Next Article in Journal
Multi-Objective Unified Optimal Control Strategy for DAB Converters with Triple-Phase-Shift Control
Next Article in Special Issue
Numerical and Experimental Study of a Novel Additively Manufactured Metal-Polymer Composite Heat-Exchanger for Liquid Cooling Electronics
Previous Article in Journal
Computing and Assessment of Discrete Angle Positions for Optimizing the Solar Energy Harvesting for Urban Sustainable Development
Previous Article in Special Issue
Performance Improvement of a Double-Layer Microchannel Heat Sink via Novel Fin Geometry—A Numerical Study
 
 
Reply published on 9 October 2021, see Energies 2021, 14(20), 6445.
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Comment

Comment on Hamayun et al. Evaluation of Two-Column Air Separation Processes Based on Exergy Analysis. Energies 2020, 13, 6361

Department of Chemical and Biochemical Engineering, Faculty of Chemical and Food Technology, Slovak University of Technology in Bratislava, Radlinského 9, 812 37 Bratislava, Slovakia
*
Author to whom correspondence should be addressed.
Energies 2021, 14(20), 6443; https://doi.org/10.3390/en14206443
Submission received: 18 June 2021 / Accepted: 27 September 2021 / Published: 9 October 2021
(This article belongs to the Special Issue Modelling of Thermal and Energy Systems)

Abstract

:
Oxygen production from air belongs to energy-intense processes and, as a result, possibilities for its decrease are a frequent topic of optimization studies, often performed with simulation software such as Aspen Plus or Aspen HYSYS. To obtain veritable results and sound solutions, a suitable calculation method hand in hand with justified assumptions and simplifications should form the base of any such studies. Thus, an analysis of the study by Hamayun et al., Energies 2020, 13, 6361, has been performed, and several weak spots of the study, including oversimplified assumptions, improper selection of a thermodynamic package for simulation and omission of certain technological aspects relevant for energy consumption optimization studies, were identified. For each of the weak spots, a recommendation based on good praxis and relevant scientific literature is provided, and general recommendations are formulated with the hope that this comment will aid all researchers utilizing Aspen Plus and Aspen HYSYS software in their work.

1. Introduction

Oxygen is among the most frequently produced and consumed technical gases. It has found wide application in various industrial branches. Oxycombustion in the heat and power production sector [1,2] benefits from higher thermal efficiency of combustors [3] and subsequently reduced fuel consumption and lower carbon footprint [4,5], simultaneously offering the possibility for carbon capture and storage [6,7]. This is even more pronounced in the production and treatment of metals [8,9], where the achievable fuel and greenhouse gas reduction potential is substantial [10,11]. Oxygen use in gasification processes instead of air leads to products with increased heating value [12] and to reduced equipment size due to lower volumetric flows of the gaseous product [13]. As a result of the current COVID crisis, its importance as a medical gas has been fully revealed [14,15] and its production capacities are strained, which draws further attention to its impact on the environment [16,17].
Its production on an industrial scale is usually performed via cryogenic air separation [18,19]. Other techniques such as separation by membranes [20,21] or adsorption [22,23] have the potential to reduce the specific energy consumption for oxygen production but are still under development [24,25].
Cryogenic air separation is an energy-intensive process, which is the reason for the ongoing research effort for its design [26,27] and operation [28,29] optimization. It is, thus, of utmost importance that mathematical modeling of such units employs suitable thermodynamic models and adopts consistent and technically feasible assumptions. In this regard, we thoroughly analyzed the paper by Hamayun et al. (2020) [30] published previously in the journal Energies. The performed analysis and the presented results are followed by generally valid recommendations for the modeling of cryogenic air separation units.

2. General Comments

Hamayun et al. (2020) [30] set up and analyzed seven different configurations of a cryogenic air separation plant with a processing capacity of 500 th−1 of inlet air. They performed the underlying steady-state mathematical modeling in the Aspen Plus V10 environment using the Peng–Robinson property package. The Petlyuk-like column configuration came out as the most energy-effective one. In this section, general comments are formulated, while Section 3 provides more detailed comments on the plant scheme incorporating the Petlyuk-like column configuration, supported by our own simulations.
Firstly, model assumptions as formulated in [30] are incomplete. The pressure drop in heat exchangers assuming a constant value of 10 kPa, regardless of the position in the process scheme, is unjustified, similar to assuming zero pressure losses as in [26]. A more feasible approach is to adopt percentual pressure losses related to air inlet pressure. The Darcy–Weisbach equation, Equation (1), and the ideal gas state equation show that frictional pressure losses of gases and vapors depend on actual pressure (see Equation (2)), which justifies the proposed approach. This is also consistent with the approach of Szablowski et al. (2021) [31] where the assumed pressure drop in heat exchangers incorporated in the mathematical model of an adiabatic liquid energy storage system is related to the material stream density.
P = f L D ρ   v 2 2
P = f L D p M R T   v 2 2 C P
where ∆P is the frictional pressure loss, Pa; f is a friction coefficient; L is the length and D is the inner diameter of the pipeline, both in m; ρ is the material stream density, kgm−3; v is the flow velocity, ms−1; P is pressure of material stream, Pa; M is its molar mass, kgmol−1; R is the universal gas constant; and T is the thermodynamic temperature of the material stream, in Kelvins. Assuming an adiabatic flow and small frictional pressure losses, then all variables in Equation (2) are almost constant and can, except for pressure, be merged into one constant, C.
Secondly, the authors omitted pressure losses in adsorbers. The issue of pressure loss in adsorbers is of serious concern and can be subject to optimization [32]. It contributes to the energy consumption of the air separation unit and should thus be considered.
Thirdly, adsorbers are modeled as component splitters, the assumption of which is oversimplified. The effect of water steam adsorption heat should be considered as it may reach up to 3000–4000 kJ·kg−1 of adsorbed steam [33] for conventional zeolites used in compressed air drying by adsorption. The resulting temperature increase in air passing through the adsorbent layer can thus exceed 10 or even 20 °C depending on the water steam content in the inlet air which, in turn, impacts the equipment downstream.
Fourthly, the energy consumption evaluation in [30] is incomplete as it does not incorporate the energy needed for adsorber regeneration. As mentioned above, a significant amount of heat is released by steam adsorption on adsorbent, and thus, its regeneration is energy-intensive and contributes to the overall energy consumption of the air separation plant. Heat recuperation is often proposed to cut down the adsorbent regeneration costs [19,34], which, however, adds further complexity to the plant.
Fifthly, moist air cooling in a multi-stream heat exchanger directly to −100 °C and below after its intake from an ambient environment, as depicted in process scheme C7 in [30], is technically infeasible. It would lead to ice formation and air path blockage, possibly followed by heat exchanger damage.
The comments above document that proper and careful formulation of assumptions accompanied with a detailed check of each equipment piece modeled is a must before mathematical modeling. Moreover, the aspects discussed above affect the final energy consumption of the air separation plant in individual configurations as the positions of the given equipment vary, as do their thermal duties and the mass flows of moisture removed from air.
Finally, the importance of using a proper thermodynamic package should be addressed. The Peng–Robinson equation is recommended for applications comprising nonpolar gases and vapors [35], which holds true for nitrogen and oxygen or for dry air, but certainly not for water steam. We deal with this issue more deeply together with the chosen approach in Section 3.

3. Data Analysis Method

Following the model description and assumptions stated in [30], an identical model was set up in Aspen Plus® (Aspen Plus® V11, Aspen Technology Inc., Bedford, MA, USA) and Aspen HYSYS (Aspen HYSYS® V11, Aspen Technology Inc., Bedford, MA, USA). The scheme of the model set up in the Aspen Plus environment is shown in Figure 1.
Models in both Aspen Plus® and Aspen HYSYS served for verification of the model calculation results shown in [30]. Peng–Robinson as well as HYSPR (Peng–Robinson for HYSYS) thermodynamic packages were tested for compression section modeling in Aspen Plus while Peng–Robinson was applied in Aspen HYSYS.
The Peng–Robinson equation of state is considered an ideal thermodynamic model to predict the phase equilibria during cryogenic air separation [36]. However, the model parameters and settings differ considerably between Aspen Plus and Aspen HYSYS.
The HYSPR property method implements the Peng–Robinson property package from Aspen HYSYS into Aspen Plus. Several enhancements to the original Peng–Robinson model were made to extend its range of applicability and to improve its predictions for non-ideal systems [36].
First, Aspen Plus and Aspen HYSYS use different databanks for the estimation of bi-nary interaction parameters. Although most of the binary parameters are similar or the same, some of them, namely the parameters regarding the binary interactions of water, are different or missing.
Second, the way each model treats the missing parameters differs significantly. HYSPR estimates the missing parameters from critical volume using Equation (3):
k i j = 1 V c , i V c , j 1 / 6 0.5 V c , i 1 / 3 + V c , j 1 / 3
where k i j is the binary interaction parameter, V c is the critical volume and i and j are the respective components. However, no information is provided on how Aspen Plus treats the missing parameters, and according to the model results, it can be assumed that the Aspen Plus Peng–Robinson model considers no interaction where the information is missing. Furthermore, each model uses a different alpha function, although according to the values of acentric factors, this shall not make a difference when considering only the main air constituents.
Finally, and most importantly, specific attention needs to be given to the way both models treat water. Even though the user is asked about the steam tables to be used in the main settings pane, the default option of the Peng–Robinson model is not to use the steam tables. Hence, in order to attain comparable results, this option needs to be altered.

4. Results and Discussion

Simulation results for phase separators F1 to F4 are provided in Table 1. As can be seen, the results from Aspen Plus match those in [30] very closely. On the contrary, Aspen HYSYS yielded significantly lower mass flows of water condensate separated from air in the phase separators. To explain the different model outcomes, the Aspen Plus model was adjusted, switching from the standard Peng–Robinson package to the HYSPR package. The simulation results for phase separators F1 to F4 from Aspen Plus using the HYSPR package were compared with those from Aspen HYSYS as well as with data provided in [30] (Table 2). Switching from the Peng–Robinson to the HYSPR package in Aspen Plus resulted in data that are almost identical to those obtained from Aspen HYSYS. Thus, it can be stated that the differences in mass flows of water condensate from air visible in Table 1 resulted from the differences in the thermodynamic Peng–Robinson package implemented in Aspen Plus and in Aspen HYSYS.
The mass flow of condensed water in exchanger E2 was further checked by hand calculation employing Equations (4)–(7). The mass flow of dry air was calculated first (Equation (4)), followed by calculation of the relative mass fraction of water vapor in air (Equation (5)). In the equations, m ˙ G and m ˙ S 7 are the mass flows of dry air and inlet air to heat exchanger E2, respectively, in kgh−1; w H 2 O is the mass fraction of water in inlet air; and   Y H 2 O is the relative mass fraction of water vapor in inlet air. Y H 2 O * is the equilibrium relative mass fraction of air leaving heat exchanger E2, which can be calculated by Equation (6), where P H 2 O ° denotes the saturated vapor pressure of water, Pa; P is the overall pressure, Pa; and M H 2 O and M G represent the molar masses of water and dry air, respectively, in kgmol−1. The saturated vapor pressure of water at 40 °C, which is the air exit temperature from heat exchanger E2, was adopted from steam tables. The mass flow of water condensate from air in heat exchanger E2, m ˙ H 2 O , kgh−1, was calculated by Equation (7), and the results are shown in Table 3, where they are compared with the mass flows of condensed water obtained by Aspen Plus using the standard Peng–Robinson package and by Aspen HYSYS.
m ˙ G = m ˙ S 7   1 w H 2 O = 500,000 kg h ·   1 0.00976 = 495.12   kg / h
Y H 2 O = w H 2 O 1 w H 2 O = 0.00976 1 0.00976 = 0.0098561
Y H 2 O * = P H 2 O ° P P H 2 O °   ·   M H 2 O M G = 7.375   kPa 590   kPa 7.375   kPa   ·   18.02 kg kmol 28.96 kg kmol = 0.0078764
m ˙ H 2 O = m G ˙ · Y H 2 O Y H 2 O * = 495,120 kg h · 0.0098561 0.0078764 = 980.2   kg / h
As can be seen in Table 3, the hand calculation yielded values that are much closer to the results from Aspen HYSYS than to those from Aspen Plus using the standard Peng–Robinson package. This further corroborates the previous findings that the Peng–Robinson package implemented in Aspen HYSYS is more suitable for modeling processes involving water substance than the standard Peng–Robinson package in Aspen Plus.
Different predicted mass flows of water condensate separated in the phase separators further affect the water load of adsorbers A1 and A2. In reality, more water steam passes through the air compression and cooling train to the adsorbers than predicted in [30] using Aspen Plus. The difference in mass flows of water steam to be removed from compressed air in adsorbers A1 and A2 can be seen in Table 4 and Table 5.
The tables show that the mass flow of the remaining water steam to be removed is around 10% higher in Aspen HYSYS than in Aspen Plus using the standard Peng–Robinson thermodynamic package. This results in the need for a larger mass of adsorbent to be used in the adsorbers to maintain their operating time. Moreover, the adsorption heat release rate increases, and dry air leaves the adsorber at a higher temperature than that corresponding to the mass flows of adsorbed water predicted in [30]. It has to be stressed that the effect of adsorption heat is fully omitted in [30] as the adsorbers are treated as isothermal component splitters there. We consider such an approach to be oversimplified.

5. Recommendations for Future Work

From the detailed analysis of the cryogenic air separation unit simulation performed in [30], the following general recommendations can be drawn to increase the feasibility of the simulation results for similar or the same processes:
  • Regarding pressure losses in gas or vapor heat exchangers, assuming a constant pressure loss value regardless of the actual thermodynamic conditions is an oversimplification. Adopting a percentual value related to the actual pressure of the given medium is recommended.
  • Assuming isobaric–isothermal operation of water steam adsorbers from compressed air is infeasible. It is recommended to consider the adsorption heat effect properly and to consider air pressure loss in the adsorber at least equal to that in the preceding air cooler.
  • Energy consumption for compressed air dryer (adsorber) regeneration should be taken into account, especially when performing an energy consumption optimization study.
  • A proper thermodynamic package should be selected for modeling to avoid imprecise simulation results which may further affect the study findings and conclusions.

6. Conclusions

A thorough analysis of the study method and results presented in [30] was performed. Weak points of the method were discussed, such as the oversimplified assumptions, omitted thermal effects of water steam adsorption in adsorbers and the energy consumption for their regeneration as well as improper selection of the thermodynamic package, and recommendations for future work were formulated. It is hoped that the authors of [30] as well as other researchers find them helpful in their future research.

Author Contributions

Conceptualization, M.V.; methodology, D.J.; validation, P.F.; investigation, D.J. and M.V.; data curation, D.J.; writing—original draft preparation, M.V.; writing—review and editing, D.J. and P.F., visualization, D.J. All authors have read and agreed to the published version of the manuscript.

Funding

This study was supported by the Slovak Research and Development Agency under the contract no. APVV-18-0134.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

All data obtained by calculations and analyses are listed directly in this study.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

References

  1. Ghoniem, A.F.; Zhao, Z.; Dimitrakopoulos, G. Gas oxy combustion and conversion technologies for low carbon energy: Fundamentals, modeling and reactors. Proc. Combust. Inst. 2019, 37, 33–56. [Google Scholar] [CrossRef]
  2. Kotowicz, J.; Michalski, S.; Brzęczek, M. The Characteristics of a Modern Oxy-Fuel Power Plant. Energies 2019, 12, 3374. [Google Scholar] [CrossRef] [Green Version]
  3. Portillo, E.; Gallego Fernández, L.M.; Vega, F.; Alonso-Fariñas, B.; Navarrete, B. Oxygen transport membrane unit applied to oxy-combustion coal power plants: A thermodynamic assessment. J. Environ. Chem. Eng. 2021, 9, 105266. [Google Scholar] [CrossRef]
  4. Normann, F.; Andersson, K.; Johnsson, F.; Leckner, B. NOx reburning in oxy-fuel combustion: A comparison between solid and gaseous fuels. Int. J. Greenh. Gas. Control 2011, 5, S120–S126. [Google Scholar] [CrossRef]
  5. Bělohradský, P.; Skryja, P.; Hudák, I. Experimental study on the influence of oxygen content in the combustion air on the combustion characteristics. Energy 2014, 75, 116–126. [Google Scholar] [CrossRef]
  6. Fogarasi, S.; Cormos, C.-C. Assessment of coal and sawdust co-firing power generation under oxy-combustion conditions with carbon capture and storage. J. Clean. Prod. 2017, 142, 3527–3535. [Google Scholar] [CrossRef]
  7. Allam, R.; Martin, S.; Forrest, B.; Fetvedt, J.; Lu, X.; Freed, D.; Brown, G.W.; Sasaki, T.; Itoh, M.; Manning, J. Demonstration of the Allam Cycle: An Update on the Development Status of a High Efficiency Supercritical Carbon Dioxide Power Process Employing Full Carbon Capture. Energy Procedia 2017, 114, 5948–5966. [Google Scholar] [CrossRef]
  8. Royo, P.; Ferreira, V.J.; López-Sabirón, A.M.; García-Armingol, T.; Ferreira, G. Retrofitting strategies for improving the energy and environmental efficiency in industrial furnaces: A case study in the aluminium sector. Renew. Sustain. Energy Rev. 2018, 82, 1813–1822. [Google Scholar] [CrossRef] [Green Version]
  9. Dzurňák, R.; Varga, A.; Kizek, J.; Jablonský, G.; Lukáč, L. Influence of Burner Nozzle Parameters Analysis on the Aluminium Melting Process. Appl. Sci. 2019, 9, 1614. [Google Scholar] [CrossRef] [Green Version]
  10. Poskart, A.; Radomiak, H.; Niegodajew, P.; Zajemska, M.; Musiał, D. The Analysis of Nitrogen Oxides Formation During Oxygen—Enriched Combustion of Natural Gas. Arch. Metall. Mater. 2016, 61, 1925–1930. [Google Scholar] [CrossRef]
  11. Dzurňák, R.; Jablonský, G.; Varga, A.; Kizek, J.; Pástor, M.; Lukáč, L. Impact of oxygen enhanced combustion of natural gas on thermal efficiency of combustion aggregate. MATEC Web Conf. 2018, 168, 07016. [Google Scholar] [CrossRef]
  12. Gabbar, H.; Aboughaly, M.; Russo, S. Conceptual Design and Energy Analysis of Integrated Combined Cycle Gasification System. Sustainability 2017, 9, 1474. [Google Scholar] [CrossRef] [Green Version]
  13. Corti, A. Biomass integrated gasification combined cycle with reduced CO2 emissions: Performance analysis and life cycle assessment (LCA). Energy 2004, 29, 2109–2124. [Google Scholar] [CrossRef]
  14. Carnero, M.C.; Gómez, A. Optimization of Decision Making in the Supply of Medicinal Gases Used in Health Care. Sustainability 2019, 11, 2952. [Google Scholar] [CrossRef] [Green Version]
  15. Feinmann, J. How covid-19 revealed the scandal of medical oxygen supplies worldwide. BMJ 2021, 373, n1166. [Google Scholar] [CrossRef]
  16. Balys, M.; Brodawka, E.; Korzeniewska, A.; Szczurowski, J.; Zarebska, K. LCA and economic study on the local oxygen supply in Central Europe during the COVID-19 pandemic. Sci. Total. Environ. 2021, 786, 147401. [Google Scholar] [CrossRef]
  17. Shakil, M.H.; Munim, Z.H.; Tasnia, M.; Sarowar, S. COVID-19 and the environment: A critical review and research agenda. Sci. Total. Environ. 2020, 745, 141022. [Google Scholar] [CrossRef]
  18. Xu, J.; Wang, T.; Chen, Q.; Zhang, S.; Tan, J. Performance design of a cryogenic air separation unit for variable working conditions using the lumped parameter model. Front. Mech. Eng. 2019, 15, 24–42. [Google Scholar] [CrossRef]
  19. Wang, C.; Akkurt, N.; Zhang, X.; Luo, Y.; She, X. Techno-economic analyses of multi-functional liquid air energy storage for power generation, oxygen production and heating. Appl. Energy 2020, 275, 115392. [Google Scholar] [CrossRef]
  20. Adhikari, B.; Orme, C.J.; Klaehn, J.R.; Stewart, F.F. Technoeconomic analysis of oxygen-nitrogen separation for oxygen enrichment using membranes. Sep. Purif. Technol. 2021, 268, 118703. [Google Scholar] [CrossRef]
  21. Dibrov, G.; Ivanov, M.; Semyashkin, M.; Sudin, V.; Fateev, N.; Kagramanov, G. Elaboration of High Permeable Macrovoid Free Polysulfone Hollow Fiber Membranes for Air Separation. Fibers 2019, 7, 43. [Google Scholar] [CrossRef] [Green Version]
  22. Banaszkiewicz, T.; Chorowski, M. Energy Consumption of Air-Separation Adsorption Methods. Entropy 2018, 20, 232. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Akulinin, E.; Golubyatnikov, O.; Dvoretsky, D.; Dvoretsky, S. Optimization and analysis of pressure swing adsorption process for oxygen production from air under uncertainty. Chem. Ind. Chem. Eng. Q. 2020, 26, 89–104. [Google Scholar] [CrossRef]
  24. Fernandez-Barquin, A.; Casado-Coterillo, C.; Valencia, S.; Irabien, A. Mixed Matrix Membranes for O(2)/N(2) Separation: The Influence of Temperature. Membranes 2016, 6, 28. [Google Scholar] [CrossRef] [PubMed]
  25. Widiastuti, N.; Gunawan, T.; Fansuri, H.; Salleh, W.N.W.; Ismail, A.F.; Sazali, N. P84/ZCC Hollow Fiber Mixed Matrix Membrane with PDMS Coating to Enhance Air Separation Performance. Membranes 2020, 10, 267. [Google Scholar] [CrossRef] [PubMed]
  26. Caspari, A.; Offermanns, C.; Schäfer, P.; Mhamdi, A.; Mitsos, A. A flexible air separation process: 1. Design and steady-state optimizations. AIChE J. 2019, 65, e16705. [Google Scholar] [CrossRef]
  27. Singla, R.; Chowdhury, K. Enhanced oxygen recovery and energy efficiency in a reconfigured single column air separation unit producing pure and impure oxygen simultaneously. Chem. Eng. Process. Process Intensif. 2021, 162, 108354. [Google Scholar] [CrossRef]
  28. Adamson, R.; Hobbs, M.; Silcock, A.; Willis, M.J. Steady-state optimisation of a multiple cryogenic air separation unit and compressor plant. Appl. Energy 2017, 189, 221–232. [Google Scholar] [CrossRef] [Green Version]
  29. Elhelw, M.; Alsanousie, A.A.; Attia, A. Novel operation control strategy for conjugate high-low pressure air separation columns at different part loads. Alex. Eng. J. 2020, 59, 613–633. [Google Scholar] [CrossRef]
  30. Hamayun, M.H.; Ramzan, N.; Hussain, M.; Faheem, M. Evaluation of Two-Column Air Separation Processes Based on Exergy Analysis. Energies 2020, 13, 6361. [Google Scholar] [CrossRef]
  31. Szablowski, L.; Krawczyk, P.; Wolowicz, M. Exergy Analysis of Adiabatic Liquid Air Energy Storage (A-LAES) System Based on Linde–Hampson Cycle. Energies 2021, 14, 945. [Google Scholar] [CrossRef]
  32. Shigaki, N.; Mogi, Y.; Haraoka, T.; Sumi, I. Reduction of Electric Power Consumption in CO2-PSA with Zeolite 13X Adsorbent. Energies 2018, 11, 900. [Google Scholar] [CrossRef] [Green Version]
  33. Kim, H.; Cho, H.J.; Narayanan, S.; Yang, S.; Furukawa, H.; Schiffres, S.; Li, X.; Zhang, Y.B.; Jiang, J.; Yaghi, O.M.; et al. Characterization of Adsorption Enthalpy of Novel Water-Stable Zeolites and Metal-Organic Frameworks. Sci. Rep. 2016, 6, 19097. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Goodarzia, G.; Dehghani, S.; Akbarzadeh, A.; Date, A. Energy Saving Opportunities in air Drying Process in High-pressure Compressors. Energy Procedia 2017, 110, 428–433. [Google Scholar] [CrossRef]
  35. Haydary, J. Chemical Engineering Process. Design and Simulation-Aspen Plus and Aspen HYSYS Applications, 1st ed.; Wiley & Sons, Inc.: Hoboken, NJ, USA, 2019. [Google Scholar]
  36. Aspen Technology Inc. Aspen Plus User Guide, Version 10.2; Aspen Technology Inc.: Cambridge, MA, USA, 2000. [Google Scholar]
Figure 1. Mathematical model of air separation unit designed in Aspen Plus. Legend: A—adsorber; E—heat exchanger; F—phase separator; HPC—high-pressure column; K—compressor; LPC—low-pressure column; MHEX—multi-stream heat exchanger; P—pump; Q—energy stream; S—material stream except cooling water; V—valve; W—cooling water stream.
Figure 1. Mathematical model of air separation unit designed in Aspen Plus. Legend: A—adsorber; E—heat exchanger; F—phase separator; HPC—high-pressure column; K—compressor; LPC—low-pressure column; MHEX—multi-stream heat exchanger; P—pump; Q—energy stream; S—material stream except cooling water; V—valve; W—cooling water stream.
Energies 14 06443 g001
Table 1. Simulation results for phase separators F1 to F4. Press. = pressure; Temp. = Temperature.
Table 1. Simulation results for phase separators F1 to F4. Press. = pressure; Temp. = Temperature.
Aspen PlusAspen HYSYSStudy [30]
StreamMass Flow (kg/h)Temp.
(°C)
Press.
(kPa)
Mass Flow
(kg/h)
Temp.
(°C)
Press.
(kPa)
Mass Flow
(kg/h)
Temp.
(°C)
Press.
(kPa)
S5 (F1)040240040240040240
S15 (F2) 459.340590244.44059050040590
S11 (F3)97640590519.340590100040590
S19 (F4)215.740740252.24074020040740
Table 2. Simulation results for phase separators F1 to F4 using the thermodynamic package HYSPR in Aspen Plus.
Table 2. Simulation results for phase separators F1 to F4 using the thermodynamic package HYSPR in Aspen Plus.
Aspen PlusAspen HYSYSStudy [30]
StreamMass Flow (kg/h)Temp.
(°C)
Press.
(kPa)
Mass Flow
(kg/h)
Temp.
(°C)
Press.
(kPa)
Mass Flow
(kg/h)
Temp.
(°C)
Press.
(kPa)
S5 (F1)040240040240040240
S15 (F2) 244.440590244.44059050040590
S11 (F3)519.240590519.340590100040590
S19 (F4)252.240740252.24074020040740
Table 3. Simulation results: mass flows of water condensate in streams S7 and S19 obtained by simulations in Aspen Plus, Aspen HYSYS and by hand calculation (Equations (4)–(7)).
Table 3. Simulation results: mass flows of water condensate in streams S7 and S19 obtained by simulations in Aspen Plus, Aspen HYSYS and by hand calculation (Equations (4)–(7)).
StreamAspen PlusAspen HYSYSHand Calculation
S7 (kg/h)1435.3763.7980.4
S19 (kg/h)215.7252.2254.7
Table 4. Simulation results for adsorber A1.
Table 4. Simulation results for adsorber A1.
A1
Aspen PlusAspen HYSYSStudy [30]
StreamMass Flow (kg/h)Temp.
(°C)
Press.
(kPa)
Mass Flow (kg/h)Temp.
(°C)
Press.
(kPa)
Mass Flow (kg/h)Temp.
(°C)
Press.
(kPa)
S10339.040590339.540590339.040590
S12336.540590336.740590336.540590
S132.5405902.8405902.540590
Table 5. Simulation results for adsorber A2.
Table 5. Simulation results for adsorber A2.
A1
Aspen PlusAspen HYSYSStudy [30]
StreamMass Flow (kg/h)Temp.
(°C)
Press.
(kPa)
Mass Flow (kg/h)Temp.
(°C)
Press.
(kPa)
Mass Flow (kg/h)Temp.
(°C)
Press.
(kPa)
S18159.340740159.540740159.340740
S20158.440740158.440740158.440740
S20A0.9407401.1407401.040740
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Variny, M.; Jediná, D.; Furda, P. Comment on Hamayun et al. Evaluation of Two-Column Air Separation Processes Based on Exergy Analysis. Energies 2020, 13, 6361. Energies 2021, 14, 6443. https://doi.org/10.3390/en14206443

AMA Style

Variny M, Jediná D, Furda P. Comment on Hamayun et al. Evaluation of Two-Column Air Separation Processes Based on Exergy Analysis. Energies 2020, 13, 6361. Energies. 2021; 14(20):6443. https://doi.org/10.3390/en14206443

Chicago/Turabian Style

Variny, Miroslav, Dominika Jediná, and Patrik Furda. 2021. "Comment on Hamayun et al. Evaluation of Two-Column Air Separation Processes Based on Exergy Analysis. Energies 2020, 13, 6361" Energies 14, no. 20: 6443. https://doi.org/10.3390/en14206443

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop