You are on page 1of 8
Available online at www sciencedirect com Energy cost ScienceDirect Proced ia ELSEVIER Energy Procedia 160 (2019) $42-S49 won elsevier.comlocate/procedia 2nd International Conference on Energy and Power, ICEP2018, 13-15 December 2018, Sydney, Australia Open-cathode PEMFC heat utilisation to enhance hydrogen supply rate of metal hydride canisters Reza Omrani*, Huy Quoc Nguyen, Bahman Shabani School of Engieermg, RMIT University, Melbourne, Australia Abstract In this paper, the hydrogen supply to an open-cathode PEM fuel cell (FC) by using metal hydride (MH) storage and thermal coupling between these two components are investigated theoretically. One of the challenges in using MH hydrogen storage ‘canisters is their limited hydrogen supply rate as the hydrogen release from MH is an endothermic reaction. Therefore, in order to ‘meet the required hydrogen supply rate, high amounts of MH should be employed that usually suggests storage of hydrogen to be higher than necessary for the application, adding tothe size, weight and cost ofthe system. On the other hand, the exhaust heat (i.e that is usually wasted if not utilised for this purpose) from open-cathode FCs is a low-grade heat. However, this heat can be transferred to MH canisters through convection to heat them up and increase their hydrogen release rate. A mathematical model is ‘used to simulate the heat transfer between PEMEC exhaust heat and MH storage, This enables the prediction of the required MH for different FC powwer levels with and without heat supply to the MH storage. A 2.5-kW open-cathode FC is used to measure the ‘exhaust air temperature at different output powers. It was found that in the absence of heat supply from the FC to the MH canisters, significantly higher number of MH canisters are required to achieve the required rate of hydrogen supply to the FC for sustained ‘operation (specially at high power outputs). However, using the exhaust hot air from the FC to supply heat to the MH storage, can reduce the number of the MH canisters required by around 40% to 70% for power output levels ranging from 500 W to 2000 W © 2019 The Authors, Published by Elsevier Ltd This isan open access article under the CC BY-NC-ND license (tips /creativecommons on icenses by-ne-nd'4. 09) Selection and peer-review under responsibility of the scientific committee of the 2nd Intemational Conference on Energy and Power, ICEP2018, Keywords: PEMEC: metal hydride: heat recovery; hydrogen release rate enhancement; thermal management * Corresponding author. Tel.: +61 3 9925 4353. E-mail address: rera.ccnrani@mit.ta a0 1876-6102 © 2019 The Authors, Published by Elsevier Ltd This is an open access article under the CC BY-NC-ND license (tps /eveativecommons.org/licenses/by-ne-nd/4.0/) ‘Selection and peet-review under responsibilty of the scientific committee ofthe 2nd International Conference on Energy anxd Power, ICEP2O18, 10.1016/}.eaypro.2019.02.204 Reza Omran eral. (Energy Procedia 160 (2019) $42-549, ss 1. Introduction Proton exchange membrane (PEM) fuel cells (FC) have been widely used as one of the most promising solutions for sustainable energy supply for both stationary [1] and mobile (2-4] applications. This growing interest is mainly ‘due to their low to zero emissions, high energy conversion efficiency, fast start-up, and quick response to load change [5-8]. Duc to different losses in PEMFCs, a considerable amount of heat is generated (around 45-60% of reacted hydrogen energy). The gencrated heat must be effectively removed from the stack to maintain the stack at optimal ‘operating temperatures and avoid FC degradation and performance drop. In order to improve the FC system overall energy efficiency, this waste heat is usually recovered for small-scale low-temperature heating applications (e.g. space heating, domestic hot water and inlet reactants preheating) [8-11]. ‘Metal hydride (MH) has been considered as an attractive solution for hydrogen storage due to its high volumetric density (ie. 100 g/L), stability, safety and, reliability [12]. In fact, MH storage systems have been integrated with PEMFC-powered systems for mobile and stationary applications [11, 13]. Hydtogen discharging from MH is an endothermic reaction, Therefore, additional heat is needed be supplied to maintain the sufficient hydrogen release rate. However, due to low thermal conductivity of the MH materials (lower than 1 Wm-IK-1) [14], matching hydrogen release rate from MH tanks at ambient temperature (i.e. 20-30 °C) with FC stack hydrogen demand is often a technical challenge [15]. Oversizing the MH system can partly address this challenge; however, it increases the overall system ‘weight, size, and cost that are not desirable. Considering the generated heat by FCs that is often wasted due to its relatively low temperature (ie, heat recovery may not be economical depending on the applications), heat recovery from FC to enhance MH storage hydrogen release rate is an attractive solution, Several published researches theoretically and experimentally investigated the possibility of capturing FC waste heat to heat up the MH storage. By testing thermal coupling of a 1.2-kW PEMFC and 17.5 kg of ABS MH storage, Forde et al. [16] found that MH canisters can supply sufficient hydrogen for the operation of FC stack at maximum power for 3 hours. They have reported that only about 25% of total cooling load of the fuel cell was sufficient to ‘maintain the MH storage temperature at 30°C and achieve the desired hydrogen flow rate. Similar study by Wilson et al. [17] showed that MH storage hydrogen release rate increases from 4 slpm to 8 slpm at 2 bar of pressure by capturing ‘waste heat from a 1-kW PEMEFC to heat up MH canisters. The possibility of hydrogen discharge rate enhancement of MH storage by thermal coupling with high-temperature PEMFC is also investigated [18-20]. Most ofthese researches focuses on liquid-cooled PEMFC which requires redesigning of heat exchangers (i.e. U-tube, spiral tubes and external heat jacket) for MH canisters. Recently, efforts have been made to reduce the mass, volume, complexity and parasitic energy consumption by introducing passive thermal coupling PEMFC and MH hydrogen storage using heat pipe. Tetuko et al. [15, 21, 22] proposed novel model of passive PEMFC heat recovery utilising heat pipes to enhance the MH canisters hydrogen release rate. In this model, heat pipes provide a thermal bridge between FC cooling plates and MH canisters; Their results showed that no more than 20-30% of cooling load is required to achieve the required hydrogen flow rate for ‘operating the FC. They have also reported that the addition of heat pipes is still needed to remove the remaining cooling load (70-80%) to maintain the FC at the desired operating temperature levels (i.e. 60-80 °C) [15, 21, 22]. However, there are few paper up to date discussed in detail about the possibility of thermal-coupling between ‘commercial PEMFC and MH canisters. Therefore, in this paper a mathematical model of an open-cathode PEMFC heat recovery for enhancement hydrogen discharge rate of MH hydride canisters was developed to calculate the required hydrogen supply from the MH canisters and hence the number of the MH canisters required at different FC power output levels as well as improving the MH storage supply rate by thermal coupling of an open-cathode PEMFC and MH storage through forced convection to reduce the required MH canisters, 2, Mathematical model 2.1. Model configuration A theoretical model of open-eathode PEMFC heat recovery for enhancement of MH canisters hydrogen release rate has been developed. In this configuration (Fig. 1), the MH canisters are considered to be placed behind the fuel cell in a way that the exhaust air passes over the MH canisters to recover the cooling heat from the FC stack. For this, su Reza Omran eral. (Energy Procedia 160 (2019) $42-549, study, a 2.5 kW open-cathode fuel cell stack (66 cells) from Horizon Fuel Cell Technologies is considered. The MH canisters considered here are the 800-s! model used by Tetuko ct al, [21, 23]. The 2.5 kW fuel cell was operated at different output powers in order to measure the exhaust air temperature. H2 inlet MBH canisters ome Hot Exhaust ait ( » > cu) re OO air — t q - a I stack oO H2 outlet Fig. 1. Schematic ofthe thermal coupling (foreed convection) between an open-cathode PEMFC and MH canisters, 2.2. Mathematical mode! of PEMFC ‘The waste heat from the FC stack can be calculated in order to be compared with the MH canisters heat demand. The factors determining the heat generation are the FC power output (Prc) and cell voltage (Vi). Heat generation is calculated by comparing the actual voltage with maximum output of a 100% efficiency fuel cell (ie. 1.48 V and 1.25 V based on HHV and LHV of hydrogen respectively). In this paper, itis assumed that the product water leaves the fuel cell is in vapor form, so the waste heat from the stack is given by Eq. 1 [24]. 1.25 (j -1fP a ‘ell In this model, the hydrogen flow rate from MH canisters is needed to be matched to the hydrogen demand by FC stack. The hydrogen flow rate required by FC can be calculated by Eq. 2 [8]. ag wt " 1000F @ where Jin is the hydrogen stoichiometry, which is equal to one when the fuel cell is operated in dead-ended anode ‘mode; 1 is the fuel utilisation coefficient and the value considered here is 95%: m is the number of the cells in the FC stack (66 for the FC stack considered here), /is the current (A); and F is the Faraday constant. 2.3. Mathematical model of MH hydrogen storage system ‘The hydrogen desorption in MH is an endothermic process, and hence their hydrogen release rate is dependent on the temperature and the supplied heat. MHs hydrogen absorption and desorption (the focus of this paper) can be enhanced by appropriate thermal management. The MH canister hydrogen release rate depends mainly on the ‘equilibrium pressure and temperature as shown in the Eq. 3 (15, 16] —*(p,-pp) @ Reza Onan etal (Energy Procedsa 160 (2019) 342-549 sas where Cis the desorption constant (s"!); E, is the desorption activation energy (J mol"); T is the MHs temperature (K); Ris the universal gas constant (8.314 J mol” K"); peyis the desorption equilibrium pressure (Pa); (kg m°) is the metal hydride desorbed density; and , (kg m*) is the metal hydride alloy density with no hydrogen (i.e. empty storage) As discussed in previous section, to maintain the required hydrogen rate for operating the FC at desired power, MH canisters absorb heat, The heat demand by the MH (Qxz) includes heat required for hydrogen desorption reactions (Qu) and heat carried out by released hydrogen gas (1) [15] rar = Qraace +O, m,(h+h,) @ ‘where i is the reaction enthalpy change (I kg"); and hy isthe discharged hydrogen gas enthalpy (J kg"). In this study, the parameters and the values reported by Tetuko et al. [15, 21, 23] are used, 2.4. Heat transfer from fuel cell exhaust to MH canisters ‘The temperature of the open-cathode PEMFC stack outlet air is moderate (~ 30-45 °C) and the mechanism of heat transfer from hot air to MH canisters is mainly convection, Therefore, the heat transfer from air to MHL canisters surface is calculated by Newton’s law of cooling. rons = PransArae Trae ~T) 6) where Auer is the effective area of MH (mm); Tags and Tar is the temperature of the MH and air stream (°C); hi i convection heat transfer coefficient (Wm? K*!, For natural convection (ie. no heat supply to the MH canisters from the FC), the correlation developed by Churchill and Chu [25] is used, is GryPr Nuj,* =0.60 +0.387| (6) hi +(0.559/Pr""] where Gro is Grashof number; Pr is Prandtl number; D is diameter of MH canister (m). In case of heat supply from the exhaust air of the fuel cell to the MH canisters, for flow actoss the cylinders like MH canisters, the convection coefficient is calculated by the correlation developed by Gnielinski (26). Nu a 0.34 (N02 joy +N ays) o where 4 = 1.D/2; kris the air thermal conductivity (W mt K"); Nic iw and Niv.nro are Nusselt numbers of laminar and turbulent flow, respectively [26]. Nit, yy =0.664(Re, }(Pr)"* @) tam 0.037(Re, )* Pr Ni = y O) °° 1+2.443(Re, )"(Pr**. 546 Reza Omran eral. (Energy Procedia 160 (2019) $42-549, where Re; is Reynolds number. 3. Results and discussion 3.1, Hydrogen supply from MH canisters without heat supply The hydrogen flow rate required by the FC is calculated using Eq, 2 and is presented in Fig. 2(a). The maximum hydrogen discharge rate of onc MH canister at different temperatures is presented in Fig. 2(b) based on the data from the paper by Tetuko et al. (21) a be Model Z x Experimenta valves a 0 500 1000 1500 2000 2500 so 5 1 1s 20 25 Power (W) MH temperature (C) Fig. 2 (a) Hydiogen flowrate requted by the call stack at diferent power (b) hydrogen discharge flow rate ofa MH canister at different temperature [21] For the case without heat supply, heat transfer through the natural convection is calculated for MH canister using Eq, 6 and for different ambient temperatures. The results are shown in Fig. 3. Also, the MH canisters heat demand while supplying the maximum flow rate at each temperature is presented in this figure, which is calculated using Eq, 3 and based on the values reported by Tetuko et al. (21). From this graph, the equilibrium temperature for the MH canister at each ambient temperature (,) can be obtained. For instance, the MH equilibrium temperature is 1.8 °C, 5.3 °C, and 8.7 °C when ambient temperature is 10 °C, 20 °C, and 30 °C, respectively. Then the flow rate at each ambient temperature can be obtained from Eq. 2 or Fig. 2(b) considering the MH equilibrium temperature. eat transfer WV MUL emperatere CC) Fig. 3. Heat transfer fiom a MH canister at different ambient temperature andthe heat demand by a MH canister [21] a different MH temperatures ‘The number of MH canisters needed to supply the required hydrogen to run the FC at different powers is calculated for different ambient temperature and presented in Table 1. As it ean be scen, by an increase in the power, a significant number of MH canisters are required, and it is more significant for lower ambient temperature, This illustrates the Reza Onan etal (Energy Procedsa 160 (2019) 342-549 oo need for heat supply requirement to the MH canisters to avoid the use of excessive number of MH canister and to reduce the cost, size, and volume of the system. ‘Table 1. Number of MH required at different power and ambient temperare without thermal heat transfer to MH storage Fuel cell Ambient Number of | Fuel eat Ambient ‘Number of power (W) temperature (°C) MHzequited } power (W) temperature (°C) MH sequized 10 9 10 D 500 20 1 1500 20 2 30 6 30 19 10 » 0 a 1000 20 18 2000 20 31 30 2 30 7 3.2. Thermal coupling between fuel cell and MH canisters through forced convection By mounting the MH canisters, arranged in-line and in two rows, on the exhaust cooling stream side of the FC stack, the heat from the FC exhaust air can be transferred to the MH canisters through forced convection. Therefore, the heat transfer rate can be calculated for different powers using the Gnielinski correlations (Eq. 7, 8, and 9) and considering the correction for flow passing over a bank of tubes [26]. The speed of air is obtained from the supplier for different powers, At ambient temperature of 20 °C, the exhaust air temperature of 30 °C + 1, 38.5 °C + 1, 46° 1, and 46 °C = 1 was achieved for output powers of 500 W, 1000 W, 1500 W, and 2000 W, respectively, by running, the Horizon FC stack, ‘The heat transfer rate from the fuel cell exhaust air to the MH canisters at different temperatures is calculated and presented in Fig. 4. For these calculations, the transverse and longitudinal spacing between the canisters are considered 4.5 em (half of the canister diameter) and 9 em, respectively to account for the flow passing over a bank of tubes; the effective length of the MH canister is considered 25 cm (the width of the FC exhaust) Mitemperatere(©) Fig. 4, Forced convection to the MH canisters fom the fel cell exhaust ais at different power outputs and the heat demand by a MH canister [21] at diferent MH temperatures Comparing the results with the ease with no heat supply (Fig. 3), it ean be seen that the equilibrium temperature of the MH is significantly increased by the heat supply from the fuel cell exhaust air. This helps to reduce the mumber of ‘MH needed to supply the hydrogen demand of the fuel cell. If we assume that the MHI temperature is increased initially to the equilibrium temperature by an external heat source, the number of MH canisters required for different powers at ambient temperature of 20 °C is calculated and presented in Table 2. Compared to the values obtained for the case without heat supply, a noticeable reduction in the number of MH canisters can be observed. Using Eq. 1, the heat sis Reza Omran eral. (Energy Procedia 160 (2019) $42-549, demand by one MH canister at equilibrium temperature is 11.7%, 8.3%, 6.8%, and 4.6% of the FC waste heat for output powers of 500 W, 1000 W, 1500 W, and 2000 W, respectively. Table 2. The number of 800-s1 MH canisters required to supply the hydrogen demand by the fuel at different power with heat supplied to the MET canisters fiom the fuel cll exhaust ar and ambient temperanuse of 20 °C. Fueleell __ Mequilituiumn Fuel cell Mifflow nie Min of MET power(W) Temp. (°C) demand (SLPM) (SLPMp) seqquired 300 165 3 129 + 1000 253 88 180 5 1500 2 136 226 1 2000 3M 1895 236 ° 4, Conclusion A mathematical model was developed to predict the number of MH canisters required for an open-cathode PEMFC bby considering the hydrogen supply rate limitation of the canisters, The number of metal hydrides required was calculated both with and without heat supply from the fuel cell exhaust in terms of the effect on hydrogen supply rate For the case without any heat supply to MH storage (ic. natural convection) is was found that for higher powers (1.¢ greater than 1000 W) significantly higher number of MH canisters are required and it hecomes even more critical at lower ambient temperatures. This can result in excessive overall system weight, size, and cost. Utilising the waste heat of fuel cell by passing the fuel cell exhaust heat over the MH canisters, was shown to help reduce the number of the MH canisters significantly. For instance, at ambient temperature of 20 °C, the number of MH canisters required was reduced by ~ 45%, 679%, 68%, and 69% for output powers of 500 W, 1000 W, 1500 W, and 2000 W respectively. Further experiments and investigations are required to model the transient response of the MH canisters in both conditions (with and without heat supply), Furthermore, to enhance the heat transfer to the MH canisters, addition of fins can be considered (i.e. to increase the effective heat transfer area) to further reduce the number of MH canisters required. It is noteworthy the total hydrogen required to be stored should be considered as another limiting factor when reducing the number of MH canisters. References [1] Shabani, Bahman, and John Andrews, "Standalone solar-hydrogen systems powering fire contingency networks.” International Journal of Hydrogen Energy 40(15) (2015): 509-5517. [2] Maniatopoutos, Paul, John Andrews, and Bahman Shabani, "Towards a sustainable strategy for road transportation ‘in Australia: the potential contribution of hydrogen.” Renewable and Sustainable Energy Reviews 52 (2015): 24 34 [3] Islam, R., B. Shabani, J. Andrews, and G. Rosengarten. "Experimental investigation of using ZnO nanofiuids as coolants ina PEM fuel cell.” International Journal of Hydrogen Energy 42(30) (2017): 19272-19286. [4] Shabani, Bahman, Reza Omrani, and John Andrews. "Energy security and sustainability for road transport sector: the role of hydrogen fuel cell technology”, in Amritanshu Shukla and Atul Sharma (Eds) Energy Security and Sustainability, (2017) Boca Raton, Taylor & Francis [5] Shabani, Bahman, and John Andrews, "An experimental investigation of a PEM fuel cell to supply both heat and power in a solar-hydrogen RAPS system." International Journal of Hydrogen Energy 36(9) (2011): 5442-5452. [6] Islam, MR, B Shabani, G Rosengarten, and J Andrews. "The potential of using nanofluids in PEM fuel cell cooling systems: A review." Renewable and Sustainable Energy Reviews 48 (2015): 523-539 [7] Shabani, Bahman, John Andrews, and Sukhvinder Badwal. "Fuel cell heat recovery, electrical load management, and the economics of solar-hydrogen systems." International Journal of Power & Energy Systems 30(4) (2010): 256. Reza Omran eral. (Energy Procedia 160 (2019) $42-549, 349 [8] Neuyen, Huy Quoc, Asma Mohamad Aris, and Bahman Shabani. "PEM fuel cell heat recovery for preheating inlet air in standalone solar-hydrogen systems for telecommunication applications: An exergy analysis.” International Journal of Hydrogen Energy 41(4) (2016): 2987-3003. [9] Assaf, Jihane, and Bahman Shabani. "Transient simulation modelling and energy performance of a standalone solar-hydrogen combined heat and power system integrated with solar-thermal collectors.” Applied Energy 178 (2016): 66-77. [10] Shabani, Bahman. "Solar-hydrogen combined heat and power systems for remote area power supply." (2010). [11] Assaf, Jihane, and Bahman Shabani. "Multi-objective sizing optimisation of a solar-thermal system integrated with a solar-hydrogen combined heat and power system, using genetic algorithm." Energy Conversion and Management 164 (2018): 518-532. [12] Chung, CA, Su-Wen Yang, Chien-Yuh Yang, Che-Weu Hsu, and Pai-Yuh Chiu. "Experimental study on the hydrogen charge and discharge rates of metal hydride tanks using heat pipes to enhance heat transfer." Applied Energy 103 (2013): 581-587 [13] Lototskyy, Mykhaylo V, Ivan Tolj, Lydia Pickering, Cordellia Sita, Frano Barbir, and Volodymyr Yartys, "The ‘use of metal hydrides in fuel cell applications." Progress in Natural Science: Materials International 21(1) (2017): 3-20. [14] Kim, Kwang J., Blanea Montoya, Arsalan Razani, and K. H. Lee. "Metal hydride compacts of improved thermal conductivity.” international journal of hydrogen energy 26(6) (2001): 609-613. [15] Tetuko, Anggito P, Bahman Shabani, and John Andrews, "Thermal coupling of PEM fuel cell and metal hydride hydrogen storage using heat pipes." international journal of hydrogen energy 41(7) (2016): 4264-4277, [16] Forde, T, J Eriksen, AG Pettersen, PJS Vie, and © Ulleberg, "Thermal integration of a metal hydride storage unit and a PEM fuel cell stack." international journal of hydrogen energy 34(16) (2009): 6730-6739, [17] Wilson, P.R., RC. Bowman, J. L. Mora, and J. W. Reiter. "Operation of a PEM fuel cell with LaNi4.8Sn0.2 hydride beds." Journal of Alloys and Compounds 446-447 (2007): 676-680. [18] Weiss-Ungethiim, J6ng, Inga Birger, Niko Schmidt, Mare Linder, and Josef Kallo, "Experimental investigation of a liquid cooled high temperature proton exchange membrane (HT-PEM) fuel cell coupled to a sodium alanate tank," international journal of hydrogen energy 3911) (2014): 5931-5941 [19] Urbanezyk, Robert, S Peil, D Bathen, C HeBike, J Burfeind, K Hauschild, M Feldethoff, and F Schiith. "HT - PEM Fuel Cell System with Integrated Complex Metal Hydride Storage Tank.” Fuel Cells 11(6) (2011): 911-920. [20] Pfeifer, P, C Wall, O Jensen, H Hahn, and M Fichtner. "Thermal coupling of a high temperature PEM fuel cell with a complex hydride tank." international journal of hydrogen energy 34(8) (2009): 3457-3466. [21] Tetuko, Anggito P, Bahman Shabani, Reza Omrani, Biddyut Paul, and John Andrews. "Study of a thermal bridging approach using heat pipes for simultaneous fuel cell cooling and metal hydride hydrogen discharge rate enhancement." Journal of Power Sources 397 (2018): 177-188, [22] P. Tetuko, Anggito, Bahman Shabani, and John Andrews. "Passive Fuel Cell Heat Recovery Using Heat Pipes to Enhance Metal Hydride Canisters Hydrogen Discharge Rate: An Experimental Simulation.” Energies 11(4) (2018): 913. [23] Tetuko, Anggito P., Bahman Shabani, and John Andrews. "Passive Fuel Cell Heat Recovery Using Heat Pipes to Enhance Metal Hydride Canisters Hydrogen Discharge Rate: An Experimental Simulation.” Energies 11(4) (2018): 913. [24] Larminie, J, and A Dicks. Fuel Cell Systems Explained, 2nd Edition, Wiley, 2003 [25] Churchill, Stwart W., and Humbert H. S. Chu, "Correlating equations for laminar and turbulent free convection from a horizontal cylinder." International Journal of Heat and Mass Transfer 18(9) (1975): 1049-1053. [26] Gnielinski, Volker. "Berechnung mittlerer Warme- und Stoffibergangskoeffizienten an laminar und turbulent liberstrimten Einzelkorpe mit Hilfe einer einheitlichen Gleichung.” Forschung im Ingeniewwesen A 41(5) (1975): 145-153

You might also like