Elsevier

Chemical Engineering Science

Volume 123, 17 February 2015, Pages 81-91
Chemical Engineering Science

Heat integration and optimization of hydrogen production for a 1 kW low-temperature proton exchange membrane fuel cell

https://doi.org/10.1016/j.ces.2014.10.036Get rights and content

Highlights

  • A hydrogen production system via MSR was designed for a 1 kW LT-PEMFC.

  • The optimized system heat efficiency (86.1%) was obtained using the pinch method.

  • The maximum energy recovery was obtained in the hydrogen production process.

  • The system efficiency was evaluated with the sensitivity analysis.

  • The device insulation and catalyst performance possess the system efficiency.

Abstract

Heat integration and process optimization for the hydrogen production units of 1 kW low-temperature proton exchange membrane fuel cells (LT-PEMFCs) were studied. Hydrogen is produced from methane in a process consisting of a steam methane reforming reaction (SMR), a high-temperature water–gas shift reaction (HWGS), a low-temperature water–gas shift reaction (LWGS), and a preferential oxidation reaction (PROX). Using commercial Aspen Plus® software, the process was designed and simulated with the objective of energy recovery. Furthermore, the unit operation parameters were discussed based on the mass/heat balances. Pinch analysis was applied for heat integration and process optimization to improve the energy recovery level of the hydrogen production system. By integrating system heating, a maximum energy recovery network (MER) and also a relaxed network were developed. The efficiencies of the hydrogen production systems designed with these MER and relaxed networks were 84.3% and 80.1%, respectively. Compared with the MER design, the relaxed network greatly simplifies the hydrogen production process and reduces the exchanger number from 18 to 9. The penalty in terms of power for relaxation is 120 W. Sensitivity analysis shows that heat loss and the steam-to-carbon (S/C) ratio are the primary factors influencing the system efficiency, i.e., a 5% heat loss may result in a 7% efficiency drop, and increasing the S/C ratio from 2.5 to 3.0 may result in a 2.5% efficiency drop. The conversion of methane (XCH4) and the relative deviation from chemical equilibrium for the WGS reactions (both HWGS and LWGS) are secondary factors: efficiency drops 3.5% at XCH4=0.8 and at 0.9 of the WGS equilibrium, compared with equilibrium conditions. Finally, the minimum pinch temperature difference (ΔTmin) and reaction temperatures have little impact on system efficiency. Therefore, the device insulation and catalyst performance are very important for system performance.

Introduction

Proton exchange membrane fuel cells (PEMFCs) are commonly used commercial fuel cells which convert hydrogen to electricity and heat with a significantly high degree of energy efficiency. The net electrical efficiency of a PEMFC can reach 40–58% (based on the lower heating value (LHV)), which is much higher than the limit of the Carnot cycle (Arsalis et al., 2011a). Today, PEMFCs are used in many wide-ranging applications such as power stations, portable devices, electric vehicles, space shuttles, and combined-heat-and-power (CHP) systems (Holladay et al., 2009, Men et al., 2008).

Currently, the steam methane reforming (SMR) reaction is the primary method for obtaining hydrogen and is used in numerous PEMFC processes (Balasubramanian et al., 1999, De Jong et al., 2009, Di Bona et al., 2011, Kamarudin et al., 2004, Stutz and Poulikakos, 2005). Even though it is arguable whether PEMFCs offer better energy efficiency than when starting from fossil fuels, the benefits of PEMFCs are clear when using distributed methane sources such as small oil fields or biomethane. Methane is stored and transported much more easily than pure hydrogen, particularly for mobile fuel cell applications (Arsalis et al., 2011b, Hubert et al., 2006, Kamarudin et al., 2004, Seo et al., 2006).

The hydrogen production process for PEMFCs includes the series of reactions listed in Table 1. The reforming reaction is strongly endothermic and consumes the most heat energy in the process. The subsequent reactions—the water–gas shift reactions (WGS), preferential oxidation reaction (PROX), and combustion of the depleted fuel—are exothermic, and the heats of these reactions can be used as SMR heating sources. If the system heat is unbalanced, a supplementary methane fuel supply is needed. Because the system is composed of many streams with different temperatures, the proper organization of heat exchangers and subunits can greatly improve system efficiency and reduce supplementary utility use.

The practice of heat integration is widely applied in chemical and petroleum processes (Kemp, 2007, Linnhoff et al., 1983, Linnhoff and Hindmarsh, 1983, March, 1998). Pinch analysis is a useful tool for analyzing the heating network and achieving more efficient heating strategies. To our knowledge, for low-temperature proton exchange membrane fuel cells (LT-PEMFCs), most research has been focused on the membrane, stack performance, unit design, or CHP system efficiency, but very few studies have been concerned with the heat integration and process optimization of the hydrogen production process. Pasdag et al. (2012) reported an analysis of the possible heat integration of a fuel processor coupled with a high-temperature (HT)/LT-PEMFC. The electrical efficiency increased about 5.5% when a condensing burner technology was adopted. Godat and Marechal (2003) also optimized a fuel cell system by using process integration techniques. The fuel cell and fuel processing subsystems were contained, and four operating conditions (i.e., steam-to-carbon (S/C) ratio, SMR temperature, PEMFC temperature, and fuel utilization) were optimized for higher efficiency. The overall efficiency was raised from 35% for the reference system to 49% in the optimized design. However, the heat exchange networks for the integrated system and process flow diagrams were not involved in their study.

In this work, the hydrogen production system for a 1 kW LT-PEMFC system was studied. Using commercial Aspen Plus® process simulation software, the hydrogen production process and operating parameters were simulated and optimized. Pinch analysis was applied to integrate the heating networks of the hydrogen production system. More practically, a heat exchange network was designed based on maximum energy recovery (MER) and relaxed networks. Sensitivity analyses that considered the minimum pinch temperature difference (ΔTmin), the relative deviation from chemical equilibrium for both SMR and WGS, the S/C ratio, the reaction temperatures, and heat losses were also investigated. Our aim was to achieve higher system efficiency and balance the investment costs. This work can further guide reformer design and structure optimization in subsequent studies.

Section snippets

Process models

Commercial process simulation software (Aspen Plus®, AspenTech, Burlington, MA, USA) was used for the optimization of operations and the calculations of mass and heat balances. A schematic flow diagram for a 1 kW LT-PEMFC system is shown in Fig. 1, and the corresponding abbreviations are defined in Table 2. The model is based on a zero-dimensional approach with steady and isothermal operation (Tanim et al., 2013). All the working fluids are calculated using the Peng–Robinson equation of state

Integration method

Pinch analysis is a widely used process integration method in the petroleum and chemical industries, as well as metallurgy, papermaking, food, beverages, and many other fields (Godat and Marechal, 2003). It was proposed by Linnhoff and his team as a heat exchange network design and optimization method in the late 1970s, with the aim of maintaining a four-dimensional balance of energy savings, productivity improvements, reduced investments, and environmental protection (Kemp, 2007, March, 1998).

Process designs

The designed hydrogen production processes according to the MER and relaxed heat network optimizations are shown in Fig. 8. The values in black indicated on the heat exchangers represent the inlet and outlet temperatures of the streams, while the red values represent the heat duties of the different heat exchangers. Obviously, process (a) is much more complex than (b). Streams in Fig. 8a flow through 18 heat exchangers which will lead to complex links and higher heat losses. In contrast,

Conclusions

A hydrogen production system for a 1 kW LT-PEMFC was designed and discussed in this paper. The system is composed of a series of units, including a steam methane reformer, water–gas shift reactors (both HWGS and LWGS), a preferential oxidation reactor, water evaporator, a fuel/anodic off-gas burner, and many heat exchangers. The operating conditions were optimized: for the SMR, a temperature, pressure, and S/C ratio of 1000 K, 1 atm, and 3.0, respectively; for the HWGS, the temperature and

Notation

    PEMFC

    proton exchange membrane fuel cell

    LT-PEMFC

    low temperature PEMFC

    HT-PEMFC

    high temperature PEMFC

    SMR

    steam methane reforming

    CHP

    combined-heat-and-power

    WGS

    water–gas shift

    HWGS

    high-temperature water–gas shift

    LWGS

    low-temperature water–gas shift

    PROX

    preferential oxidation

    XCH4

    conversion of methane

    S/C ratio

    mole ratio of steam to methane

    ΔT

    heat transfer temperature difference

    ΔTmin

    minimum pinch temperature difference

    T–H

    temperature–enthalpy diagram method

    CP

    heat capacity flow rate

    Cp

    the specific heat capacity

Acknowledgments

This work is sponsored by the National 863 Program of China (Grant no. 2012AA053402) and the Strategic Program of Sichuan Province (Grant no. 12XXCP0096). The authors would like to acknowledge China Chengda Engineering Co., Ltd. for its software support in this work.

References (31)

Cited by (8)

  • Process simulation and energy integration in the mineral carbonation of blast furnace slag

    2019, Chinese Journal of Chemical Engineering
    Citation Excerpt :

    According to the pinch rule, this heat transfer across the pinch point may cause extra energy consumption [40,41]. These exchangers can be eliminated by replacing the network with a more complex network, for example, by splitting stream 1 into four streams and matching it with other streams to eliminate heat exchanger E4, as recommended by the Aspen energy analyzer, which means more exchangers are used [54]. To reuse the waste heat of the solid streams, the sensible heat of hot solid was used to heat the air; then, the hot air releases the heat to the cold solid and absorbs the heat of the hot solid streams again.

  • Thermo-environmental life cycle assessment of hydrogen production by autothermal reforming of bioethanol

    2017, Energy for Sustainable Development
    Citation Excerpt :

    For this reason, carbon is excluded from the component list as plausible product since all of the considered configurations have SC > 3 and T > 300 °C. In the second step, the SG exiting the reformer is passed through a CO cleanup section introduced to guarantee H2 production with a CO content compatible with fuel cell specifics (levels below 10 ppm (Lei et al., 2015). The clean-up section is made up by WGS and CO preferential oxidation reactors (COPROX) (Salemme et al., 2009).

  • Energetic, exergetic and environmental life cycle assessment analyses as tools for optimization of hydrogen production by autothermal reforming of bioethanol

    2016, International Journal of Hydrogen Energy
    Citation Excerpt :

    CO + 0.5 O2 → CO2 In a COPROX reactor, a noble metal catalyst is employed and CO is selectively oxidized to CO2 with trace air [57]. Meanwhile, a small amount of H2 is also oxidized to H2O (Eq. (13)).

  • A thermo-environmental study of hydrogen production from the steam reforming of bioethanol

    2016, Journal of Energy Storage
    Citation Excerpt :

    Consequently, the gas leaving the LT-WGS is mixed with air before entering the COPROX reactor, where the remaining CO is oxidized to CO2 via Eq. (7).4 CO + 0.5O2 → CO2 In a COPROX reactor a noble metal catalyst is employed, and CO is selectively oxidized to CO2 with trace air [57]. Meanwhile, a small amount of H2 is also oxidized to H2O (Eq. (8)).

View all citing articles on Scopus
View full text