la page 1sur 6

Thermoelectric Power by the Diffusion of Protons in a Nanoporous Structure Michael A.

Reznikov Physical Optics Corporation, 20600 Gramercy Place, Torrance, CA 90501, U.S.A. ABSTRACT The work presents the model for the ionic thermoelectric phenomenons, which is based on the consideration of Grotthuss, hopping mechanism for the proton conductivity. The Seebeck coefficient and figure of merit are estimated in good agreement with experimental data. INTRODUCTION Most solid-state thermoelectric (TE) generators, based on the Seebeck effect, suffer from power loss due to the parasitic flow of heat. While low-dimensional structures, such as quantum wells, superlattices, quantum wires, and quantum dots, offer new ways to decrease the thermal (phonon) conductivity while improving the electric conductivity [1,2], their energy conversion efficiency is relatively low and fabrication is very expensive. Therefore, there is a need for new TE materials for waste heat recovery and electricity generation that have a high figure of merit (ZT) and the potential for large-scale production at costs competitive with conventional technologies, considering the entire system over its lifetime. It is known that the receptors of marine sharks and skates (rays), which contain the extracellular hydrogel, are the most exquisite sensors of thermal fluctuations [3-5]. We demonstrated that these ionic structures can be mimicked by natural and synthetic gels [6], and exhibit very high TE voltage, a few mV/K at low, room temperatures. These materials currently are commercially available due to progress in fuel cell technology, where they are used as proton conductive membranes. THEORY Theoretical models for thermoelectricity [7, 8] based on electronic structure are presumably inappropriate for proton-conductive substances. First of all, the density of mobile ions depends much less on the temperature than the density of free electrons. Second, the mobility of ions is very sensitive to temperature, due to the thermodiffusion mechanism of charge transfer [9]. Dynamic capturing and release of protons is the base mechanism of protonic transfer in nature. A mechanism for proton mobility was suggested to be comprised of the cyclic isomerization between the two forms of protonated water, H3O+ and H5O2+, which is coupled to hydrogen-bond dynamics in the second solvation shell of the H3O+. The formation of a high fraction of pore bulk water in proton exchange membranes (PEMs) is desirable for high conductivity because of the dominance of the Grotthuss diffusion mechanism in conductivity, which occurs in bulk water rather than at the surface. The Grotthuss mechanism is the protonhopping-mechanism where each oxygen atom simultaneously passes and receives a single hydrogen atom. . Due to interaction with the wall of a pore, localized water molecules create deeper traps for protons than those in the bulk water. This is illustrated in figure 1 (courtesy of [10]).

(a) (b) Figure 1. A simplified picture of structure and proton transfer in Nafion (a) in fully hydrated state, and (b) electrical analog of the proton transport in Nafion [10]. Proton transfer in hydrogel structures accounts for the hopping conductivity in the disordered structure with the energy barrier, Wb, between local states separated by distance . In the steady state, if there are no gradients of temperature, T, proton concentration, n, or electric potential, , there is no net exchange flow between the two states, 1 and 2: (j = j12 j21 = 0), as shown in figure 2(a).

(a) (b) Figure 2. Hopping proton exchange between two neighboring states: (a) no electric field; (b) electric field applied. Therefore, the average exchange flow (per unit of volume), jij, can be described as

j12

j21 Const n f T exp

Wb , kT

(1)

where f(T) is the frequency of oscillations of ion (attempts to go over the barrier). Gradient of electric potential, E = - d /dx, leads to the asymmetry of the energy barrier (see figure 2(b)) and the net exchange flow is

j12

j21 Const f n exp

Wb kT

sinh

q d 2kT dx

(2)

The condition of continuity for current at the steady state at the uniform gradient of temperature (dT/dx = ) leads to the thermoelectric generation expression. If the DC current is blocked (open circuit), the material is polarized until j = 0, and the alternative way is the consideration of microscopic equilibrium (j = j1,2 j2,1 = 0) and extrapolation of local deviations ( T and n) to the macroscopic range, which leads to the direct expression
n2 n1 Wd exp q E 2 kT 1 Wd q E 2 k T2 2Wd q E dT 2k T T dx q E kT

exp

exp

(3)

where multiplier exp temperature gradient

q E kT

can be neglected. This provides the solution for proton density at

n n0 exp

2Wd q E 1 2k T0 x

1 T0

(4)

where n0 and T0 are the proton density and temperature at the cold end, correspondently, and Wd = Wdis+Wb includes the enthalpy of dissociation, Wdis, to account for proton generation at varied temperature. Now the gradient of electric field can be expressed by the charge density, , due to 2Wd q E dE q n n0 the Poisson equation , and supposing that 1 we obtain dx 2 2kT 0 0 the equation for the electric field

q E ln 1 2Wd

q 2 n0 T0 x ln 1 4 0 kT0 T0

(5)

where the logarithm at the left and right sides can be replaced by a series up to the linear and quadratic members, correspondently. Therefore, the final expression for the gradient of electric potential is

d dx

Ex

qWd n0 2 x 2 4 0 kT0

(6)

and the thermoelectric voltage at the small distance, d, is


Ud qWd n0 T 2 d . 12 0 kT T

(7)

Here, T is the average temperature on the distance d, T Th Tc 2 , and T = Th-Tc. The distance d is small enough to satisfy our suppositions of weak electric field. Therefore, equation 9 can be considered as a linear approximation of the thermoelectric voltage per the unit of length in the temperature gradient. Equation 7 can be numerically evaluated without necessity to simplify both sides of the expression. This derives function E(x) and therefore the integral of it, U(x). Figure 3 presents the result of such numerical simulation for distilled water (initial concentration of protons, n0 = 6 107 cm-3) at temperature gradient 2 K/cm, hopping activation energy 0.26 eV (10 kT at the temperature 300 K), and average spacing between hopping places 1 nm.

Figure 3. Numerically simulated Electric Field and Thermoelectric Voltage in the nanoionic material at gradient of temperature 2 K/cm. Voltage quadratically increases with the distance (according to equation 9), and achieves a value close to 68 mV at distance 1 cm. The electric field doesnt exceed 20 mV/cm and therefore the hopping conductivity (see figure 2(b)) can be ignored. DISCUSSION The result of estimation of the thermoelectric voltage with an open circuit (no load) at distance 0.5 cm, as shown in figure 3, delivers the Seebeck coefficient, S, ~ 3 mV/K, which is in good agreement with the results of our tests for a Nafion thermal sensor with distance between the electrodes of 5 mm [6], which demonstrated thermoelectric voltage 2.36 mV/K. A nanoporous thermoionic (NPTI) element can be considered as an electrochemical cell that converts the thermal energy into electrical energy. The gradient of electrochemical potential

is established due to the separation of ions, hydronium H3O+ and hydroxyl OH-, in the selective ionic conductor over the temperature gradient. These ions are produced from a water molecule, H2O, with a temperature-dependent constant of dissociation, K K 0 exp Wdis kT , and following the association of a free proton, H+, with the water molecule H++H2O H3O+. The thermoactivated separation of ions creates an excess of cations, H3O+, at the cathode and anions, OH-, at the anode. Because the thermal energy is consumed mostly for the dissociation of water molecules, the generation of a single electron in the external circuit requires energy that is equal to the change of electrochemical potential in the electrolyte or Gibbs free energy, G, per single molecule of water, which is G = H - T S. Since entropy increases in the process of dissociation, it also consumes energy from the heat source at temperature T. The enthalpy change, H, for water dissociation is 285.83 kJ per one mole of water or 4.75 10-19 J per single molecule. Entropy change, S, is 69.91 J/K or T S = 48.7 kJ at 298K, which is 5.6 10-19 J per molecule. This energy is moved from the anode to the cathode in latent form (as a dissociated water molecule). Therefore, the thermoelectric current of 1 A (or 6.25 1018 electrons per second) at 298 K requires at least 3.5 W of heat flux. The efficiency of conversion for such a consideration depends on the voltage at the terminals of the NPTI cell, and the temperature. Figure 3 shows the voltage value of 68 mV at distance 1 cm and a difference of temperature 2 K. Therefore, the Seebeck coefficient, S, is 34 mV/K, which allows estimating the efficiency of the NPTI device by using the dimensionless figure of merit, ZT. Here, Z = S2/ where and are correspondently the electric and thermal conductivities, and T is the average temperature, (TH-TC)/2 between hot (TH) and cold (TC) ends. Assuming a polymeric or ceramic host, we use electric conductivity for the wet Nafion , 0.1 S/cm and thermal conductivity, 5.5 10-3 W/(K cm), of the water because the thermal conductivity of Nafion is even less, 1.3 10-3 W/(K m). Using the computed Seebeck coefficient, S=34 mV/K and temperature, T=300 K, the corresponding ZT = 6.3. The maximal achievable efficiency of a thermoelectric generator is defined as
max

T H TC TH

1 ZT 1 ZT

1 , TC TH

(8)

where the first term is the maximal Carnot efficiency. For TH = 90C = 363K and TC = 27C = 300K, max 8.7% or ~ 50% of the Carnot limit, 17.4%. This proportion depends on the TH, as one can see from equation 8. CONCLUSIONS The described model of thermoelectric effects in porous ionic conductors accounts the Grotthuss, hopping mechanism for the proton conductivity. The good agreement with experimental data is demonstrated. The relatively high ZT and correspondently max for the nanoionic device result from the low thermal conductivity of the host material and the relatively high electric conductivity of Nafion in combination with the high Seebeck coefficient.

The presented preliminary model should be expanded for elevated concentrations of protons and higher temperature gradients, when the electric field factor increases and must be accounted for. REFERENCES 1. M.S. Dresselhaus, Y.M. Lin, S.B. Cronin, M.R. Black, O. Rabin, G. Dresselhaus, Investigation of Low-dimensional Thermoelectric, Nonlithographic and Lithographic Methods for Nanofabrication: Symposium Proceedings, edited by Steven Smith, Technomic Publishing Co., Inc., Lancaster, PA, 2001. 2. S.B. Cronin, Y.-M. Lin, O. Rabin, M.R. Black. M.S. Dresselhaus, Thermoelectric Transport Properties of Bismuth Nanowires, The 21st International Conference on Thermoelectrics: ICT Symposium Proceedings, Long Beach, CA, edited by T. Caillat and J. Snyder, IEEE, Piscataway, NJ, pp. 243-248, 2002. 3. J.B.R. Brown, J.C. Hutchison, M.E. Hudges, D.R. Kellog, R.W. Murray, Electrical Characterization of Gel Collected from Shark Electrosensors, Phys. Rev. E, vol. 65, 061903, 2002. 4. B.R. Brown, Sensing Temperature without ion Channels, Nature, vol. 421, p. 495, 2003. 5. B.R. Brown, M.E. Hughes, C. Russo, Thermoelectricity in Natural and Synthetic Hydrogels, Phys. Rev. E, 70, 031917, 2004. 6. M. Reznikov, A. Kolessov, Space Charge Effects in the Electrolytes, Proc. of Electrostatics Joint Conference, ESA/IEA/IEJ/IAS/SFE, Boston, June 2009. 7. CRC Handbook of Thermoelectrics/ edited by D.M. Rowe, CRC Press, 1995. 8. G.S. Nolas, J. Sharp, and H.J. Goldsmid, Thermoelectrics. Basic Principles and New Materials Developments, Springer, 2001. 9. J. Janek and C. Korte, Thermal Diffusion in Mixed Conductors, Proc. Electrochem. Soc., vol. 97, no. 24, pp. 304-328, 1998. 10. P. Choi, N. H. Jalani, and R. Datta, Thermodynamics and Proton Transport in Nafion II. Proton Diffusion Mechanisms and Conductivity, Journal of the Electrochemical Society, vol. 152 (3), E123-E130, 2005.

Vous aimerez peut-être aussi