Next Article in Journal
Improved Catalytic Transfer Hydrogenation of Levulinate Esters with Alcohols over ZrO2 Catalyst
Previous Article in Journal
Integrating Diphenyl Diselenide and Its Mehg+ Detoxificant Mechanism on a Conceptual DFT Framework
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Proceeding Paper

Removal of Hydrogen Sulfide (H2S) Using MOFs: A Review of the Latest Developments †

by
Amvrosios G. Georgiadis
1,
Nikolaos D. Charisiou
1,
Ioannis V. Yentekakis
2 and
Maria A. Goula
1,*
1
Laboratory of Alternative Fuels and Environmental Catalysis (LAFEC), Department of Chemical Engineering, University of Western Macedonia, GR-50100 Koila, Greece
2
School of Environmental Engineering, Laboratory of Physical Chemistry & Chemical Processes, Technical University of Crete, GR-73100 Chania, Greece
*
Author to whom correspondence should be addressed.
Presented at the 1st International Electronic Conference on Catalysis Sciences, 10–30 November 2020; Available online: https://eccs2020.sciforum.net.
Published: 9 November 2020
(This article belongs to the Proceedings of The 1st International Electronic Conference on Catalysis Sciences)

Abstract

:
The removal of hydrogen sulfide (H2S) from gas streams with varying overall pressure and H2S concentrations is a long-standing challenge faced by the oil and gas industries. The present work focuses on H2S capture using metal–organic frameworks (MOFs), in an effort to shed light on their potential as adsorbents in the field of gas storage and separation. MOFs hold great promise as they make possible the design of structures from organic and inorganic units, but also, they have provided an answer to a long-time challenging issue, i.e., how to design extended structures of materials. Moreover, the functionalization of the MOF’s surface can result in increased H2S uptake. For example, the insertion of 1% of a fluorinated linker in MIL-101(Cr)-4F(1%) allows for enhanced H2S capture. Although noticeable efforts have been made in studying the adsorption capacity of H2S using MOFs, there is a clear need for gaining a deeper understanding in terms of their thermal conductivities and specific heats in order to design more stable adsorption beds, experiencing high exothermicity. Simply put, the exothermic nature of adsorption means that sharp rises in temperature can negatively affect the bed stability in the absence of sufficient heat transfer. The work presented herein provides a detailed discussion by thoroughly combining the existing literature on new developments in MOFs for H2S removal, and tries to provide insight into new areas for further research.

1. Introduction

The removal of hydrogen sulfide (H2S), released from different industrial sources, is a matter of great importance as it can cause corrosion and environmental damage even at low concentrations. This work thoroughly focuses on H2S capture using a relatively new type of material, namely, metal–organic frameworks (MOFs), with a view to shed light on their desulfurization performance (i.e., principally via adsorption) [1]. Crystalline MOFs are formed by reticular synthesis, resulting in strong bonds between organic and inorganic units. The proper selection of MOF constituents can lead to crystals of ultrahigh porosity and high chemical, thermal, and mechanical stability. These properties allow the interior of MOFs to be chemically modified for the use in the field of gas storage and separation, among other applications. The ability to expand their metrics without changing the underlying topology and the precision generally exercised in their chemical modulation has not been achieved in other materials. The work presented herein discusses the existing literature on new developments in MOFs for H2S removal, opening new avenues for further research in terms of desulfurization processes [2].

2. H2S Capture via Materials of the Institute Lavoisier (MILs)

Hamon et al. [3] pioneered the investigation of H2S adsorption at room temperature by using different MIL-series MOFs, including MIL-47(V), MIL-53(Al, Cr, Fe), MIL-100(Cr), and MIL-101(Cr). The authors observed that larger-pore MOFs such as MIL-100 (16.7 mmol g−1) and MIL-101 (38.4 mmol g−1) exhibited higher H2S uptake in comparison to smaller-pore MOFs such as MIL-47 (14.6 mmol g−1) and MIL-53 (Al, Cr and Fe, 13.1 mmol g−1, 11.8 mmol g−1, and 8.5 mmol g−1, respectively). However, large-pore MIL-100 and MIL-101 MOFs demonstrated irreversibility, which was due either to the strong interaction of H2S with the framework, or structural collapse after H2S exposure. Soon thereafter, these findings were confirmed by the same group using a combination of IR measurements and modeling [4].
H2S capture using MIL-53(Al), in both powder and pellet form, was also examined by Heymans et al. [5] who carried out a joint experimental/theoretical approach. Their focus was on the synchronous removal of H2S and CO2 from biogas streams. The results showed that MOF-53(Al) was fully regenerable at moderate temperatures (200 °C), indicating that no chemisorption took place. It was also reported that the powdered form of MOF-53(Al) exhibited a higher desulfurization performance in comparison to that of the pelleted form, probably due to its increased specific surface area (SSA) and pore volume.
MIL-68(Al) was probed at high H2S pressures up to 12 bar at room temperature by Yang et al. [6] using both experimental and theoretical (Grand Canonical Monte Carlo; GCMC) approaches. Given the results obtained, one can conclude that the triangular pores of this MOF were locked by some remaining organic or solvent molecules, because of the incomplete activation of the material. This partially activated sorbent was proven to be fully regenerable for at least five consecutive sulfidation cycles. This notwithstanding, it needs further elucidation whether the MIL-68(Al) can resist corrosiveness following H2S exposure if fully activated.
Vaesen et al. [7] tested the adsorption performance of the amino-functionalized titanium terephthalate MIL-125(Ti)-NH2 towards its parent MIL-125(Ti) analogue, for the simultaneous removal of CO2 and H2S from biogas and natural gas, applying a joint experimental/modeling approach. The pure-component adsorption runs at a low temperature (30 °C), and low pressures resulted in decreased H2S capture for both MOFs. A key finding was that compared to other sorbents, such as 13X zeolites, these materials exhibited lower H2S adsorption enthalpies, and thus a lower energy footprint for recycling the sorbent.
Recently, Díaz-Ramírez et al. [8] reported the partial functionalization of MIL-101(Cr) with fluorine using 2,3,5,6-tetrafluoro-1,4-benzenedicarboxylate (BDC-4F). The authors aimed to investigate the adsorption performance of MIL-101(Cr)-4F(1%) at a low temperature (30 °C) and 15% of H2S volume. The results showed that this MOF outperformed other mesoporous MOFs mentioned in the literature. However, a serious downside was that H2S exposure partially led to structural degradation.

3. H2S Capture via HKUST-1 (Hong King University of Science and Technology)

Petit et al. [9] prepared HKUST-1 composites with graphite oxide (GO) (5 to 46 wt. %) and reported a synergistic effect on H2S capture for the hybrid materials. The authors reported that the composite material GO/MOF (5 wt. % of GO) outperformed both GO and HKUST-1 sorbents, with an H2S adsorption capacity of 199 mg g−1. The enhanced desulfurization performance of the GO/MOF solid was ascribed to the formation of newly formed pores in its structure. H2S molecules were captured through physisorption and reactive adsorption. Nevertheless, it was claimed that HKUST-1 suffers from structural collapse due to H2S molecules that strongly bind to the unsaturated copper centers of the MOF, resulting in the formation of CuS.
Pokhrel et al. [10] also studied the H2S adsorption on HKUST-1 and HKUST-1/GO. The authors claimed that GO did not enhance the adsorption of H2S molecules, but it was the presence of well dispersed crystals of the MOF that promoted the H2S uptake. It was also shown that both physical and reactive adsorption took place, due to the unsaturated Cu sites in the MOF structure, which interact with H2S. Regardless, since physical adsorption predominates, an increasing temperature resulted in favorable kinetics but a reduced H2S uptake. In the presence of water, the stability of both materials presented gradual degradation, indicating that chemisorption occurred.
An interesting theoretical study was carried out by Watanabe et al. [11] who tried to calculate the binding energies of different molecules, namely, H2S, H2O, CO, NO, pyridine, C2H2, and NH3, using HKUST-1. The results showed that H2S exhibited a binding strength of 0.49 eV on Cu dimers, quite close to that of H2O, which demonstrated a large affinity for the metal center of HKUST-1.
In general, the results obtained from the theoretical studies differ from the ones adopted from the experimental results. Modeling suggests that H2S physisorption prevails on HKUST-1, thus failing to explain the conversion of H2S to CuS in the presence of moisture. Contrary to experimental approaches, theoretical studies do not consider the host–guest interactions which are required to explain the CuS formation.
All things considered, experimental studies denote that H2S molecules bind stronger with Cu atoms in the center of HKUST-1, thereby displacing the existing H2O molecules.

4. H2S Capture via Isoreticular Metal–Organic Frameworks (IRMOF-n)

Isoreticular MOFs (IRMOF-n, where n = 1–16) based on a skeleton of Zn-based MOF were first prepared by Eddaoudi et al. [1] To exemplify the structure of IRMOF−1 (also known as MOF-5), Figure 1 is presented; this solid has a stable cube-like structure with a regular, three dimensional cubic lattice, with BDC as the edges and Zn4O clusters as vertexes 5.
In a theoretical study by Gutiérrez-Sevillano et al. [12], the adsorption of H2S on MOF-5 was examined. The results showed a lower heat of adsorption for the material under consideration (~−15 kJ mol−1) compared to that of HKUST-1 (~−30 kJ mol−1), probably because of the wider pores of the MOF-5. Moreover, the energy of adsorption of H2O (~−22.5 kJ mol−1) on MOF-5 was higher than that of H2S (~−16.7 kJ mol−1), suggesting that the presence of moisture negatively affects the H2S uptake.
Another study was carried out by Huang et al. [13], who prepared composites of Zn-based MOF (MOF-5) and GO in the presence of glucose for H2S removal (Figure 2). The results showed that the glucose-promoted Zn-based sample exhibited increased H2S uptake at 5.25% of GO loading, reaching a maximum capacity of 130.1 mg g−1. Nevertheless, even though the loading of GO enhanced the dispersive force in the porous structure, when the GO loading surpassed the optimum value of 5.25%, it led to the crystal distortion of the MOF-5. It was also mentioned that the insertion of glucose can help maintain structural stability and prevent distortion.

5. H2S Capture via M-MOF-74

M-CPO-27, also known as M-MOF-74 [M2(2,5-dhbdc)(H2O)2], (2,5-dhbdc = 2,5-dihydroxyterephthalate M = Ni2+, Zn2+), was investigated by Allan et al. [7] because of its strong affinity for H2S. It was reported that the H2S uptake on Ni-MOF-74 was approximately 6.4 mmol g−1 at room temperature and relative pressures (under 5 kPa). The highest H2S removal of 12 mmol g−1 was attained at 100 kPa and 25 °C. Howbeit, after regeneration, the H2S adsorption capacity was reduced in the second run, corroborating the irreversibility of H2S binding on the Ni sites.
Chavan et al. [14] also studied H2S removal (relative pressures, 10 mbar) on Ni-MOF-74 and reported the formation of H2S adducts on almost 80% of the Ni sites. It was also claimed that Ni-MOF-74 had a reversible behavior upon thermal activation at 200 °C for 12 h, and that after desorption, there was an increase in H2S adsorption capacity probably because of the additional active sites produced by means of heat treatment.

6. H2S Capture via Universitetet I Oslo MOF (UiO-66)

UiO-66 (Universitetet I Oslo) is an MOF built of [Zr6O4(OH)4] clusters (octahedra) that are 12-fold connected with adjacent octahedra through BDC struts (linkers), leading to a highly face-centered cubic structure [15].
Li et al. [16] performed a theoretical study to probe the adsorption performance of the pristine UiO-66(Zr) and its functionalized derivatives in removing sulfur from binary gas mixtures. UiO-66-(COOH)2 and UiO-66-COOH displayed the highest H2S uptake compared to that of the other tested solids, probably owing to their higher adsorption isosteric heats. The isosteric heat of adsorption under infinite dilution and radial distribution functions suggests that the hydrophilic groups and polar H2S molecules strongly interact with one another, favoring H2S removal.
Huang et al. [17] studied the antagonistic adsorption between CO2 and H2S by synthesizing core–shell-structure H2S-imprinted polymers (PMo12@UiO-66@H2S-MIPs) based on the surface of UiO-66 modified by phosphomolybdic acid hydrate. At the outset, it was mentioned that the use of H2O as a substitution template for H2S can surmount the limitations associated with H2S molecules, such as their toxic and instable nature. It was reported that PMo12@UiO-66@H2S-MIPs presented an increased H2S adsorption capacity (24.05 mg g−1) compared to that of the carrier PMo12@UiO-66, suggesting that the capacity of the latter in capturing H2S was further improved by the H2S-imprinted polymers. In addition, PMo12@UiO-66@H2S-MIPs exhibited a decent H2S adsorption capacity at room temperature and in the presence of water, while it successfully separated H2S/CO2 mixtures.

7. H2S Capture via Zeolitic Imidazolate Frameworks (ZIFs)

It is widely known that MOFs exhibit hydrothermal stability. For example, MOF-5′s structure irreversibly collapses after only 10 min of H2O exposure, even at mild conditions (low pressure and temperature) [18]. Given that MOFs have hydrophilic properties, they strongly interact with H2O. That said, even small amounts of moisture can disintegrate the coordination bonds, resulting in framework degradation. Another downside that is associated with the hydrophilic properties of MOFs is that the access of hydrophobic organic substrates is hampered, compromising the catalytic activity of some reactions. In this regard, many scientists trying to take advantage of the stability of zeolites, combined with the diverse structures and the MOFs tailorability in terms of chemical functionality, synthesized zeolitic imidazolate frameworks (ZIFs), which are classified as an MOF subclass.
ZIFs are zeolite-like structures which are built of transition metal ions that replace aluminum or silica atoms and maintain the topology of a zeolitic material. Furthermore, the organic ligands displace the oxygen atoms in the lattice of the zeolite. For instance, ZIF-8 was thoroughly examined owing to its high thermal stability (up to 550 °C), high surface area (1630 m2 g−1) and notable chemical resistance to boiling organic solvents and alkaline H2O [19].

8. Conclusions

H2S removal using MOFs can be limited by the formation of strong and often irreversible bonds. To avoid this issue, one can regulate the host–guest binding interaction between MOFs and H2S.
Reversibility after H2S sulfidation can be achieved through noncovalent bonding between functionalized ligands H2S molecules. However, the need to elucidate further the preferred H2S adsorption sites arises in order to optimize this kind of H2S separation.
Moreover, a deeper understanding of the structural characteristics of MOFs is key. For example, the structure of MOFs with open metal sites (i.e., HKUST-1, IRMOF-3 and MIL-53(Fe)) degrades when challenged with the toxic H2S, generating metal sulfides. A solution to this shortcoming may be the use of MOF composites, such as MOF/GO, wherein graphene oxide is used as a support. However, these composite materials suffer from poor H2S uptake.
Conversely, mild interactions between H2S and the open metal sites can promote the adsorption of H2S molecules without breaking the material’s structure, offering the opportunity of reversible adsorption processes.
In addition, the functionalization of the MOFs’ surface can lead to increased H2S adsorption capacities, as in the case of MIL−101(Cr)-4F(1%), which exhibited noticeable H2S uptake at low temperatures and pressures.
At this point, it is worth mentioning the lack of studies in the literature in terms of MOFs’ thermal conductivities. The designing of adsorption beds with high stability when experiencing high exothermicity is critical.
Further research should focus on the benefits provided by reticular chemistry for the developments of stable porous MOFs, suitable for sweetening applications.
Moreover, several theoretical studies studied the mechanism of H2S capture in MOFs in the presence of moisture. However, the results obtained are somehow misleading, as they contradict the ones obtained from the experimental H2S adsorption tests.
Finally, even though significant improvements have been made in terms of MOFs’ structural characteristics in the past two decades, the development of MOF structures with high H2S selectivity, higher H2S adsorption capacities, regenerability, long-term stability, and lower cost remains a challenge to be addressed in order to reach industrialization standards.

Author Contributions

Conceptualization, A.G.G.; data curation, A.G.G.; formal analysis, A.G.G.; funding acquisition, I.V.Y. and M.A.G.; investigation and methodology, M.A.G.; project administration, M.A.G. and I.V.Y.; project coordination, I.V.Y.; resources, M.A.G.; supervision, N.D.C.; writing—original draft, A.G.G. and N.D.C.; writing—review and editing, N.D.C., I.V.Y., and M.A.G. All authors have read and agreed to the published version of the manuscript.

Funding

This research has been co-financed by the European Union and Greek national funds through the operational program Competitiveness, Entrepreneurship, and Innovation, under the call Research-Create-Innovate (Project code: T1EDK-00782).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Eddaoudi, M.; Kim, J.; Rosi, N.; Vodak, D.; Wachter, J.; O’Keeffe, M.; Yaghi, O.M. Systematic Design of Pore Size and Functionality in Isoreticular MOFs and Their Application in Methane Storage. Science 2002, 295, 469–472. [Google Scholar] [CrossRef] [PubMed]
  2. Zárate, J.A.; Sánchez-González, E.; Jurado-Vázquez, T.; Gutiérrez-Alejandre, A.; González-Zamora, E.; Castillo, I.; Maurin, G.; Ibarra, I.A. Outstanding Reversible H2S Capture by an Al(III)-Based MOF. Chem. Commun. 2019, 55, 3049–3052. [Google Scholar] [CrossRef] [PubMed]
  3. Hamon, L.; Serre, C.; Devic, T.; Loiseau, T.; Millange, F.; Férey, G.; Weireld, G.D. Comparative Study of Hydrogen Sulfide Adsorption in the MIL-53(Al, Cr, Fe), MIL-47(V), MIL-100(Cr), and MIL-101(Cr) Metal–organic Frameworks at Room Temperature. J. Am. Chem. Soc. 2009, 131, 8775–8777. [Google Scholar] [CrossRef] [PubMed]
  4. Hamon, L.; Leclerc, H.; Ghoufi, A.; Oliviero, L.; Travert, A.; Lavalley, J.-C.; Devic, T.; Serre, C.; Férey, G.; De Weireld, G.; et al. Molecular Insight into the Adsorption of H2S in the Flexible MIL-53(Cr) and Rigid MIL-47(V) MOFs: Infrared Spectroscopy Combined to Molecular Simulations. J. Phys. Chem. C 2011, 115, 2047–2056. [Google Scholar] [CrossRef]
  5. Heymans, N.; Vaesen, S.; De Weireld, G. A complete procedure for acidic gas separation by adsorption on MIL-53 (Al). Microporous Mesoporous Mater. 2012, 154, 93–99. [Google Scholar] [CrossRef]
  6. Zhang, H.-Y.; Yang, C.; Geng, Q.; Fan, H.-L.; Wang, B.-J.; Wu, M.-M.; Tian, Z. Adsorption of hydrogen sulfide by amine-functionalized metal organic framework (MOF-199): An experimental and simulation study. Appl. Surf. Sci. 2019, 497, 143815. [Google Scholar] [CrossRef]
  7. Allan, P.K.; Wheatley, P.S.; Aldous, D.; Mohideen, M.I.; Tang, C.; Hriljac, J.A.; Megson, I.L.; Chapman, K.W.; De Weireld, G.; Vaesen, S.; et al. Metal–organic Frameworks for the Storage and Delivery of Biologically Active Hydrogen Sulfide. Dalton Trans. 2012, 41, 4060–4066. [Google Scholar] [CrossRef] [PubMed]
  8. Díaz-Ramírez, M.L.; Sánchez-González, E.; Álvarez, J.R.; González-Martínez, G.A.; Horike, S.; Kadota, K.; Sumida, K.; González-Zamora, E.; Springuel-Huet, M.A.; Gutiérrez-Alejandre, A.; et al. Partially Fluorinated MIL-101(Cr): From a Minuscule Structure Modification to a Huge Chemical Environment Transformation Inspected by 129Xe NMR. J. Mater. Chem. A 2019, 7, 15101–15112. [Google Scholar] [CrossRef]
  9. Petit, C.; Mendoza, B.; Bandosz, T.J. Hydrogen Sulfide Adsorption on MOFs and MOF/Graphite Oxide Composites. Chem. Phys. Chem. 2010, 11, 3678–3684. [Google Scholar] [CrossRef] [PubMed]
  10. Pokhler, J.; Bhoria, N.; Wu, C.; Reddy, K.S.K.; Margetis, H.; Anastasiou, S.; George, G.; Mittal, V.; Romanos, G.; Karonis, D.; et al. Cu- and Zr-Based Metal Organic Frameworks and Their Composites with Graphene Oxide for Capture of Acid Gases at Ambient Temperature. J. Solid State Chem. 2018, 266, 233–243. [Google Scholar]
  11. Watanabe, T.; Sholl, D.S. Molecular Chemisorption on Open Metal Sites in Cu3(Benzenetricarboxylate)2: A Spatially Periodic Density Functional Theory Study. J. Chem. Phys. 2010, 133, 094509. [Google Scholar] [CrossRef] [PubMed]
  12. Gutiérrez-Sevillano, J.J.; Martín-Calvo, A.; Dubbeldam, D.; Calero, S.; Hamad, S. Adsorption of Hydrogen Sulphide on Metal–organic Frameworks. RSC Adv. 2013, 3, 14737–14749. [Google Scholar] [CrossRef]
  13. Huang, B.L.; Ni, Z.; Millward, A.; McGaughey, A.J.H.; Uher, C.; Kaviany, M.; Yaghi, O. Thermal conductivity of a metal–organic framework (MOF-5): Part II. Measurement. Int. J. Heat Mass Trans. 2007, 50, 405–411. [Google Scholar] [CrossRef]
  14. Chavan, S.; Bonino, F.; Valenzano, L.; Civalleri, B.; Lamberti, C.; Acerbi, N.; Cavka, J.H.; Leistner, M.; Bordiga, S. Fundamental Aspects of H2S Adsorption on CPO-27-Ni. J. Phys. Chem. C 2013, 117, 15615–15622. [Google Scholar] [CrossRef]
  15. Winarta, J.; Shan, B.; McIntyre, S.M.; Ye, L.; Wang, C.; Liu, J.; Mu, B. A Decade of UiO-66 Research: A Historic Review of Dynamic Structure, Synthesis Mechanisms, and Characterization Techniques of an Archetypal Metal–Organic Framework. Cryst. Growth Des. 2020, 20, 1347–1362. [Google Scholar] [CrossRef]
  16. Li, Z.; Liao, F.; Jiang, F.; Liu, B.; Ban, S.; Chen, G.; Sun, C.; Xiao, P.; Sun, Y. Capture of H2S and SO2 from trace sulfur containing gas mixture by functionalized UiO-66(Zr) materials: A molecular simulation study. Fluid Phase Equilibria 2016, 427, 259–267. [Google Scholar] [CrossRef]
  17. Huang, Z.; Liu, G.; Kang, F. Glucose-Promoted Zn-Based Metal–organic Framework/Graphene Oxide Composites for Hydrogen Sulfide Removal. ACS Appl. Mater. Interfaces 2012, 4, 4942–4947. [Google Scholar] [CrossRef]
  18. Kaye, S.S.; Dailly, A.; Yaghi, O.M.; Long, J.R. Impact of Preparation and Handling on the Hydrogen Storage Properties of Zn4O(1,4-benzenedicarboxylate)3(MOF-5). J. Am. Chem. Soc. 2007, 129, 14176–14177. [Google Scholar] [CrossRef]
  19. Park, K.S.; Ni, Z.; Côté, A.P.; Choi, J.Y.; Huang, R.; Uribe-Romo, F.J.; Chae, H.K.; O’Keeffe, M.; Yaghi, O.M. Exceptional Chemical and Thermal Stability of Zeolitic Imidazolate Frameworks. Proc. Natl. Acad. Sci. USA 2006, 103, 10186–10191. [Google Scholar] [CrossRef]
Figure 1. The cubic structure of metal-organic framework 5 (MOF-5). The lattice constant at 27 °C is 25.85 Å. The diameter is 7.16 Å. Reproduced with permission from Ref. [1]. Copyright 2007 International Journal of Heat and Mass Transfer.
Figure 1. The cubic structure of metal-organic framework 5 (MOF-5). The lattice constant at 27 °C is 25.85 Å. The diameter is 7.16 Å. Reproduced with permission from Ref. [1]. Copyright 2007 International Journal of Heat and Mass Transfer.
Chemproc 02 00027 g001
Figure 2. Schematic view of the glucose-promoted MOF-5/GO structure unit: (A) GO layer, (B) MOF-5, and (C) glucose polymer. Reproduced with permission from ref. [13]. Copyright 2012 Applied Materials & Interfaces.
Figure 2. Schematic view of the glucose-promoted MOF-5/GO structure unit: (A) GO layer, (B) MOF-5, and (C) glucose polymer. Reproduced with permission from ref. [13]. Copyright 2012 Applied Materials & Interfaces.
Chemproc 02 00027 g002
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Georgiadis, A.G.; Charisiou, N.D.; Yentekakis, I.V.; Goula, M.A. Removal of Hydrogen Sulfide (H2S) Using MOFs: A Review of the Latest Developments. Chem. Proc. 2020, 2, 27. https://0-doi-org.brum.beds.ac.uk/10.3390/ECCS2020-07586

AMA Style

Georgiadis AG, Charisiou ND, Yentekakis IV, Goula MA. Removal of Hydrogen Sulfide (H2S) Using MOFs: A Review of the Latest Developments. Chemistry Proceedings. 2020; 2(1):27. https://0-doi-org.brum.beds.ac.uk/10.3390/ECCS2020-07586

Chicago/Turabian Style

Georgiadis, Amvrosios G., Nikolaos D. Charisiou, Ioannis V. Yentekakis, and Maria A. Goula. 2020. "Removal of Hydrogen Sulfide (H2S) Using MOFs: A Review of the Latest Developments" Chemistry Proceedings 2, no. 1: 27. https://0-doi-org.brum.beds.ac.uk/10.3390/ECCS2020-07586

Article Metrics

Back to TopTop