Next Article in Journal
Assessment of the Current Potential of Hydropower for Water Damming in Poland in the Context of Energy Transformation
Next Article in Special Issue
Smart Grid, Demand Response and Optimization: A Critical Review of Computational Methods
Previous Article in Journal
Collection and Utilization of Indoor Environmental Quality Information Using Affordable Image Sensing Technology
Previous Article in Special Issue
A Review on the Dispersion and Distribution Characteristics of Pollutants in Street Canyons and Improvement Measures
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Olive Mill Wastewater Valorization through Steam Reforming Using Multifunctional Reactors: Challenges of the Process Intensification

by
Cláudio Rocha
1,
Miguel Angel Soria
2 and
Luís M. Madeira
2,*
1
LSRE-LCM—Laboratory of Separation and Reaction Engineering-Laboratory of Catalysis and Materials, Associate Laboratory, Department of Chemical Engineering, Faculty of Engineering, University of Porto, Rua Dr. Roberto Frias s/n, 4200-465 Porto, Portugal
2
LEPABE—Laboratory for Process Engineering, Environment, Biotechnology and Energy, Department of Chemical Engineering, Faculty of Engineering, University of Porto, Rua Dr. Roberto Frias s/n, 4200-465 Porto, Portugal
*
Author to whom correspondence should be addressed.
Energies 2022, 15(3), 920; https://doi.org/10.3390/en15030920
Submission received: 1 December 2021 / Revised: 10 January 2022 / Accepted: 13 January 2022 / Published: 27 January 2022
(This article belongs to the Collection Review Papers in Energy and Environment)

Abstract

:
Olive oil mill wastewater (OMW) is a polluting stream derived from the production of olive oil and is a source of environmental pollution; this is relevant in many countries around the world, but particularly in all the Mediterranean region where major producers are located. In this effluent, several pollutants are present—namely, sugars, fatty acids, and polyphenols, among others. Nowadays, to reduce the pollutant load, several treatment techniques are applied, but these technologies have numerous cost and efficiency problems. For this reason, the steam reforming of the OMW (OMWSR) presents as a good alternative, because this process decreases the pollutant load of the OMW and simultaneously valorizes the waste with the production of green H2, which is consistent with the perspective of the circular economy. Currently, the OMWSR is an innovative treatment alternative in the scientific field and with high potential. In the last few years, some groups have studied the OMWSR and used innovative reactor configurations, aiming to improve the process’ effectiveness. In this review, the OMW treatment/valorization processes, the last developments on catalysis for OMWSR (or steam reforming of similar species present in the effluent), as well as the last advances on OMWSR performed in multi-functional reactors are addressed.

Index
Abstract
1
1. Introduction3
 1.1. OMW Composition8
 1.2. OMW Treatment/Valorization9
 1.3. Olive Mill Wastewater Steam Reforming
12
2. Omwsr Catalysts15
 2.1. Catalysts Used for Steam Reforming of OMW15
 2.2. Catalysts Used for Steam Reforming of Model Molecules of OMW19
   2.2.1. Effect of the Active Phase19
   2.2.2. Effect of the Support20
   2.2.3. Effect of Promoters21
 2.3. Summary
27
3. Multifunctional Reactors28
 3.1. Sorption-Enhanced Reactor28
   3.1.1. CO2 Sorbents30
   3.1.2. Hydrotalcites34
   3.1.3. OMWSR in Sorption-Enhanced Reactors39
   3.1.4. Summary40
 3.2. Membrane Reactor40
   3.2.1. Hydrogen Perm-Selective Membranes41
   3.2.2. OMWSR in Membrane Reactors46
   3.2.3. Summary49
 3.3. Sorption-Enhanced Membrane Reactor49
   OMWSR in Sorption-Enhanced Membrane Reactors
50
4. Conclusions
51
Notation and Glossary
52
References54

1. Introduction

The olive oil industry is very relevant in many countries around the world—particularly in the Mediterranean countries (cf. Figure 1), both economically and traditionally. Because of its excellent nutritional properties, olive oil contributes to cholesterol control and promotes cardiovascular health. This significance is explained in numbers: about 9 million hectares of olive trees have been cultivated worldwide during the 2019/2020 crop year [1]. For the 2019/2020 crop year, it was estimated that 70% of the production of olive oil around the world was performed in European Union countries, with the Iberian Peninsula (Portugal and Spain) being a leader zone (76% of the European Union’s total)—see Figure 1 [1,2]. Nevertheless, there are other countries with high levels of olive oil production: Tunisia, Morocco, Syria, and the USA. Besides that, other countries—such as Australia, Argentina and South Africa—are also employing this type of crop [3].
In this way, it is observed that all these regions are significantly affected by the pollution resulting from olive oil production. The process yields massive amounts of liquid (olive oil mill wastewater—OMW) and solid waste, which have raised severe disposal difficulties for the producers. This agroindustry represents a huge environmental problem, where big quantities of effluents are produced in a small period. This environmental concern is emphasized by the large amount of word olive oil production/consumption (close to 3 million metric tons per year produced/consumed—this value has remained nearly constant in the last 10 years) [2,4,5]—see Figure 2.
The aqueous residues related to this agro-industrial activity are produced using either the traditional pressing or the centrifugation methods (three- or two-phase systems) because, in all of them, water is used to generate the olive oil from the olive [1,6]. In Figure 3, a scheme of these processes of olive oil production can be seen [7,8].
The discontinuous pressing procedure (traditional method) is the oldest and most well-known method for processing olives. In the traditional pressing (nowadays adopted only in small mills/companies), the ground olives are pressed and then the liquid mixture is rested in a sequence of decanters to separate the olive oil. To separate the oil from the other phases, a small amount of water is added. This traditional process also produces a solid fraction denominated by olive pomace (or olive cake) constituted by water, olive stone, and olive pulp. This traditional method presents several advantages such as low-priced equipment and technological simplicity. Nevertheless, it also has handicaps, such as the fact of running discontinuity and high hand-labor costs [7]. Still, because it adds water in several steps of the production process (previous to the decantation step), a large quantity of OMW (approximately 40–60 L/100 kg olives) is produced [7]. The OMW generated has a higher chemical oxygen demand (COD) compared with the OMW produced by the other methods [9]. Until 40 years ago, this was the only process used, although financial problems pushed olive mills to gradually abandon the traditional process [10,11].
In the centrifugation systems, crushed olive fruits are fed into a two-phase or three-phase decanter centrifuge and then the impure olive oil is centrifuged [12]. These processes are based on the differences in the density of the several components obtained (solids, olive oil, and water) and can produce three distinct fractions: pomace, olive oil, and OMW—see Figure 3.
Using the three-phase process (route a) in Figure 3, the olive paste is divided into three phases in the horizontal centrifugation (water added in this step): olive oil, OMW, and pomace [7,8]. Then, the liquid fraction generated (olive oil + OMW) is centrifuged (vertical centrifugation) with additional water to recover a high-quality olive oil, producing a high quantity of OMW—see Figure 3. Since the continuous three-phase centrifugation method adds water in the two centrifugation steps, a high volume of OMW is produced (about 100–120 L/100 kg olives—1.7–3 times more compared to the traditional method) [8]. Nevertheless, this three-phase system has various advantages, such as better quality of the olive oil and complete automation. Nonetheless, this system also presents several disadvantages, such as a higher level of water and energy consumption, higher volume of OMW produced (as previously mentioned), and more expensive installations [3].
To minimize the impact of OMW, the two-phase centrifugation process (route b) in Figure 3 was developed during the 1990s [7]. Using this two-phase process, the olive paste is divided only into two phases in the horizontal centrifugation (no water added in this step): wet pomace and olive oil, the first known as two-phase olive mill waste (TPOMW), a combination of pomace and OMW—see Figure 3 [7]. Then, the olive oil generated is washed (vertical centrifugation—oil washing) with a very small amount of water, producing only a small OMW stream—see Figure 3 [8]. Despite the lower volume of OMW produced in comparison with the three-phase process, the resulting TPOMW (10 L/100 kg olives) is harder to manage and treat, because its pollutant load is more concentrated [7]. Nevertheless, some quantity of OMW is drained from the TPOMW in the storage tanks and is collected using draining valves. Besides that, it is possible to extract some oil from TPOMW through several techniques (e.g., using solvents), valorizing, in this way, this by-product [7]. In this way, the two-phase system, besides presenting the same advantages of the three-phase system, also produces a lower quantity of OMW (about 8.5–11 L/100 kg olives) and reduces water consumption; thus, this centrifugation system has been called “ecological” [3,13]. Nevertheless, like the three-phase system, the two-phase system needs expensive installations.
Figure 4 illustrates a scheme quantifying the OMW production (for all the extraction processes used in the countries of the Mediterranean area) in relation to the quantities of olives and extractive water used, and olive oil/olive husk (pomace) produced; one can see that big ratios between generated OMW and olive oil are obtained [14]. Moreover, it is estimated that 30 Mm3 of OMW per year are produced around the world [6]. The OMW results from both the water present in the olives and the water used in the production process of the olive oil, especially with the three-phase centrifugation system and pressing method [1,6].
These big amounts of OMW cause large environmental problems [15,16,17,18]. It is reported that such residues have an effect on soil/air quality and on the aquatic ecosystems due to the deposition of toxic streams directly in the receiving bodies. Some of the main negative impacts are the increase of the organic charge concentration and the existence of phytotoxicity in the water systems. This happens due to the high values of COD (60–200 g·L−1), total organic carbon (TOC—> 9 g·L−1), and biochemical oxygen demand (BOD—14-100 g·L−1) present in the OMW [6]. Pollution from OWM is expected to be 200–400 times higher than from urban wastewater [1,19,20]. The uncontrolled disposal of OMW represents economic, social, and environmental problems that must be resolved.

1.1. OMW Composition

The composition of OMW is extremely variable and suggests that minor or major compounds can be present, depending on numerous factors, which include the age of the olive tree, region of cultivation, maturation level of the olive, weather conditions that the olive was subjected to in the ripening process, treatment of the tree, and method of extracting the oil [21]. El-Abbassi et al. [22] showed that OMW resulting from the traditional method had a higher phenolic content compared to that obtained from the three-phase centrifugal system. The investigation of these effects in the OMW composition is very important, in order to develop sustainable OMW management strategies.
In general, OMW consists of solid matter and liquid, with a slightly acidic pH. The latter is composed of water, polyphenols, fatty acids, and sugars, among other organic and inorganic species [6,23]. The content of water in this stream is usually between 80–95 wt.% [6,24,25,26]. In Table 1, it is possible to see a typical physicochemical characterization of the OMW [23].
The compounds most reported are the following: caffeic acid, vanillic acid, tyrosol, cinnamic acid, p-coumaric acid, D-arabinose, D-galactose, D-galacturonic acid, acetic acid, syringic acid, gallic acid, protocatechuic acid, guaiacol, phenol, phenethyl alcohol, and benzyl alcohol [1,6,7,14,21,22,26,27,28,29,30,31,32,33,34,35,36,37]. However, it is important to mention that more than 30 polyphenols and sugars have been noticed in OMW [38]. In Table 2 an example of OMW composition on a dry basis is represented [36].

1.2. OMW Treatment/Valorization

As the substances present in OMW are very toxic, the effluent needs to be correctly treated. A great variety of treatment/valorization processes (see Figure 5) have been investigated in the past years to reduce OMW pollutant load or to allow its disposal, but research efforts turn nowadays toward other aspects. Still, those methods of OMW load reduction are not often used due to technical and economic reasons. The trivial management of the OMW consists of the direct discharge into sewer systems and rivers, disposal in evaporation lagoons or dilution with other effluents [4,39]. In addition, the latter practice requires large areas and produces a black foul-smelling sludge that is difficult to handle, with inherent pollutant infiltration into ground water and insect proliferation. Land disposal is also a common practice considering OMW as a fertilizer, if it is well managed. Nevertheless, to avoid the contamination of the soil, land disposal must be quantitatively controlled by releasing small amounts of wastewater. Thus, nowadays, an organized and controlled release of OMW into the environment is still permitted, according to specific regulations of the different countries.
Therefore, solutions for adding value to OMW or for extracting valuable compounds from it before treatment/discharge are required. The OMW valorization/treatment methods can be organized into several general categories (see Figure 5). Among all the existing sub-categories, the following stand out:
  • Thermal Processes [4]:
    Distillation and Evaporation Processes: these processes concentrate the inorganic and organic contents of the OMW by water evaporation. These processes have high operation costs and the organic content must be treated afterwards;
    Combustion: this process is very efficient but requires a large amount of energy.
  • Membrane Processes [4,26,40,41,42]: these processes separate the organic and inorganic contents of the OMW by the utilization of membranes. However, the cost of the membrane and its use might be high, particularly if high pressures are required, and the process has operation problems (membrane fouling, concentration polarization, etc.). There are several types of membrane-based processes: ultrafiltration, nanofiltration, reverse osmosis, and microfiltration.
  • Physicochemical Processes [43]: these processes consist of physical and chemical reactions involving molecules present in the OMW composition. Sometimes, they are conjugated to maximize efficiency. The following processes are included in this category:
    Electrochemical Oxidation [44,45]; Ozonation [46,47]; Adsorption [48,49,50]; Ion-Exchange [51,52]; Flocculation [4,53,54]; UV Photocatalysis [45,55,56]; Wet Oxidation [57,58,59]; Fenton Oxidation [60,61,62,63,64]; Photo-Fenton [38,65]; Electro-Coagulation [66,67]; Enzymatic Catalysis [68]; Supercritical Water Gasification [69].
  • Biological Processes [70]: these processes use microorganisms to decrease the pollutant load and, simultaneously, provide products with value (for instance, formation of CH4 or H2). The following are included in this category:
    Aerobic Processes [70,71,72,73]; Anaerobic Processes [50,74,75,76].
Figure 5. General scheme of different techniques applicable for OMW treatment/valorization [77].
Figure 5. General scheme of different techniques applicable for OMW treatment/valorization [77].
Energies 15 00920 g005
Most of the methods studied aiming OMW valorization include the recovery of the polyphenols fraction owing to their interesting pharmacological properties [78,79,80,81]; besides that, OMW phenolic extracts can be used in the food industry, in order to improve the oxidative status of the products [77]. For instance, in recent years, membrane processes have been used for polyphenols recovery [26]—namely microfiltration, ultrafiltration, reverse osmosis, nanofiltration, membrane distillation, and osmotic distillation [4,40,41]. However, membrane-based processes have a range of limitations; e.g., may be limited by the OMW’s high concentration of suspended solids. Adsorption processes [48,50] have been also applied to the recovery of phenolic compounds to valorize the OMW effluent.
Concerning the treatment processes, the most commonly utilized physicochemical treatments with this purpose include electrochemical oxidation [44,45], ozonation [46,47], electro-coagulation [66], adsorption [49], and advanced oxidation processes [43,82,83] (particularly Fenton oxidation [60,61]). Since the polyphenols have strong resistance to bacteria, the physicochemical methods can be extremely important, because they can reduce the organic content to make the effluent suitable for biological processes (two different techniques combined for OMW treatment/valorization).
In general terms, several conventional valorization/treatment technologies with varying degrees of success have been proposed. However, most of them require considerable financial input and show low efficiency. In addition, they frequently focus on removing the polyphenols and end up producing a secondary waste that needs additional treatment.
Therefore, it is necessary to improve the scientific and technological solutions for the treatment and valorization of these wastes to decrease the environmental impacts and to allow sustainable use of resources, in line with international waste management practices. In this perspective, the possibility of waste valorization for green energy production has been considered [84], namely through OMW steam reforming (OMWSR) [85]. This technology would permit removing the high pollutant components present in the OMW while producing H2, therefore providing added-value. This biofuel is environmentally attractive once it is renewable and reduces the CO2 emission into the atmosphere [84]. Unlike fossil fuels, H2 burns cleanly, without producing any environmental pollutants. The OMWSR would help reducing the pollution load of the effluent as well as to economically valorize a by-product of the olive oil industry without any value so far [85]. The H2 can be applied in several technologies like in combustion engines or electrochemical cells. In this perspective, it is worth mentioning that hydrogen technologies proved extremely resilient during the COVID-19 pandemic, with their momentum staying strong in 2020 [86]. Moreover, a recent publication of the International Energy Agency reported that the global H2 demand in the Net Zero Emission Scenario (between 2020 and 2030) will increase significantly in the next years, in order to reach international climate and sustainable energy goals—see Figure 6 [86,87]. To meet this growing demand for H2, the environmentally friendly production of H2 has become an important part of the H2 market. However, this progress falls well short of what is needed in the Net Zero Emissions Scenario.

1.3. Olive Mill Wastewater Steam Reforming

The OMWSR (Equation (1)) consists of the sum of two independent reactions: decomposition of high molecular weight molecules in the presence of steam (Equation (2)) and the water–gas shift (WGS—Equation (3)) reaction. The values of Δ H r 25   ° C were taken from the literature [84].
C x H y O z + 2 x z   H 2 O 2 x z + y 2 H 2 + xCO 2 Δ H r 25   ° C   ( kJ · mol 1 )   >   0
C x H y O z + x z H 2 O x z + y 2   H 2 + xCO Δ H r 25   ° C   ( kJ · mol 1 )   >   0
CO +   H 2 O H 2 + CO 2 Δ H r 25   ° C   ( kJ · mol 1 )   =   41
The reason why the OMWSR was chosen as a method of OMW valorization is mainly due to the fact that the steam reforming is extensively used in industry, apart from the reasons given above (i.e., environmental concern and production of green H2). Besides that, syngas (H2 and CO) produced can be applied as a resource for the fabrication of valuable chemicals and fuels through Fischer–Tropsch or dimethyl ether synthesis routes [88].
The OMWSR process (in a Traditional Reactor—TR—see Figure 7) has, however, some challenges that need to be overcome and optimized to be a valid approach for the OMW valorization, namely:
  • The OMWSR is an endothermic reaction, thus needing high temperatures and intrinsically high operating costs; moreover, producing H2 by this process also yields CH4, CO2, and CO, whose production is a target of environmental alarm and limited by legislation;
  • The WGS reaction is thermodynamically limited (Equation (3));
  • This process has several side reactions, which affect both the formation and purity of H2. The main side reactions are represented by Equation (4)—methanation of carbon monoxide, Equation (5) (sum of Equation (3) and reverse of Equation (4))—steam reforming of CH4, and Equation (6) (sum of reverse of Equation (3) and reverse of Equation (4))—dry reforming of CH4;
  • Other products apart from the previous ones—such as butane, propane, acetone, methanol, benzene, ethanol, cyclopentadiene, ethylene, butanol, and acetaldehyde—could be present in the reaction medium. These species are referred by many works as possible secondary products of some reactions (e.g., Equation (7)—cracking of oxygenates) or as intermediates of the reactions involved in this process [1,27,28];
CO + 3 H 2 H 2 O + CH 4 Δ H r 25   ° C   ( kJ · mol 1 )   =   206
2 H 2 O + CH 4 CO 2 + 4 H 2 Δ H r 25   ° C   ( kJ · mol 1 )   =   165
CO 2 + CH 4 2 CO + 2 H 2 Δ H r 25   ° C   ( kJ · mol 1 )   =   247
C x H y O z C n H m O k + gas H 2 , CO 2 ,   CO , CH 4 ,   CH n + C Δ H r 25   ° C   ( kJ · mol 1 )   >   0
  • The formation of coke is a problem since it deactivates the catalysts (Equations (7) and (8)—cracking of hydrocarbons, Equation (9)—Boudouard reaction, Equation (10)—methane cracking, Equation (11)—carbon monoxide reduction, and Equation (12)—carbon dioxide reduction), thus affecting H2 yield and purity and long-term operation of the process;
CH x coke   precursors   alofins + aromatics x 2 H 2 + C Δ H r 25   ° C   ( kJ · mol 1 )   >   0
2 CO C + CO 2 Δ H r 25   ° C   ( kJ · mol 1 )   =   172
CH 4 2 H 2 + C Δ H r 25   ° C   ( kJ · mol 1 )   =   74
CO +   H 2 H 2 O + C Δ H r 25   ° C   ( kJ · mol 1 )   =   131
CO 2 + 2 H 2 2 H 2 O + C Δ H r 25   ° C   ( kJ · mol 1 )   =   90
With the aim of resolving these difficulties, new catalyst formulations must be developed. Furthermore, new reactor configurations merging the OMWSR catalytic process and H2 and/or CO2 selective removal should be addressed, due to their potential to solve the restrictions mentioned above (namely shifting in the desired direction the reversible reactions, like the WGS). In this work, some of the important progress regarding the OMWSR process (developed catalysts) and its combination with a H2 perm-selective membrane (in a membrane reactor—MR), in the perspective of process intensification, are reviewed; still, the potential of the utilization of the sorption-enhanced reactor (SER with a CO2-selective sorbent) and the sorption-enhanced membrane reactor (SEMR—combining in the same device the separation of H2 and capture of CO2) are also addressed.

2. OMWSR Catalysts

Heterogeneous catalysis applied to the OMWSR is not very developed until now. The catalysts investigated so far have been Ni-, Pd-, Pt-, and Ru-based. However, several catalysts were extensively studied for the steam reforming of individual components that are similar to the components present in the OMWSR (like acetic acid, or similar molecules). Currently, the key goal is to prepare catalysts with high catalytic activity and high stability.

2.1. Catalysts Used for Steam Reforming of OMW

Firstly, this reaction was studied by Tosti et al. [1,6,89] and by Llorca et al. [28]. Recently, Alique et al. [89] also studied the steam reforming of the OMW in an MR, using a catalyst developed and tested in a previous work of Tosti et al. [89].
In the first study, Tosti et al. [6] used a commercial Pt-based catalyst (1 wt.% Pt on Al2O3—BASF SP-01 T) in an MR (150 μm thick Pd-Ag dense tubes). The experimental tests were carried out at 450 °C and in the range of 1–5 bar, at different space velocities (Weight Hourly Space Velocity—WHSV = 2.78 × 10−3–8.33 × 10−3 mol·h−1·gcat−1. The gas chromatographic analysis of the retentate side revealed the existence of H2, CO2, and CH4, thus showing the presence of side reactions such as the hydrogenolysis and cracking of high weight molecules. Additionally, catalyst deactivation due to coke formation was detected. Then, this research group prepared a new noble metals-based catalyst [89] and compared it with the commercial Pt-based catalyst used in the previous work under similar operation conditions (using the same reactor configuration—MR). The new catalyst consists of Al2O3 pellets coated with a CeO2-ZrO2 layer. The active species (Pd, Rh, and Pt) loading was lower than 3 wt.%. CeO2-ZrO2 presents several benefits in the steam reforming reactions: oxidation reactions and WGS reaction are enhanced by the high oxygen mobility, which is also related to the decrease of the coke formation and sintering inhibition. Moreover, when these supports are employed, Rh can exhibit high catalytic stability during the reaction, low Pt loadings (<1 %) are enough for limiting deactivation phenomena and Pd can show good stability even at very low steam-to-carbon feed ratios (SCFRs). This catalyst exhibited no substantial loss of activity after 6–8 h of operation, in contrast with the commercial Pt-based catalyst whose activity reduced after 1–2 h of experimental test. When compared to the commercial material, the new catalyst exhibited higher selectivity, allowing it to significantly reduce the production of CH4 and coke; in this way, it is obtained a higher H2 production and higher durability and stability. This new catalyst was also used in the recent work of Alique et al. [90], who reported that, in a MR, higher pressures increased the permeation of H2, prevailing the membrane shift effect against the negative effect on thermodynamics of the reactions involved in the steam reforming process. In this way, an increase in H2 production was obtained as the pressure increased.
Llorca and co-workers are another research group that studied this reaction [28] in the last few years in a TR. Several catalytic honeycombs loaded with Ni or noble metals (Ru, Pt, Rh, or Pd) over La-stabilized CeO2 were prepared. The experiments were carried out at 1 bar and in the range of 600–750 °C, at different space velocities (Gas Hourly Space Velocity—GHSV = 4500–16,000 h−1). Their catalytic performance in terms of production of H2, stability and selectivity towards reforming products followed this trend: Pt/CeLa > Rh/CeLa > Ru/CeLa >> Ni/CeLa. The trend of coke formation on the surface of these catalysts was: Ru/CeLa > Rh/CeLa > Pt/CeLa, which is contrary to that of the catalytic performance, strongly suggesting that coke deposition was the cause of catalyst deactivation during OMWSR—see Figure 8. Besides, it was found that the selectivity for secondary reactions increases with time, for all the catalysts, and the catalyst with higher selectivity for H2 formation was Rh/CeLa. In addition, Ca deposition was also observed after the stability tests in all samples; seemingly, calcium is present in the OMW and its possible influence on the long-term catalytic activity must be further studied. The 1 wt.% Pt/CeLa catalyst showed a very stable distribution of products in long term runs. An increase of the Pt load up to 2 wt.% had a positive effect on the product distribution, particularly on the suppression of CH4.
Finally, Rocha et al. [85,91] tested several commercial and prepared catalysts (Cu-Zn-, Ni-, and noble metal-based) for the OMWSR process (using a synthetic OMW effluent—a complex mixture of model oxygenates, representative of a real effluent) to compare their catalytic performances. A catalytic screening study with all the materials was performed (@ 1 bar and 350/400 °C), and stability tests (@ 1 bar and 400 °C) were conducted with the most promising samples. Although there are a few materials with good catalytic performance, the commercial Rh-based catalyst stood out, exhibiting high performance (catalytic activity and stability): the H2 yield (over 9 molH2·mol−1OMW) and total TOC conversion (>98%) were high during the long-term experiment (24 h of the stability test)—see Figure 9.
Besides that, several studies regarding the steam reforming of guaiacol (component typically present in the composition of the OMW) using Ni-based catalysts were also reported [92,93]. Finally, it is important to emphasize that several works have already studied the steam reforming of effluents (e.g., bio-oil and palm oil mill) with compositions very similar to OMW (oxygenated species), using Ni and Co-based catalysts [94,95,96,97,98,99,100,101,102,103].
A summary of some results obtained with the more relevant catalysts reviewed in this section is presented in Table 3.
New catalysts must be developed for future catalytic screening and improvement. These catalysts will need to maximize the reaction conversion, decrease the coke formation, and increase the H2 selectivity under milder operating conditions.

2.2. Catalysts Used for Steam Reforming of Model Molecules of OMW

Among the materials above-mentioned, numerous catalysts (Ni-, Pt-, Ru-, Co-, and Rh-based) have been extensively studied for the steam reforming of individual components similar to the species present in OMW. Thus, in this work, three different molecules were defined as OMW model compounds: acetic acid, phenol, and toluene. The catalytic performances of the catalysts used for the steam reforming of these species are affected by the active phase, support, and promoter agents. These topics are discussed in the following sections.

2.2.1. Effect of the Active Phase

The active phase is the metal element of the catalyst that catalyzes the reaction. Numerous catalysts (Ni-, Pt-, Ru-, Co-, and Rh-based) have been extensively studied for the steam reforming of oxygenates.
The Ni-based catalysts are the most used in the steam reforming processes because of their effectiveness (they have the potential to be highly active and stable), being commercially available and relatively cheap [104,105,106,107,108]. Nevertheless, the Ni loading, support, promoter(s), and synthesis method used in the catalyst preparation have a crucial impact on both catalytic activity and price [109]. In the study of Zhang et al. [110], it was concluded that the increase of Ni loading from 10 wt.% did not significantly increase the activity of the catalysts, but enhanced the stability, and especially the resistivity towards coking in the steam reforming of acetic acid (@ 1 bar, Liquid Hourly Space Velocity—LHSV = 4.2 h−1, SCFR = 5 and 300–600 °C). The coke formed over the catalyst with the lower Ni loading was more amorphous and the coke formed over higher Ni loading was more fibrous [110]. It was verified by Borges et al. [111] that the higher the Ni loading is, the lower the temperature and the time necessary for the catalyst reduction. The catalyst with a molar ratio of Ni/Co = 3 exhibited the highest activity and the best carbon deposition resistance due to the high density of basic sites for the steam reforming of phenol. In a study by Yu et al. [112], the effect of the type of anion in the Ni precursors in the catalytic behavior of Ni/Al2O3 catalysts in steam reforming of acetic acid was studied. It was found that the NO3 and CH3COO as the anions of the Ni precursors did not cause negative effects on the catalytic performance of the catalyst.
Güell et al. [113] reported that the Pt catalyst was active for the steam reforming of oxygenates compounds, as shown in the study of de Castro et al. [114]. Pant et al. [115] reported that Ni-Co, Ni-Co/CeO2-ZrO2, and Ni/La2O3-Al2O3 catalysts can catalyze the steam reforming reaction of acetic acid. Among the three catalysts investigated, the Ni-Co catalyst was found to be more effective for this reaction. Several authors [116,117,118] found that Rh-based catalysts have a high performance in the steam reforming of oxygenates and several of these materials exhibit superior performance when compared to a commercial Ni-based sample. A series of Ni-Cu bimetallic catalysts supported on sepiolite (Nix-Cuy/SEP) were prepared by Liang et al. [119] for the steam reforming of phenol. The results show that the Ni-Cu alloys were successfully produced in bimetallic catalysts, and the addition of Cu decreased the size of the Ni particles, and improved the redox ability and metal dispersion of bimetallic materials (in this case, Cu can also be considered as a promoter of the catalyst).
The Fe2O3/SiO2−Al2O3 catalyst was used by Adnan et al. [120] for the steam reforming of toluene. The results indicate that Fe2O3/SiO2−Al2O3 is a promising catalyst (in terms of activity and stability, @ 500 °C, 3.8 bar and SCFR = 0.14) since it has a large quantity of acidic active sites. However, the study of Wang et al. [121] (steam reforming of acetic acid) reported that, in the Co-Fe catalysts, the more active component is Co rather than Fe.
Furthermore, several authors [122,123,124] show that multi-metallic catalysts show high activity and improve the H2 yield due to synergistic effects of the multiple components of the active phase.
Finally, mixed oxides such as hydrotalcite-type oxides have been reported as promising catalyst precursors for the steam reforming of these types of oxygenates [111,125].

2.2.2. Effect of the Support

In several studies, it was verified that the support has a significant effect on the performance of the catalysts. Different catalyst supports significantly influenced the reforming reaction.
A series of Ni core-shell catalysts with various shell species (SiO2, Al2O3, CeO2, and TiO2) were prepared by Pu et al. [126]. By comparing the catalytic activities of the catalysts with various shell materials, it was concluded that the Ni/Al2O3 catalyst was the most suitable material for the steam reforming of acetic acid, showing much higher catalytic activity than the other catalysts. It was also observed in the work of Li et al. [127] that Al2O3 support has a high specific area that can enhance the catalytic activity, but the acidity of Al2O3 promotes the carbon deposition, reducing the stability of the materials.
The nature of the support was also studied by Zhang et al. [128], using attapulgite (ATTP) and Al2O3 as support of Ni-based catalysts. It was verified that ATTP has a lower specific surface area and lower thermal stability than Al2O3. The stability of the Ni/Al2O3 catalyst was higher than that of the Ni/ATTP catalyst, not only due to the higher surface area of Al2O3, but also due to the type of coke formed on the surface of the Ni/Al2O3 catalysts (fibrous) instead of the amorphous coke formed on Ni/ATTP catalyst.
Chen et al. [129] studied the effect of biochar as support in Ni-based catalysts for the steam reforming of acetic acid. The results of the catalyst characterization showed that the porosity of biochar was increased significantly after activation, increasing the surface area of the materials and the dispersion of the Ni particles on the catalyst’s surface. Besides that, in the work of Wang et al. [130]—study of the steam reforming of acetic acid—it was found that the treatment of biochar support with HNO3 enhanced the quantity and the strength of the acidic active sites as well as the catalytic activity of the Ni/biochar catalysts.
Additionally, Yang et al. [131] using a NiO/MgO catalyst have concluded that the interaction between the support and active phase can prevent the material from sintering.
Finally, silicate structured mesoporous materials were reported to be less vulnerable to deactivation due to carbon formation than the traditional microporous catalyst supports. They also cause less resistance to the diffusion of reactants to the active sites [88].

2.2.3. Effect of Promoters

In several studies of the steam reforming of oxygenates, the catalysts were doped with promoters. It was reported that Cr inhibited the formation of Ni3C and can modify the metal sites forming alloys with Ni [132,133], La2O3 led to a decrease in coke formation [105,132,133,134,135,136], K and MgO improved the stability [105,117,132,133,136,137,138,139], and TiO2 decreased the conversion [132]. It was also observed that the addition of CeO2 improved the activity and stability of the catalyst [106,108,114,135,140,141,142].
The work of Li et al. [143] reported that the presence of CeO2 in the catalyst increased the support–metal interaction, restricting the growth of metal particles. Besides that, it was found that the oxygen vacancies of the CeO2 are responsible for the activation of the water to generate OH- groups, which can react with intermediate products to produce CO2 and H2. In the study of Iiada et al. [118], Ru/SrCO3-Al2O3 catalysts with a low Ru loading exhibited higher catalytic activity for steam reforming of toluene compared to a commercial Ru/Al2O3 catalyst (@ 600 °C and SCFR = 2). In this way, it was possible to see that the SrCO3 increases the performance of the catalyst. Additionally, in the study of Zhang et al. [144], it was found that the addition of KOH to Ni/Al2O3 with the lowest Ni loading significantly enhances the catalytic activity and promotes gasification of the reactive intermediates such as methyl group, carbonyl group, etc.
Choi et al. [145] studied the steam reforming of acetic acid by using catalysts modified by MgO, LaO3, Cu, and KOH. Ni/Al2O3 modified with Mg showed the best performance at low temperatures. It was possible to verify that a catalyst with a large quantity of weak basic sites and few middle/strong basic sites is necessary to improve the catalytic performance and minimize the coke formation. Finally, it was observed that the toluene conversion and H2 yields increased with the addition of alkaline-earth metals [146]. In a general way, the addition of basic oxides (e.g., MgO, CeO2, La2O3, etc.) enhanced the coking resistance, thus improving the catalytic performance. These oxides aim to enhance the redox properties of the catalyst [143,147,148,149,150], which increases the oxidation of carbon deposits. For instance, Kumar et al. [151] reported that the coke formation was also significantly suppressed by the γ-Al2O3-La2O3-CeO2 support.
Finally, Yu et al. [152] reported that the addition of Mn enhanced the specific area of the catalyst, by the formation of more pores and the addition of Fe, Mn, La, and Ce could enhance the dispersion of Ni particles in the support. Besides that, Co, K, La, Na, or Ce enhance the activity of Ni-based catalysts and K or Na addition enhanced the reduction of Ni oxides.
However, Galdamez et al. [153] concluded, in the steam reforming of bio-oil, that the addition of La2O3 in the Ni catalyst does not increase the H2 yield. It was also found that ZrO2 had a negligible effect on the performance of the catalyst [138].
A summary of the more relevant catalysts reviewed in this section for the steam reforming of individual components (or similar) of the OMW and their best catalytic performances are presented in Table 4. It is possible to verify that Ni-based catalysts are effective for these reactions (they exhibit similar catalytic performance compared to noble metal-based catalysts) and are relatively cheaper. Therefore, more detailed information about the catalytic stability and preparation method of these catalysts is reported in Table 5. It is possible to see that several Ni-based catalysts present no deactivation until the end of the experimental test, even though noble metal-based catalysts are less susceptible to coke formation [94,154,155]. Furthermore, the most used preparation method of these catalysts is the impregnation of the Ni on the support; however, there are only a few studies about the effect of the preparation method in the catalytic behavior of the materials for the steam reforming of long-chain oxygenates.

2.3. Summary

Until now, only a few catalysts have been developed for the OMWSR process: Ni-, Pd-, Pt-, Rh-, and Ru-based. However, several catalysts were studied for the steam reforming of components that are similar to species present in the OMW.
It is possible to conclude that Ni-based catalysts are promising for the steam reforming of oxygenated species. However, almost all these catalysts showed conversions close to 100% and high H2 yields and selectivities, making it difficult to choose the best one. These results show how attractive Ni-based catalysts are for the OMWSR process, in comparison with the noble metal-based catalysts, due to their lower costs.
Almost all Ni-based catalysts were prepared by impregnation, appearing to be the best method for the preparation of these materials. It is still necessary to study the effect of the preparation method on the catalytic behavior of the materials.
A bi-metallic catalyst with Ni and one noble metal (e.g., Ru) as active phases, supported in an adequate material (e.g., Al2O3 or SiO2) and doped with a promotor (e.g., CeO2) is a potential catalyst to achieve high catalytic performance (H2 yield and OMW conversion) for OMWSR during long reaction times.
In future studies, it is necessary to rigorously assess the costs associated with the preparation of the different catalysts (with different preparation methods, reagents, active phases, etc.), to better benchmark the numerous catalytic formulations available, i.e., make a selection (for industrial applications) based on both techno and economic criteria.

3. Multifunctional Reactors

In the last few years, different reactor configurations for improving the catalytic steam reforming processes have been investigated and debated, since this process involves reversible reactions, and thus equilibrium limitations. Thus, the utilization of multifunctional reactors in the OMWSR process can become a very promising approach [160]. Even so, and despite the significant number of studies addressing such devices for other applications, in future works more attention should be paid to detailed techno-economic analyses to assess the economic feasibility of this process, as well as the difference, in financial terms, resulting from the application of different reactor configurations (traditional vs. multifunctional reactors). To the authors’ knowledge, this type of study has not yet been developed for the OMWSR process.

3.1. Sorption-Enhanced Reactor

One of those alternatives for the process intensification of the OMWSR is the utilization of a SER (see Figure 10), which consists of combining the TR and CO2 sorption in the same reactor. This innovative configuration allows shifting the thermodynamic equilibrium of the steam reforming reaction by removing one of the products (in this case CO2) from the reaction medium—Le Chatelier principle [160,161]. This reactor configuration was already used in multiple applications [109,162,163,164,165,166,167,168,169,170,171,172,173,174,175,176,177,178,179,180,181,182,183,184,185,186,187,188,189,190,191,192,193] in the last decade. The SER can shift the equilibrium of the reversible reactions to higher conversions of the reagents. In this way, this reactor configuration will enhance the production of H2, since the removal of CO2 shifts the WGS reaction (Equation (3)) to the forward direction and inhibits CH4 and coke production (Equations (5) and (12)).
However, the drawback of this technology is that after some time the sorbent gets saturated with CO2 and then regeneration is required. To make it possible to run continuously, the system must have at least two parallel devices, so that while one of the reactors is utilized as SER (reaction stage), the other is being regenerated. It is possible to distinguish three regimes along time: pre-breakthrough, breakthrough, and post-breakthrough. In the pre-breakthrough region, the CO2 is being removed, thus enhancing H2 production. In the breakthrough region, the H2 concentration at the reactor outlet begins decreasing. In the pos-breakthrough region, the sorbent is already saturated and so the improvement of the H2 production due to CO2 capture disappears—Figure 11.
The effect of the capture of CO2 from the reaction medium was already discussed for the OMWSR from the thermodynamic point of view [194], though this effect has already been analyzed for other similar reactions [84,195,196]. In a previous work of Rocha et al. [197], a comparison of the performance between a SER and a TR was already performed for this process. This reactor configuration intensifies the performance of the process by combining the reaction and CO2 capture in a single device. In this way, a SER allows the production of more high-purity products at the outlet (in this case, pure H2 before the breakthrough time).
In this reactor configuration, the decrease of the CO2 emissions also presents a very interesting environmental advantage. Since the industrial revolution, the global CO2 concentration increased from about 280 ppm to 415 ppm in 150 years. The higher concentration of pollutants in the atmosphere increases the greenhouse effect with a huge environmental impact. To avoid irreversible changes to the global climate, the United Nations in the COP26 have agreed to establish goals to reduce the CO2 emissions to limit its concentration in the atmosphere [198]. Therefore, this reactor configuration is also in line with this premise, in order to decrease the environmental impact of CO2.

3.1.1. CO2 Sorbents

Sorption-enhanced technologies require the mixture of a catalyst, that is stable and with high catalytic activity in the temperature range considered (in this case 300–500 °C, or even higher), and a solid sorbent, in this case with high CO2 sorption capacity and with adequate sorption–desorption kinetics.
Several materials—such as CaO-based materials [199,200,201,202,203,204,205,206,207,208], hydrotalcites (HTCs) [209,210,211,212,213,214,215,216,217,218,219,220,221,222,223,224], lithium silicates [225,226,227,228,229,230,231,232,233,234,235,236,237,238], lithium cuprate [239] and lithium zirconates [240,241,242], among others—have been reported in the literature as high-temperature CO2 sorbents. It is expected from a good CO2 sorbent for the application envisaged herein to present high CO2 capture capacity at moderate temperatures (300–500 °C), fast sorption–desorption kinetics [243], good mechanical and hydrothermal stability, and low price. A solid with high CO2 sorption capacity with slow kinetics is not suitable since it requires a lot of time for the gas to get into the particle and long times are needed to reach saturation [244], as well as for sorbent regeneration.
Although metal oxides (e.g., CaO) have a higher sorption capacity at high temperatures and can be stabilized when used as a mixture of oxides, they lose their properties very rapidly throughout the sorption–desorption cycles, an issue that is more marked when the adsorbent is in the presence of steam. Besides that, the sorption capacity of these materials decreases because of sintering phenomena [109]. Conversely, HTCs have practically constant capacity over the cycles, increase their capacity in a wet environment, and have great thermal and mechanical stability. Moreover, compared to other possible sorbents such as CaO and lithium zirconates, HCTs show the best adsorption–desorption kinetics and the energy involved in their regeneration is considerably lower [244].
Nevertheless, some research was performed with other sorbents to compete with the HTCs, which is shortly described herein below.
Akgsornpeak et al. [245] avoided the sintering phenomena by preparing CaO materials through the sol–gel synthesis with cetyltrimethyl ammonium bromide. Besides that, the incorporation of Y2O3 has been lately stated to significantly enhance the carbonation (sorption) rate and adsorption capacity of CaO-based materials [206]. The addition of optimized quantities of K to Li2ZrO3 sorbents was reported to enhance their CO2 sorption capacity, sorption rate, and stability in sorption–desorption cycles [226]. Seggiani et al. [226] examined the performance of several doped-Li4SO4 samples, having noticed that the K-doped and Na-doped sorbents presented the highest sorption capacities and sorption rates. In terms of stability, only the K-doped sorbent was capable of maintaining its properties after 25 sorption/desorption cyclic runs. However, Iwand et al. [240] concluded that the traditional preparation method of K-doped materials results in sorbents with slow regeneration step.
The CO2 sorption capacities and the sorption and regeneration temperatures of several materials are summarized in Table 6.

3.1.2. Hydrotalcites

HTC is a layered double hydroxide. This natural or synthetic lamellar hydroxide has two kinds of metallic cations (divalent and trivalent) in the main layers and interlayer space containing anionic species (compensating/interlayer anions)—Figure 12 [248,249]. These materials are used in a wide range of applications as catalysts, precursors, and sorbents [214]. Several thousands of tons of HTCs are produced annually by different chemical companies—such as BASF, SASOL, Clariant, Kisuma Chemicals, and Sakai Chemical [250]. The general formula is M 1 x 2 + M x 3 + OH 2 x + A x / n n · yH 2 O x , where M2+ is a divalent metal cation and M3+ is a trivalent metal cation. An- is a charge balancing anion (usually CO32−) located in the interlayer volume and x is generally between 0.2 and 0.4, while y is the number of moles of water located in the interlayer space [209,210]. Similar layered structures of HTCs could be produced with several interlayer anions [222,224,246,247,250,251,252,253,254,255,256,257].
Co-precipitation is the most used method for the preparation of this type of material. It is based on the slow addition of a mixed solution of divalent and trivalent metals salts in adequate proportions into a reactor. An alkaline solution (e.g., NaOH) is added to the reactor in order to maintain the pH at a selected value, leading to the co-precipitation of the two metallic salts [248]. Therefore, there is often competition between some anionic species, and the tuning of the synthesis experimental conditions can lead to the selective intercalation of one of them as the primary interlamellar anion. Secondary interlamellar anions can replace them with further treatments such as an anionic exchange. In addition, it is not clear that a pure HTC phase is always obtained: for instance, although the overall composition of a certain precipitate corresponds to the estimated value, secondary phases such as hydroxides or basic salts of the divalent or trivalent metal can be present in the material [248]. HTCs can, however, be synthesized by various techniques depending on the specific requirement and properties of the compounds. Other methods include combustion, recrystallization, urea hydrolysis, ion exchange, reverse-microemulsion, sol–gel, and hydrothermal [209,258,259,260,261,262].
The morphology and the thermal stability of the synthesized HTCs affect CO2 sorption performance. The physicochemical properties of HTCs can be modified by combining adequate cations and anions [222,224,246,247,250,251,252,253,254,255,256,257,263,264], by varying the molar ratio of M2+/M3+ [211,212,246], by changing the preparation conditions such as pH [222] (pH values between 8–10 can be used to synthesize HTCs [258]), calcination temperature [246,265,266], or aging process (e.g., microwave, ultrasonication, etc.) [209,223,267] among other parameters. On the other hand, the basicity of the sorbents can be improved by promotion with alkaline species [268], this way increasing their CO2 sorption capacity.
Fresh HTCs as such are not excellent CO2 sorbents due to poor basic properties and the presence of species that hinder CO2 adsorption. Therefore, they are subjected to thermal treatment (calcination) to obtain a nearly amorphous mixed solid solution—with a poorly defined 3D network—with more favorable and stable properties for CO2 sorption [248]. HTCs undergo different stages of transition during the calcination. These transitions depend on many aspects such as the nature and relative quantities of cations, type of anions, method of synthesis, and crystallinity. Several authors have methodically investigated the thermal decomposition process [214,261,263,265,269]. The optimum temperature of calcination is usually 400 °C, since it allows the decomposition of the interlayer anions (and then, the active sites can be occupied by the CO2) and the formation of the amorphous phase (mixed oxides) with sorption properties, without the complete and irreversible destruction of the structure. Apparently, such optimum temperature results from the balance between surface area and the number of basic active sites [214,263]. In the calcination program, the first weight loss is detected between 70 °C and 190 °C, attributed to the loss of interlayer water; then, in the temperature range of 190 °C to 280 °C, the OH group bonded with Al3+ disappears [270]. The OH group bonded with Mg2+ disappears from the structure of the material in the temperature range 280 °C to 405 °C and the interlayer anions are usually released between 405 °C and 580 °C [270]; in this way, the material becomes an amorphous mixed oxide solution, which is reversible if the calcination temperature is not higher than 550–600 °C [270,271]—see Figure 13. The use of very high calcination temperatures leads to the development of spinel phases (e.g., MgAl2O4), which should be prevented in order to conserve the platelet morphology of HTCs and an acceptable basicity [244,246]. The original structures of the calcined samples can be recovered again by exposure to steam and atmospheric CO2 (or through anion intercalation)—“memory effect” [264,272,273,274].
Different optimum Mg2+/Al3+ ratios have been reported in the literature, which varies from 1:1 to 3:1—the optimum ratio depends on the type of interlayer anion used, preparation method, etc. [212,276,277]. Although high Mg2+ content allows the formation of a higher number of basic active sites in the material, at higher Mg2+/Al3+ ratios MgCO3 may also form at high partial pressures of water (steam) and CO2, which can lead to mechanical stability problems [278,279]. In the work of Yong et al. [212], the authors studied the effect of different Mg2+/Al3+ ratios (0.5, 1.3, and 3.0) on the sorption capacity of three commercial HTCs from SASOL. It was verified that the HTC with Mg2+/Al3+ ratio of 1.3 had the highest performance. However, there is a lowest limit of the Mg2+/Al3+ ratio (=2.0) to obtain a crystallographically pure HTC, since this ratio influences the electrostatic repulsion between neighboring trivalent metals in the layers [248,280]. The optimum Mg2+/Al3+ molar ratio reported in several studies is often equal to 2 [209,210,212,255] for the sorbents impregnated with potassium.
As previously mentioned in this section, the CO2 sorption capacity of the HTCs can be enhanced by impregnating with an alkali metal such as Na, K, or Cs [281,282]. It was verified by several works that HTCs modified with alkali promoters like K2CO3 (or Na/KNO3) showed best CO2 sorption performance than the base HTCs [209,210,218,243,281,282,283,284,285,286]—see, for instance, Figure 14. Besides that, several authors [209,210,218,287] have also shown that the simultaneous impregnation with K and substitution of Al3+ with Ga3+ markedly improved the sorption capacity of HTCs, the first modification acting as a chemical promoter and the latter providing a superstructure, robust and stable at elevated temperatures. For K2CO3-modified HTCs with high Mg2+/Al3+ ratios, the doping is mostly concentrated in the bulk phase and behaves as a reactant to form high stable K-Mg double carbonates after sorbed CO2. With the increase of the Al3+ content, surface modifications occur and become the dominant improvement mechanism via the interaction between the K+ ions and unsaturated oxygen sites [283].
It has also been found that the physicochemical properties of HTCs are highly influenced by the interlayer anions used. The anion influences the crystallinity and layer spacing in HTCs. In this way, materials with different interlayer anions present different morphologies and chemical surfaces, influencing the sorption performance of the HTCs-derived mixed oxides. For instance, Yong and Rodrigues [288] reported that the sorption capacity of HTC containing Fe(CN)64- was higher(1.3 mmol·g−1) than the one containing NO3 (0.3 mmol·g−1) at 25 °C and 1 bar of CO2. However, Wang et al. [222] analyzed numerous charge interlayer anions (SO42−, Cl, CO32−, HCO3, and NO3) having noticed that when using CO32−, a spheroidal “sand rose” type of HTC with a very high surface area was produced. On the other hand, the other compensating anions led to the formation of “stone” type HTCs with very low surface areas. For this reason, the HTC with CO32− as interlayer anion presented the highest CO2 capture capacity. However, larger size anions tend to increase the interlayer spacing; this space can help the K ions to disperse into the HTC interlayer and, thus, create more Al(Mg)−O−K bonds to enhance the performance of the sorbent [247].
Several works have shown that the existence of steam in the feed gas increases the CO2 sorption capacity of potassium doped-HTCs by 10–17 wt.% and enhances the stability of these materials (in the temperature range of 200–400 °C and in the pressure range of 1–3.5 bar) [193,216,218,221,279,289,290,291]. It was observed that at least four different adsorption sites take part in the sorption/desorption of CO2 and H2O and the presence of steam activates some active sites for the CO2 sorption. Two adsorption sites can be restored with N2 (or other inert gas), whereas the other adsorption sites require the presence of H2O to be activated/regenerated [279,289,290]. These features are particularly important for the application envisaged herein, once the OMWSR occurs in the presence of large excess of water.
The CO2 sorption kinetics in dry conditions for these materials was studied by Miguel et al. [210] and Silva et al. [209] and it was noticed that the kinetics is dependent on the morphology/composition of the material, with the physical sorption being faster than the chemical one.
The conclusion on which type of CO2 sorbent is more suitable to be used in the SER strongly depends on the operating temperature at which it will be used. Since one of the main goals is to reduce the temperature of the steam reformer to reduce the operation costs, it is attractive to work at relatively low temperatures (300–500 °C). For such temperatures, HTCs have been reported to present higher CO2 sorption capacities and faster sorption–desorption kinetics, as well as easier regeneration, and lower loss of sorption capacity along consecutive cycles. HTCs are also better than other sorbents due to nearly infinite selectivity for CO2 sorption over gases like CH4, CO, and N2 [250]. Moreover, the fact that the CO2 sorption capacity of HTCs is strongly improved under wet conditions (steam reforming process conditions) makes them an even better option for use in the OMWSR process using a SER (for instance, Lu et al. [207] reported that the sorption capacity of a CaO sorbent over the carbonation/decarbonation cycles decreased faster when steam was present in the system).

3.1.3. OMWSR in Sorption-Enhanced Reactors

As already mentioned, in a previous work of Rocha et al. [197], a comparison of the performance between different reactor configurations was performed for the OMWSR process. In a general way, it was verified that the conversion of the organic carbon (above 99%) remains almost complete along all the experimental tests when using a SER (@ 1 bar and 350/400 °C), and it was verified an improvement of the H2 production during the pre-breakthrough of CO2 as compared to the conventional reactor (TR)—see Figure 15.
In this work [197], the experimental tests with the SER configuration were separated into two different sets: a first set of experimental tests using a commercial sorbent, and a second set with a sorbent prepared in the laboratory. It was observed that, at both operating conditions and using the commercial sorbent (Figure 15), the H2 yields obtained in the first cycle were higher than those reached for the first hour in the TR. Furthermore, it was noticed that the utilization of the commercial sorbent (that promotes the removal of CO2 from the reaction medium) increased the stability of the catalyst in comparison with the performance of the TR—lower formation of coke.

3.1.4. Summary

The OMWSR through a SER consists of combining the reaction and CO2 separation in the same device. This innovative reactor configuration requires, besides a catalyst, a solid sorbent with high CO2 sorption capacity, which must be simple to regenerate and have good performance, particularly in the presence of steam. Several materials—such as CaO-based materials, lithium cuprate, lithium silicates, and lithium zirconates, among others—have been reported in the literature as high-temperature CO2 sorbents. However, it is possible to conclude that HTCs have adequate properties to be used in a SER for OMWSR: practically constant and high CO2 sorption capacity over the cycles at moderated temperatures, increased performance in a wet environment, adequate sorption–desorption kinetics, and great thermal and mechanical stability.
In this way, an HTC impregnated with potassium is a potential sorbent to be used to capture from the reaction medium the CO2 produced in the OMWSR (sorption-enhanced process), increasing the purity of the H2 and shifting the thermodynamic equilibrium to the production of more H2 (increasing the H2 yield).

3.2. Membrane Reactor

Another alternative to enhance the OMWSR performance is the utilization of a MR (see Figure 16), which consists of combining the TR and H2 separation in the same device. In such an MR, the membrane selectively removes one or more species, thus allowing to overcome the thermodynamic boundaries of equilibrium-limited reactions—Le Châtelier principle [160,163,164,165,166,167,168,169,170,171,172,173,174,175,176,177,178,179,180,181,182,183,184,185,186,187,188,189,190,191,192,193,194,195,196,197,198,199,200,201,202,203,204,205,206,207,208,209,210,211,212,213,214,215,216,217,218,219,220,221,222,223,224,225,226,227,228,229,230,231,232,233,234,235,236,237,238,239,240,241,242,243,244,245,246,247,248,249,250,251,252,253,254,255,256,257,258,259,260,261,262,263,264,265,266,267,268,269,270,271,272,273,274,275,276,277,278,279,280,281,282,283,284,285,286,287,288,289,290,291,292]—as is the case of the reversible WGS, which is, therefore, shifted forward. Regarding its advantages compared to TRs, they are the same as SERs, except for the reduction of the emissions of CO2 in the pre-breakthrough zone, being that the capital costs reduction can be higher in the case of MRs since only one device is needed (instead of the parallel reactors required for the SER), as long as low-cost and long-lasting H2 perm-selective membranes can be obtained.
This reactor configuration was already used in other applications [173,293,294,295,296,297,298,299,300,301,302,303,304] and for the OMWSR [1,6,89,90,197] in the last decade.
One of the advantages of the MR for this particular application is that it is possible to produce ultra-pure H2 streams if membranes with a nearly infinite H2 perm-selectivity are used (e.g., dense Pd-based membrane). Thus, a MR is an important technology to increase the energy performance of the process by combining the reaction and recovery stages in a single device (avoiding the construction of two unit operations), leading to considerably smaller dimensions of plant layout, producing one stream with high-purity H2 (in the permeate side—see a scheme of a MR in Figure 16). The effect of the removal of H2 from the reaction medium, from a thermodynamic point of view, was already studied for the OMWSR [194], though it was already analyzed for other applications [84,196,293,299,305,306,307].
In order to use a MR in the OMWSR process, it is fundamental to select a suitable H2 perm-selective membrane. Consequently, the next sections comprise a short overview of H2 perm-selective membranes described in the literature, as well as an evaluation of the experimental tests performed in MRs for OMWSR. The performance of a MR depends considerably on the flux of permeated H2 obtained and on its purity. The stability of a membrane is here considered as the capability of a membrane to maintain a constant flux of H2, as well as a constant perm-selectivity over time.

3.2.1. Hydrogen Perm-Selective Membranes

H2 perm-selective membranes may be categorized as dense metallic membranes, proton-conducting dense ceramic membranes, dense polymeric membranes, microporous ceramic membranes, and porous carbon membranes [308,309]. The main benefit of dense metallic membranes is that typically they tolerate high temperatures (e.g., as compared to polymeric ones) and present high selectivity towards H2, thus being possible to obtain high purity H2-permeate streams [308]. Thus, and for the application envisaged herein, dense metallic membranes are preferable for ultra-pure H2 production and will be the main target of focus. Nevertheless, dense ceramic membranes (proton conducting membranes) have also achieved high performance at high operation temperatures [310,311]. Although Pd-based membranes have been more used in the past for H2 permeation, the advances in dense ceramic membranes in the last few years have been outstanding.
When selecting a membrane to be used in a MR, it is required to take into consideration the operating conditions under which the membrane will be applied. In the last years, there has been an incredible effort in order to develop appropriate membranes for MR applications. Pd-based membranes are presently the most promising for high purity H2 production in MRs, especially Pd alloys (with transition metals—e.g., Pd-Ag, Pd-Ru, Pd-Cu, and Pd-Au), which are less sensitive than pure Pd to embrittlement (caused by the presence of H2 at temperatures below 300 °C and H2 pressures below 20 bar) and poisoning (caused by contact with CO, CO2, or H2S) [308,312,313,314,315,316,317]. The embrittlement phenomenon may produce pinholes/cracks on the membrane, which will negatively affect the H2 perm-selectivity of the membrane. In addition, one of the biggest inhibitory factors for H2 permeation across a Pd membrane is concentration polarization and it is necessary to avoid that (this topic will be more deeply discussed later in this section) [318]. However, and in spite of these negative aspects, the ability of Pd to dissociate H2 to its atomic form makes this metal ideal for H2 separation applications [319]. Pd-Ag membranes show good stability, relatively lower material costs, and better mechanical properties than pure Pd membranes. Besides, it has been shown that the addition of a third element to the Pd-Ag alloy (e.g., Au) improves the tolerance to poisoning with H2S [320].
Several studies reported that the Pd alloys can obtain high permeances [321,322,323,324,325]. Moreover, their H2 permeability increases with the percentage of Ag in the alloy up to a maximum, which occurs at around 23 wt.% Ag [326]. In previous studies, it was found that the maximum recommended operating temperature for long-term stable tests with Pd-Ag membranes was close to 500 °C [297,327,328,329,330]. Above this temperature, besides the existence of leakages through the sealings, it is observed the formation of surface defects (small pinholes) that leads to the decrease of the membrane performance [327]. Tosti et al. [331] demonstrated that self-supported dense 23 wt.% Pd-Ag tubular membranes with finger-type configuration are extremely durable and reliable since these membranes allow to attain complete H2 selectivity and no malfunctions were observed after one year of hydrogenation and thermal cycles. These properties, together with high permeability, make such materials very promising for ultra-pure H2 production in energetic and industrial applications [331]. Nevertheless, several works reported that the H2 removal fraction never reaches a value higher than 0.80 [305,326,332] for low-to-medium pressures. Besides that, the application of these dense membranes is still limited; for instance, self-supported Pd membranes are considered by several authors not being suitable in MR processes because of their large thickness required for reaching mechanical stability and corresponding low fluxes across the membrane and high costs. Due to these reasons, a strong effort has been put in the preparation of thinner Pd films over different supports and in the development of new membranes with less expensive metals. For instance, group V metals (such as V, Nb, and Ta) are promising alternative materials to Pd because of their lower costs and higher H2 permeabilities. However, the main problem of these elements as membrane materials is their too high solubility to H2, making the membranes susceptible to severe H2 embrittlement, which can lead to the formation of cracks [333]. Still, direct use of these materials as a membrane is hindered due to the formation of oxide layers and the occurrence of surface reactions, which promote the reduction of the H2 permeability through the membrane. Thus, the main criteria for membrane selection are high permeability, high selectivity, resistance to reactive gases—particularly CO, CO2, CH4 and H2O—and resistance to coking [334].
Nowadays, many researchers are pursuing the possibility of using supported membranes with as thin as possible Pd-alloy layers [335], thus reducing costs and increasing hydrogen permeation fluxes. Concerning these supported membranes, it is important to emphasize that metallic supports are mechanically stronger and can be more easily assembled into metallic reactors than ceramics supports (lower probability of leakages). However, the metallic supports have low porosity, their pores are larger, the surface is rough and, at this moment, they are much more costly than ceramic supports [333]. Several works [329,336] reported that the materials (namely, the support) that contact with the membrane must be meticulously chosen. For instance, Okazaki et al. [329] stated that the decrease of the H2 permeation observed in their work was related to the interaction between the Al2O3 support and the Pd layer of the membrane.
In terms of preparation methods of these H2-selective membranes, namely supported ones, chemical vapor deposition (CVD) and electroless plating (EP) and are the most used ones. The last method (EP) normally consists of the production of Pd particles through the reduction of a plating solution containing Pd-amine complexes. This method can provide high coating adhesion, requires low operation costs and allows simple operation. Though, the impurities present in the plating solution may lead to the presence of defects on the Pd layer. The CVD technique can very easily allow to deposit a metal film on a support with the utilization of several reactions with water [293].
Generally, the permeation of H2 through a Pd-based membrane is described by a solution-diffusion mechanism composed by several steps, as shown in Figure 17.
The simplest model used to explain the mass transport of H2 permeating through a dense membrane assumes that the H2 diffuses through the membrane according to Fick’s first law of diffusion for a pure component [338]. Still, in the permeation of H2 through Pd-based membranes is commonly assumed that the diffusion coefficient is constant along the membrane and the mass transport kinetics is controlled by the diffusion step; this implies that the sorption and desorption process should be considered always to be at equilibrium (concentration of H2 in the surface of the membrane in equilibrium with the gas phase—in the retentate and permeate sides); in the case of Pd-based membranes, the sorption isotherm is generally described by the Sieverts’ law [339]. Taking into account all these considerations, the flux of H2 that permeates through the membrane ( J H 2 ) can be determined by the integration of Fick’s first law of diffusion along the thickness of the membrane (Equation (13)) [339,340,341].
J H 2 = L H 2 δ × p H 2 ,   feed 0.5 p H 2 ,   permeate 0.5
where L H 2 is the permeability of the membrane, δ is the membrane thickness, p H 2 is the partial pressures of H2 in the retentate (inlet—feed side) or in the permeate (outlet) side. The ratio L H 2 δ is normally called permeance.
The permeability is described by an Arrhenius-type dependency on the temperature (T)—Equation (14)
L H 2 = L H 2 0 × e E a R T
where L H 2 0 is the pre-exponential factor, E a is the activation energy of the membrane permeation process, and R is the universal gas constant.
Despite the extensive use of Equation (13), it is necessary to emphasize that this equation is valid only for certain operation conditions: for instance, the expression is only applicable to ideal H2/metal mixtures and when the resistance in the diffusion of the H2 along the thickness of the membrane is the rate-dominant process [323]. In general, the flux of H2 can be influenced by several factors such as temperature, partial pressure difference, membrane thickness, membrane diffusivity, adsorption/desorption kinetics, and concentration polarization at the surface of the membrane. Ward and Dao [342] reported that for a clean Pd membrane, in the absence of external mass transfer resistance, the diffusion is the dominating step at moderately high temperatures, even for membrane thicknesses close to 1 μm.
However, as also reported by Ward and Dao [342] (and shown in Figure 17), beyond the H2 diffusion through the membrane, other phenomena occur in the permeation process and the resistances to mass transport in the external gas phase may also be significant in some cases. The concentration polarization phenomenon is related to a decreased H2 concentration in the feed side of the membrane in a multi-component system. In certain cases, due to the fluid dynamics inside the device, there is an accumulation of several species in the adjacent zone to the membrane surface, increasing the concentration of these species in the boundary film nearby to the membrane surface [343]. This causes a H2 partial pressure gradient in the boundary layer and prevents the H2 adsorption on the membrane surface (increase of the mass transport resistance). Thus, the H2 concentration adjacent to the surface of the membrane is lower than that in the bulk gas and thus decreases the flux of H2 permeated through the membrane [344,345,346,347,348]. The concentration polarization is affected not only by the external hydrodynamic conditions of the module, but also by the mass transport process inside of the membrane layer [349]. While most of the existing permeation results for thick Pd membranes are consistent with the considerations for the diffusion-limited permeation, significant differences exist for membranes with less than 10 μm thickness, since the very high fluxes made possible with such thin membranes makes the external mass transfer resistances relevant [342].
In this manner, due to the several possible phenomena that occur in this separation process, it is typical to use the empirical Equation (15) to describe the permeating H2 flux [340,341]. It was reported in the literature that the pressure exponent n, in a general way, can provide insight regarding which transport step(s) control the overall permeation process. Several works studied and evaluated the physical meaning of the exponent n [342,350,351] for H2 permeation through Pd-based membranes.
J H 2   = L H 2 δ × p H 2 ,   feed n p H 2 ,   permeate n
As already mentioned, the parameter n obtained is near 0.5 for a situation of no external mass transfer resistance (and the diffusion through the membrane controls the permeation process) [342,351,352]. It was also observed that the required value of the exponent n usually varies between 0.5 and 1 in the presence of external mass transfer resistances (effects are particularly strong on the permeate side) [342]. Moreover, it was reported that, in the absence of external mass transfer resistances, the n value approaches zero when the H2 permeation is desorption-limited and varies between 0 and 0.5 when the desorption and diffusion are both influencing the H2 flux [342]. It was also stated that (again in the absence of external resistances) the value of the exponent n is equal to the unity if only the adsorption process in the feed side limits the H2 flux, and varies between 0.5 and 1 when adsorption and diffusion in the metal bulk are both limiting the permeation [352]. In line with the aforementioned, the study of Caravella et al. [351] verified that, at low temperature and using a thin membrane, the value of the exponent n decreases toward values lower than 0.5—desorption-limited process. However, higher values (>1) for the pressure exponent have been determined in previous works [352,353]. Besides that, the H2 diffusion coefficient may also exhibit a concentration dependence and the membrane may present defects on the membrane films which may lead, for instance, to Knudsen diffusion.
Table 7 and Table 8 show a summary of different Pd-based membranes that have been described in the last few years with different parameters and performances.

3.2.2. OMWSR in Membrane Reactors

Just a few experimental works regarding the OMWSR in MRs are reported in the literature [1,6,89,90]; in such studies, the authors assessed the effects of pressure and gas/weight hourly space velocity (GHSV/WHSV) and will be summarized in this section.
In the first work of Tosti et al. [6], filtered and concentrated OMW was used for feeding a MR consisting of a dense Pd-Ag permeator tube packed with a Pt-based catalyst. Reforming tests were performed at 450 °C in the range of 1–5 bar total pressure. The experimental tests have shown the ability of the membrane device to selectively separate the H2 produced and to provide a retentate stream rich in non-permeated H2, CO2, and CH4. A maximum H2 yield equivalent to 2 kg of H2 permeated per ton of OMW has been measured, when operating at 3 bar with a space velocity of 5.56 × 103 mol·h−1·g−1cat. The gas chromatographic analysis of the retentate stream showed the presence of H2, CO2, and CH4, thus proving evidence of the presence of side reactions such as the hydrogenolysis and cracking of alcohols. No detectable quantities of CO were determined due to the completion of the WGS promoted by the presence of the membrane. Additionally, a deactivation of the catalyst due to coke deposits was observed.
In the second work, Tosti et al. [89] conducted again the OMWSR process in a MR (using a Pd membrane) to investigate the capability to recuperate pure H2 and H2-rich gas mixtures from OMW. Specifically, a Pt-Rh-Pd-based catalyst supported on rare earth oxides was used in order to validate the capability of this reactor configuration of maximizing the H2 yield and decreasing side reactions. As a main result, this work permitted verifying the following capabilities of the studied membrane process: the production of H2 from waste (up to 3.25 kg of H2 per ton of OMW) and a reduction in the pollution potential of the OMW (phenols and total carbon concentrations are reduced by ≈90%). Compared to a commercial Pt-based material that was used in the previous work, the new catalyst demonstrated higher selectivity toward the steam reforming reaction, permitting a substantial decrease in the formation of CH4 and coke and thus exhibiting a higher H2 yield and higher stability of the catalyst. In fact, in all experiments, low CH4 concentrations were found in the retentate stream, while no significant loss of catalyst activity was detected after 6–8 h of operation. The low selectivity of the noble metals-based catalyst toward CH4 formation was related to its intrinsic high reforming activity.
In the last work of the same research group [1], they tested the reforming of OMW and CH4 in a Pd-membrane reformer and demonstrated that the steam reforming of CH4 + OMW exhibits about the same H2 yield of the reforming of the OMW alone. Besides that, the addition of CH4 to OMW does not affect the ability of the steam reforming process to reduce pollution. Finally, it was possible to verify that the H2 yield reached in the steam reforming of OMW together with CH4 was higher than the H2 yield obtained for the steam reforming of CH4 alone.
More recently, Alique et al. [90], using a catalytic material prepared by Tosti et al. in a previous study [89] reported that, in a MR, higher pressures were beneficial, dominating the membrane shift effect (with enhanced H2 permeating fluxes) against the negative influence of thermodynamics on the steam reforming reactions—see Figure 18.
In addition, and as previously mentioned, in the work of Rocha et al. [197], several reactor configurations were tested for the OMWSR process—namely the SER and the SEMR (@ 1 bar and 350/400 °C). The difference between the H2 yields obtained in the SER and SEMR is considerable, even after saturation of the sorbent (instant after which the system began to behave like a MR), showing that the impact of the membrane is very critical in this multifunctional system.

3.2.3. Summary

The OMWSR through a MR consists of combining the reaction and H2 separation in the same device. This innovative reactor configuration requires a H2 perm-selective membrane.
Pd-based membranes are currently promising for high purity H2 production in MRs. Since the price of these dense membranes is still a restricting factor, a major effort has been put into the preparation of thinner Pd films over different supports, by several techniques. This decrease of the thickness of the membrane will allow the increase of the permeation flux, but the selectivity will be lower.
In this way, a thin supported Pd membrane is a potential membrane to separate from the reaction medium the H2 produced in the OMWSR, in order to increase the H2 yield (since it shifts the thermodynamic equilibrium to the production of more H2) and produce an ultra-pure H2 stream. However, it is necessary to further study the effect of the species present in the OMW on the membrane performance.

3.3. Sorption-Enhanced Membrane Reactor

Another technological alternative is the multifunctional hybrid reactor combining capture of CO2 and H2 separation: the SEMR (see Figure 19). This hybrid reactor configuration allows shifting the thermodynamic equilibrium of the steam reforming reaction by removing two of the reaction products from the reaction medium: hydrogen and carbon dioxide. This reactor configuration, intensifies the process even further by combining the reaction, H2 separation, and CO2 capture in a single device. This reactor configuration was not studied in detail until now for the OMWSR with the exception of a recent experimental work [197] and one (theoretical) thermodynamic study [194]. The SEMR combines the advantages of a SER and a MR; however, like in a SER, the CO2 sorbent gets saturated at a certain point and then regeneration is needed (see details in Section 3.1). At this point, this reactor configuration is similar to a MR, since the CO2 is not captured, but the H2 is continuously separated from the reaction medium to the permeate side.

OMWSR in Sorption-Enhanced Membrane Reactors

As already stated, in the study of Rocha et al. [197], a comparison of the performance between different reactor configurations was performed for the OMWSR process (TR, SER, MR, and SEMR) @ 1 bar and 350/400 °C.
It was observed that the removal of H2 and capture of CO2 could provide higher H2 yields in the SEMR during the pre-breakthrough region in comparison with the other hybrid multifunctional reactors. This reactor configuration allows the simultaneous production of highly pure H2 in both the retentate and permeate sides. An H2 yield improvement of about 44% was seen compared to the TR, approximately 39% regarding the SER and about 16% compared to the MR.
The H2 yield observed in the SEMR at 350 °C and 1 bar was permanently between 8.0 and 6.0 molH2·molOMW−1, and always between 11.5 and 9.0 molH2·molOMW−1 at 400 °C and 4 bar. This way, during the pre-breakthrough region, the H2 yield reached practically the maximum theoretical yield (11.84 molH2·molOMW−1) in all cycles at 400 °C and 4 bar—see Figure 20. In the first cycle, the H2 yield inclusively goes over the thermodynamic equilibrium yield calculated for the TR in these operating conditions (cf. Figure 20). Besides that, CH4, CO, and CO2 are not detected in the pre-breakthrough region, thus preventing the poisoning of the membrane and minimizing the coke deposits formation (particularly via CH4 cracking).
Finally, with these experimental tests, it was observed that the OMWSR, when carried in a SEMR, can competently treat real OMW effluents, with high productions of H2, with complete decrease of the pollutant load.

4. Conclusions

OMW is a polluting stream derived from the olive oil industry and is an important source of environmental pollution, particularly in the Mediterranean countries. This stream is typically constituted by polyphenols, sugars, fatty acids, and water. Nowadays, and in order to reduce the pollutant load, several treatment/valorization techniques are applied, but these technologies have large cost and efficiency problems. Thus, OMWSR presents as a good alternative, since this process decreases the pollutant load of the OMW and simultaneously the waste is valorized with the production of green H2.
Currently, the OMWSR is an innovative treatment with high potential; however, alternatives for improving the process have been studied, since this process involves reversible reactions. In this way, it is expected that several reactor configurations (SER, MR, and SEMR) could work at lower temperatures than the conventional steam reformers and attain similar or even better performances than TRs operating at higher temperatures. Nevertheless, for the successful implementation of these reactor configurations, CO2 sorbents with high sorption capacity, easy regeneration, and high stability—as well as H2 perm-selective membranes with high H2 permeance and selectivity—have to be used. Nowadays, HTCs and Pd-based membranes are seen as promising systems for use in the SEMR at temperatures in the range of 300–400 °C.

Author Contributions

Conceptualization, M.A.S. and L.M.M.; investigation, C.R.; resources, L.M.M.; writing—original draft preparation, C.R.; writing—review and editing, M.A.S. and L.M.M.; supervision, M.A.S. and L.M.M.; funding acquisition, M.A.S. and L.M.M. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Base Funding—UIDB/00511/2020—of the Laboratory for Process Engineering, Environment, Biotechnology and Energy (LEPABE) funded by national funds through the FCT/MCTES (PIDDAC), and by the Project NORTE-01-0247-FEDER-39789, funded by European Regional Development Fund (ERDF) through the Programa Operacional Regional do Norte (NORTE 2020). This work was also financially supported by: Base Funding—UIDB/50020/2020—of the Associate Laboratory LSRE-LCM, funded by national funds through FCT/MCTES (PIDDAC); and project HyGreen&LowEmissions, with reference NORTE-01-0145-FEDER-000077, through the Programa Operacional Regional do Norte (NORTE 2020), under the PORTUGAL 2020 Partnership Agreement, funded by the European Regional Development Fund (ERDF). M.A. Soria thanks the Portuguese Foundation for Science and Technology (FCT) for the financial support of his work contract through the Scientific Employment Support Program (Norma Transitória DL 57/2017).

Conflicts of Interest

The authors declare no conflict of interest.

Notation and Glossary

List of Variables
E a Activation energy (kJ·mol−1)
J H 2 Flux of H2 (mol·m−2·s−1)
L H 2 Permeability (mol·m−1·s−1·Pa−0.5)
L H 2 0 Pre-exponential factor (mol·m−1·s−1·Pa−0.5)
nPressure exponent
p H 2 x Partial pressures of H2 (Pa)
RUniversal gas constant
TTemperature (°C or K)
Δ H r 0   ° C Enthalpy of reaction at standard conditions (kJ·mol−1)
δ Membrane thickness (m)
List of Acronyms
BODBiochemical oxygen demand
CODCarbon oxygen demand
CVDChemical vapor deposition
EPElectroless plating
HTCHydrotalcite
LHSVLiquid hourly space velocity
MRMembrane reactor
OMWOlive oil mill wastewater
OMWSROlive oil mill wastewater steam reforming
SCFRSteam-to-carbon feed ratio
SEMRSorption-enhanced membrane reactor
SERSorption-enhanced reactor
TOCTotal organic carbon
TPOMWTwo-phase olive mill waste
TRTraditional reactor
WGSWater–gas shift

References

  1. Tosti, S.; Fabbricino, M.; Pontoni, L.; Palma, V.; Ruocco, C. Catalytic reforming of olive mill wastewater and methane in a Pd-membrane reactor. Int. J. Hydrogen Energy 2016, 41, 5465–5474. [Google Scholar] [CrossRef]
  2. Council, I.O. Prices & Balances. 2021. Available online: https://www.internationaloliveoil.org/wp-content/uploads/2021/12/IOC-Olive-Oil-Dashboard-1.html (accessed on 18 October 2021).
  3. Roig, A.; Cayuela, M.L.; Sánchez-Monedero, M.A. An overview on olive mill wastes and their valorisation methods. Waste Manag. 2006, 26, 960–969. [Google Scholar] [CrossRef]
  4. Rozzi, A.; Malpei, F. Treatment and disposal of olive mill effluents. Int. Biodeterior. Biodegrad. 1996, 38, 135–144. [Google Scholar] [CrossRef]
  5. Dawson, D. Global Olive Oil Production to Dip in 2018/2019. 2018. Available online: https://www.oliveoiltimes.com/olive-oil-business/global-olive-oil-production-to-dip-in-2018-19/66039 (accessed on 25 October 2021).
  6. Tosti, S.; Accetta, C.; Fabbricino, M.; Sansovini, M.; Pontoni, L. Reforming of olive mill wastewater through a Pd-membrane reactor. Int. J. Hydrogen Energy 2013, 38, 10252–10259. [Google Scholar] [CrossRef]
  7. Dermeche, S.; Nadour, M.; Larroche, C.; Moulti-Mati, F.; Michaud, P. Olive mill wastes: Biochemical characterizations and valorization strategies. Process Biochem. 2013, 48, 1532–1552. [Google Scholar] [CrossRef]
  8. Alburquerque, J.A.; Gonzálvez, J.; García, D.; Cegarra, J. Agrochemical characterisation of “alperujo”, a solid by-product of the two-phase centrifugation method for olive oil extraction. Bioresour. Technol. 2004, 91, 195–200. [Google Scholar] [CrossRef]
  9. Di Giovacchino, L.; Sestili, S.; di Vincenzo, D. Influence of olive processing on virgin olive oil quality. Eur. J. Lipid Sci. Technol. 2002, 104, 587–601. [Google Scholar] [CrossRef]
  10. Niaounakis, M.; Halvadakis, C.P. Characterization of olive processing waste. Olive Process. Waste Manag. Lit. Rev. Pat. Surv. 2006, 5, 23–64. [Google Scholar]
  11. Sanchez Moral, P.; Méndez, M.R. Production of pomace olive oil. Grasas Y Aceites 2006, 57, 47–55. [Google Scholar] [CrossRef]
  12. Aktas, E.S.; Imre, S.; Ersoy, L. Characterization and lime treatment of olive mill wastewater. Water Res. 2001, 35, 2336–2340. [Google Scholar] [CrossRef]
  13. Kapellakis, I.E.; Tsagarakis, K.P.; Crowther, J.C. Olive oil history, production and by-product management. Rev. Environ. Sci. Bio/Technol. 2008, 7, 1–26. [Google Scholar] [CrossRef]
  14. Caputo, A.C.; Scacchia, F.; Pelagagge, P.M. Disposal of by-products in olive oil industry: Waste-to-energy solutions. Appl. Therm. Eng. 2003, 23, 197–214. [Google Scholar] [CrossRef]
  15. Paredes, M.J.; Moreno, E.; Ramos-Cormenzana, A.; Martinez, J. Characteristics of soil after pollution with wastewaters from olive oil extraction plants. Chemosphere 1987, 16, 1557–1564. [Google Scholar] [CrossRef]
  16. DellaGreca, M.; Monaco, P.; Pinto, G.; Pollio, A.; Previtera, L.; Temussi, F. Phytotoxicity of low-molecular-weight phenols from olive mill wastewaters. Bull. Environ. Contam. Toxicol. 2001, 67, 352–359. [Google Scholar] [CrossRef]
  17. Rana, G.; Rinaldi, M.; Introna, M. Volatilisation of substances alter spreading olive oil waste water on the soil in a Mediterranean environment. Agric. Ecosyst. Environ. 2003, 96, 49–58. [Google Scholar] [CrossRef]
  18. Anzelmo, B.; Wilcox, J.; Liguori, S. Hydrogen production via natural gas steam reforming in a Pd-Au membrane reactor. Investigation of reaction temperature and GHSV effects and long-term stability. J. Membr. Sci. 2018, 565, 25–32. [Google Scholar] [CrossRef]
  19. Tsagaraki, E.; Lazarides, H.N.; Petrotos, K.B. Olive Mill Wastewater Treatment. In Utilization of By-Products and Treatment of Waste in the Food Industry; Springer: Boston, MA, USA, 2007; pp. 133–157. [Google Scholar]
  20. Jaouad, Y.; Villain-Gambier, M.; Mandi, L.; Marrot, B.; Ouazzani, N. Comparison of aerobic processes for olive mill wastewater treatment. Water Sci. Technol. 2020, 81, 1914–1926. [Google Scholar] [CrossRef]
  21. Aggoun, M.; Arhab, R.; Cornu, A.; Portelli, J.; Barkat, M.; Graulet, B. Olive mill wastewater microconstituents composition according to olive variety and extraction process. Food Chem. 2016, 209, 72–80. [Google Scholar] [CrossRef]
  22. El-Abbassi, A.; Kiai, H.; Hafidi, A. Phenolic profile and antioxidant activities of olive mill wastewater. Food Chem. 2012, 132, 406–412. [Google Scholar] [CrossRef]
  23. Zbakh, H.; el Abbassi, A. Potential use of olive mill wastewater in the preparation of functional beverages: A review. J. Funct. Foods 2012, 4, 53–65. [Google Scholar] [CrossRef]
  24. Nasr, B.; Ahmed, B.; Abdellatif, G. Fenton treatment of olive oil mill wastewater-applicability of the method and parameters effects on the degradation process. J. Environ. Sci. 2004, 16, 942–944. [Google Scholar]
  25. Katsoyannos, E.; Hatzikioseyian, A.; Remoundaki, E.; Tsezos, M. Photocatalytic treatment of olive mill wastewaters (OMW) in pilot scale. In Proceedings of the 13th International Symposium on Pollution and its Impact on Life in Mediterranean Region, Thessaloniki, Greece, 8–12 October 2005; National Technical University of Athens Greece: Athens, Greece, 2005. [Google Scholar]
  26. Gebreyohannes, A.Y.; Mazzei, R.; Giorno, L. Trends and current practices of olive mill wastewater treatment: Application of integrated membrane process and its future perspective. Sep. Purif. Technol. 2016, 162, 45–60. [Google Scholar] [CrossRef]
  27. Montero, C.; Oar-Arteta, L.; Remiro, A.; Arandia, A.; Bilbao, J.; Gayubo, A.G. Thermodynamic comparison between bio-oil and ethanol steam reforming. Int. J. Hydrogen Energy 2015, 40, 15963–15971. [Google Scholar] [CrossRef]
  28. Casanovas, A.; Galvis, A.; Llorca, J. Catalytic steam reforming of olive mill wastewater for hydrogen production. Int. J. Hydrogen Energy 2015, 40, 7539–7545. [Google Scholar] [CrossRef] [Green Version]
  29. Daâssi, D.; Lozano-Sánchez, J.; Borrás-Linares, I.; Belbahri, L.; Woodward, S.; Zouari-Mechichi, H.; Mechichi, T.; Nasri, M.; Segura-Carretero, A. Olive oil mill wastewaters: Phenolic content characterization during degradation by Coriolopsis gallica. Chemosphere 2014, 113, 62–70. [Google Scholar] [CrossRef] [PubMed]
  30. Fki, I.; Allouche, N.; Sayadi, S. The use of polyphenolic extract, purified hydroxytyrosol and 3,4-dihydroxyphenyl acetic acid from olive mill wastewater for the stabilization of refined oils: A potential alternative to synthetic antioxidants. Food Chem. 2005, 93, 197–204. [Google Scholar] [CrossRef]
  31. Kyriacou, A.; Lasaridi, K.E.; Kotsou, M.; Balis, C.; Pilidis, G. Combined bioremediation and advanced oxidation of green table olive processing wastewater. Process Biochem. 2005, 40, 1401–1408. [Google Scholar] [CrossRef]
  32. Araújo, M.; Pimentel, F.B.; Alves, R.C.; Oliveira, M.B.P.P. Phenolic compounds from olive mill wastes: Health effects, analytical approach and application as food antioxidants. Trends Food Sci. Technol. 2015, 45, 200–211. [Google Scholar] [CrossRef]
  33. Kaleh, Z.; Geißen, S.U. Selective isolation of valuable biophenols from olive mill wastewater. J. Environ. Chem. Eng. 2016, 4, 373–384. [Google Scholar] [CrossRef]
  34. Vlyssides, A.G.; Loizides, M.; Karlis, P.K. Integrated strategic approach for reusing olive oil extraction by-products. J. Clean. Prod. 2004, 12, 603–611. [Google Scholar] [CrossRef]
  35. Paredes, C.; Cegarra, J.; Roig, A.; Sánchez-Monedero, M.A.; Bernal, M.P. Characterization of olive mill wastewater (alpechin) and its sludge for agricultural purposes. Bioresour. Technol. 1999, 67, 111–115. [Google Scholar] [CrossRef]
  36. Feki, M.; Allouche, N.; Bouaziz, M.; Gargoubi, A.; Sayadi, S. Effect of storage of olive mill wastewaters on hydroxytyrosol concentration. Eur. J. Lipid Sci. Technol. 2006, 108, 1021–1027. [Google Scholar] [CrossRef]
  37. Hamden, K.; Allouche, N.; Damak, M.; Elfeki, A. Hypoglycemic and antioxidant effects of phenolic extracts and purified hydroxytyrosol from olive mill waste in vitro and in rats. Chem. Biol. Interact. 2009, 180, 421–432. [Google Scholar] [CrossRef] [PubMed]
  38. Justino, C.I.; Duarte, K.; Loureiro, F.; Pereira, R.; Antunes, S.C.; Marques, S.M.; Gonçalves, F.; Rocha-Santos, T.A.P.; Freitas, A.C. Toxicity and organic content characterization of olive oil mill wastewater undergoing a sequential treatment with fungi and photo-Fenton oxidation. J. Hazard. Mater. 2009, 172, 1560–1572. [Google Scholar] [CrossRef]
  39. Zghari, B.; Pierre, D.; Abderrahmane, R.; Boukir, A. GC-MS, FTIR and 1H,13C NMR Structural Analysis and Identification of Phenolic Compounds in Olive Mill Wastewater Extracted from Oued Oussefrou Effluent (Beni Mellal-Morocco). J. Mater. Environ. Sci. 2017, 8, 4496–4509. [Google Scholar] [CrossRef]
  40. El-Abbassi, A.; Khayet, M.; Kiai, H.; Hafidi, A.; García-Payo, M.C. Treatment of crude olive mill wastewaters by osmotic distillation and osmotic membrane distillation. Sep. Purif. Technol. 2013, 104 (Suppl. C), 327–332. [Google Scholar] [CrossRef]
  41. El-Abbassi, A.; Kiai, H.; Hafidi, A.; García-Payo, M.C.; Khayet, M. Treatment of olive mill wastewater by membrane distillation using polytetrafluoroethylene membranes. Sep. Purif. Technol. 2012, 98 (Suppl. C), 55–61. [Google Scholar] [CrossRef]
  42. Allaoui, S.; Bennani, M.N.; Ziyat, H.; Qabaqous, O.; Tijani, N.; Ittobane, N. Removing polyphenols contained in olive mill wastewater by membrane based on natural clay and Hydrotalcite Mg-Al. Moroc. J. Chem. 2020, 8, 318–325. [Google Scholar] [CrossRef]
  43. Ochando-Pulido, J.M.; Pimentel-Moral, S.; Verardo, V.; Martinez-Ferez, A. A focus on advanced physico-chemical processes for olive mill wastewater treatment. Sep. Purif. Technol. 2017, 179, 161–174. [Google Scholar] [CrossRef]
  44. Un, U.T.; Altay, U.; Koparal, A.S.; Ogutveren, U.B. Complete treatment of olive mill wastewaters by electrooxidation. Chem. Eng. J. 2008, 139, 445–452. [Google Scholar] [CrossRef]
  45. Rioja-Cabanillas, A.; Valdesueiro, D.; Fernández-Ibáñez, P.; Byrne, J.A. Hydrogen from wastewater by photocatalytic and photoelectrochemical treatment. J. Phys. Energy 2020, 3, 012006. [Google Scholar] [CrossRef]
  46. Cañizares, P.; Paz, R.; Sáez, C.; Rodrigo, M.A. Costs of the electrochemical oxidation of wastewaters: A comparison with ozonation and Fenton oxidation processes. J. Environ. Manag. 2009, 90, 410–420. [Google Scholar] [CrossRef] [PubMed]
  47. Lafi, W.K.; Shannak, B.; Al-Shannag, M.; Al-Anber, Z.; Al-Hasan, M. Treatment of olive mill wastewater by combined advanced oxidation and biodegradation. Sep. Purif. Technol. 2009, 70, 141–146. [Google Scholar] [CrossRef]
  48. Frascari, D.; Rubertelli, G.; Arous, F.; Ragini, A.; Bresciani, L.; Arzu, A.; Pinelli, D. Valorisation of olive mill wastewater by phenolic compounds adsorption: Development and application of a procedure for adsorbent selection. Chem. Eng. J. 2019, 360, 124–138. [Google Scholar] [CrossRef]
  49. Al Bsoul, A.; Hailat, M.; Abdelhay, A.; Tawalbeh, M.; Jum’h, I.; Bani-Melhem, K. Treatment of olive mill effluent by adsorption on titanium oxide nanoparticles. Sci. Total Environ. 2019, 688, 1327–1334. [Google Scholar] [CrossRef] [PubMed]
  50. Vavouraki, A.I.; Dareioti, M.A.; Kornaros, M. Olive Mill Wastewater (OMW) Polyphenols Adsorption onto Polymeric Resins: Part I—Batch Anaerobic Digestion of OMW. Waste Biomass Valorization 2020, 12, 2271–2281. [Google Scholar] [CrossRef]
  51. Ochando-Pulido, J.M.; Vellido-Pérez, J.A.; González-Hernández, R.; Martínez-Férez, A. Optimization and modeling of two-phase olive-oil washing wastewater integral treatment and phenolic compounds recovery by novel weak-base ion exchange resins. Sep. Purif. Technol. 2020, 249, 117084. [Google Scholar] [CrossRef]
  52. Ochando-Pulido, J.M.; González-Hernández, R.; Martinez-Ferez, A. On the Effect of the Operating Parameters for Two-Phase Olive-Oil Washing Wastewater Combined Phenolic Compounds Recovery and Reclamation by Novel Ion Exchange Resins. Sep. Purif. Technol. 2018, 195, 50–59. [Google Scholar] [CrossRef]
  53. Sarika, R.; Kalogerakis, N.; Mantzavinos, D. Treatment of olive mill effluents: Part II. Complete removal of solids by direct flocculation with poly-electrolytes. Environ. Int. 2005, 31, 297–304. [Google Scholar] [CrossRef]
  54. Alver, A.; Baştürk, E.; Kılıç, A.; Karataş, M. Use of advance oxidation process to improve the biodegradability of olive oil mill effluents. Process Saf. Environ. Prot. 2015, 98, 319–324. [Google Scholar] [CrossRef]
  55. Uğurlu, M.; Kula, İ. Decolourization and removal of some organic compounds from olive mill wastewater by advanced oxidation processes and lime treatment. Environ. Sci. Pollut. Res. Int. 2007, 14, 319–325. [Google Scholar] [CrossRef] [PubMed]
  56. Chatzisymeon, E.; Xekoukoulotakis, N.P.; Mantzavinos, D. Determination of key operating conditions for the photocatalytic treatment of olive mill wastewaters. Catal. Today 2009, 144, 143–148. [Google Scholar] [CrossRef] [Green Version]
  57. Maduna, K.; Kumar, N.; Aho, A.; Wärnå, J.; Zrnčević, S.; Murzin, D.Y. Kinetics of Catalytic Wet Peroxide Oxidation of Phenolics in Olive Oil Mill Wastewaters over Copper Catalysts. ACS Omega 2018, 3, 7247–7260. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  58. Najjar, W.; Azabou, S.; Sayadi, S.; Ghorbel, A. Screening of Fe-BEA catalysts for wet hydrogen peroxide oxidation of crude olive mill wastewater under mild conditions. Appl. Catal. B: Environ. 2009, 88, 299–304. [Google Scholar] [CrossRef]
  59. Azabou, S.; Najjar, W.; Bouaziz, M.; Ghorbel, A.; Sayadi, S. A compact process for the treatment of olive mill wastewater by combining wet hydrogen peroxide catalytic oxidation and biological techniques. J. Hazard. Mater. 2010, 183, 62–69. [Google Scholar] [CrossRef]
  60. Kallel, M.; Belaid, C.; Boussahel, R.; Ksibi, M.; Montiel, A.; Elleuch, B. Olive mill wastewater degradation by Fenton oxidation with zero-valent iron and hydrogen peroxide. J. Hazard. Mater. 2009, 163, 550–554. [Google Scholar] [CrossRef]
  61. Maamir, W.; Ouahabi, Y.; Poncin, S.; Li, H.-Z.; Bensadok, K. Effect of Fenton pretreatment on anaerobic digestion of olive mill wastewater and olive mill solid waste in mesophilic conditions. Int. J. Green Energy 2017, 14, 555–560. [Google Scholar] [CrossRef]
  62. Esteves, B.M.; Rodrigues, C.S.D.; Madeira, L.M. Synthetic olive mill wastewater treatment by Fenton’s process in batch and continuous reactors operation. Environ. Sci. Pollut. Res. 2018, 25, 34826–34838. [Google Scholar] [CrossRef]
  63. Hodaifa, G.; Ochando-Pulido, J.M.; Rodriguez-Vives, S.; Martinez-Ferez, A. Optimization of continuous reactor at pilot scale for olive-oil mill wastewater treatment by Fenton-like process. Chem. Eng. J. 2013, 220, 117–124. [Google Scholar] [CrossRef]
  64. Nieto, L.M.; Hodaifa, G.; Rodríguez, S.; Giménez, J.A.; Ochando, J. Degradation of organic matter in olive-oil mill wastewater through homogeneous Fenton-like reaction. Chem. Eng. J. 2011, 173, 503–510. [Google Scholar] [CrossRef]
  65. Hodaifa, G.; García, C.A.; Borja, R. Study of Catalysts’ Influence on Photocatalysis/Photodegradation of Olive Oil Mill Wastewater. Determination of the Optimum Working Conditions. Catalysts 2020, 10, 554. [Google Scholar] [CrossRef]
  66. Adhoum, N.; Monser, L. Decolourization and removal of phenolic compounds from olive mill wastewater by electrocoagulation. Chem. Eng. Processing Process Intensif. 2004, 43, 1281–1287. [Google Scholar] [CrossRef]
  67. Elkacmi, R.; Boudouch, O.; Hasib, A.; Bouzaid, M.; Bennajah, M. Photovoltaic electrocoagulation treatment of olive mill wastewater using an external-loop airlift reactor. Sustain. Chem. Pharm. 2020, 17, 100274. [Google Scholar] [CrossRef]
  68. D’Annibale, A.; Stazi, S.R.; Vinciguerra, V.; Sermanni, G.G. Oxirane-immobilized Lentinula edodes laccase: Stability and phenolics removal efficiency in olive mill wastewater. J. Biotechnol. 2000, 77, 265–273. [Google Scholar] [CrossRef]
  69. Casademont, P.; García-Jarana, M.B.; Sánchez-Oneto, J.; Portela, J.R.; de la Ossa, E.J.M. Hydrogen production by catalytic conversion of olive mill wastewater in supercritical water. J. Supercrit. Fluids 2018, 141, 224–229. [Google Scholar] [CrossRef]
  70. García-Gómez, A.; Roig, A.; Bernal, M.P. Composting of the solid fraction of olive mill wastewater with olive leaves: Organic matter degradation and biological activity. Bioresour. Technol. 2003, 86, 59–64. [Google Scholar] [CrossRef]
  71. Fadil, K.; Chahlaoui, A.; Ouahbi, A.; Zaid, A.; Borja, R. Aerobic biodegradation and detoxification of wastewaters from the olive oil industry. Int. Biodeterior. Biodegrad. 2003, 51, 37–41. [Google Scholar] [CrossRef]
  72. Beltran-Heredia, J.; Torregrosa, J.; Dominguez, J.R.; Garcia, J. Treatment of black-olive wastewaters by ozonation and aerobic biological degradation. Water Res. 2000, 34, 3515–3522. [Google Scholar] [CrossRef]
  73. Hajjouji, H.E.; Bailly, J.R.; Winterton, P.; Merlina, G.; Revel, J.C.; Hafidi, M. Chemical and spectroscopic analysis of olive mill waste water during a biological treatment. Bioresour. Technol. 2008, 99, 4958–4965. [Google Scholar] [CrossRef] [Green Version]
  74. Bertin, L.; Berselli, S.; Fava, F.; Petrangeli-Papini, M.; Marchetti, L. Anaerobic digestion of olive mill wastewaters in biofilm reactors packed with granular activated carbon and “Manville” silica beads. Water Res. 2004, 38, 3167–3178. [Google Scholar] [CrossRef]
  75. Gonçalves, M.R.; Costa, J.C.; Pereira, M.A.; Abreu, A.A.; Alves, M.M. On the independence of hydrogen production from methanogenic suppressor in olive mill wastewater. Int. J. Hydrogen Energy 2014, 39, 6402–6406. [Google Scholar] [CrossRef] [Green Version]
  76. Eroğlu, E.; Eroğlu, İ.; Gündüz, U.; Türker, L.; Yücel, M. Biological hydrogen production from olive mill wastewater with two-stage processes. Int. J. Hydrogen Energy 2006, 31, 1527–1535. [Google Scholar] [CrossRef]
  77. Caporaso, N.; Formisano, D.; Genovese, A. Use of phenolic compounds from olive mill wastewater as valuable ingredients for functional foods. Crit. Rev. Food Sci. Nutr. 2018, 58, 2829–2841. [Google Scholar] [CrossRef] [PubMed]
  78. Agalias, A.; Magiatis, P.; Skaltsounis, A.-L.; Mikros, E.; Tsarbopoulos, A.; Gikas, E.; Spanos, I.; Manios, T. A New Process for the Management of Olive Oil Mill Waste Water and Recovery of Natural Antioxidants. J. Agric. Food Chem. 2007, 55, 2671–2676. [Google Scholar] [CrossRef] [PubMed]
  79. Gonçalves, C.; Lopes, M.; Ferreira, J.P.; Belo, I. Biological treatment of olive mill wastewater by non-conventional yeasts. Bioresour. Technol. 2009, 100, 3759–3763. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  80. Servili, M.; Esposto, S.; Veneziani, G.; Urbani, S.; Taticchi, A.; di Maio, I.; Selvaggini, R.; Sordini, B.; Montedoro, G. Improvement of bioactive phenol content in virgin olive oil with an olive-vegetation water concentrate produced by membrane treatment. Food Chem. 2011, 124, 1308–1315. [Google Scholar] [CrossRef]
  81. Mazzei, R.; Drioli, E.; Giorno, L. Enzyme membrane reactor with heterogenized β-glucosidase to obtain phytotherapic compound: Optimization study. J. Membr. Sci. 2012, 390–391 (Suppl. C), 121–129. [Google Scholar] [CrossRef]
  82. Kestioğlu, K.; Yonar, T.; Azbar, N. Feasibility of physico-chemical treatment and Advanced Oxidation Processes (AOPs) as a means of pretreatment of olive mill effluent (OME). Process Biochem. 2005, 40, 2409–2416. [Google Scholar] [CrossRef]
  83. Drouiche, M.; le Mignot, V.; Lounici, H.; Belhocine, D.; Grib, H.; Pauss, A.; Mameri, N. A compact process for the treatment of olive mill wastewater by combining OF and UV/H2O2 techniques. Desalination 2004, 169, 81–88. [Google Scholar] [CrossRef]
  84. Leal, A.L.; Soria, M.A.; Madeira, L.M. Autothermal reforming of impure glycerol for H2 production: Thermodynamic study including in situ CO2 and/or H2 separation. Int. J. Hydrogen Energy 2016, 41, 2607–2620. [Google Scholar] [CrossRef]
  85. Rocha, C.; Soria, M.A.; Madeira, L.M. Screening of commercial catalysts for steam reforming of olive mill wastewater. Renew. Energy 2021, 169, 765–779. [Google Scholar] [CrossRef]
  86. Agency, I.E. Hydrogen. 2021. Available online: https://www.iea.org/reports/hydrogen (accessed on 29 November 2021).
  87. Agency, I.E. Global Hydrogen Demand by Sector in the Net Zero Scenario, 2020–2030. 2021. Available online: https://www.iea.org/data-and-statistics/charts/global-hydrogen-demand-by-sector-in-the-net-zero-scenario-2020-2030 (accessed on 29 November 2021).
  88. Cakiryilmaz, N.; Arbag, H.; Oktar, N.; Dogu, G.; Dogu, T. Effect of W incorporation on the product distribution in steam reforming of bio-oil derived acetic acid over Ni based Zr-SBA-15 catalyst. Int. J. Hydrogen Energy 2018, 43, 3629–3642. [Google Scholar] [CrossRef]
  89. Tosti, S.; Cavezza, C.; Fabbricino, M.; Pontoni, L.; Palma, V.; Ruocco, C. Production of hydrogen in a Pd-membrane reactor via catalytic reforming of olive mill wastewater. Chem. Eng. J. 2015, 275 (Suppl. C), 366–373. [Google Scholar] [CrossRef]
  90. Alique, D.; Bruni, G.; Sanz, R.; Calles, J.A.; Tosti, S. Ultra-Pure Hydrogen via Co-Valorization of Olive Mill Wastewater and Bioethanol in Pd-Membrane Reactors. Processes 2020, 8, 219. [Google Scholar] [CrossRef] [Green Version]
  91. Rocha, C.; Soria, M.A.; Madeira, L.M. Use of Ni-containing catalysts for synthetic olive mill wastewater steam reforming. Renew. Energy 2022, 185, 1329–1342. [Google Scholar] [CrossRef]
  92. Zhang, Z.; Zhang, L.; Liu, F.; Sun, Y.; Shao, Y.; Sun, K.; Zhang, S.; Liu, Q.; Hu, G.; Hu, X. Tailoring the surface properties of Ni/SiO2 catalyst with sulfuric acid for enhancing the catalytic efficiency for steam reforming of guaiacol. Renew. Energy 2020, 156, 423–439. [Google Scholar] [CrossRef]
  93. Zhang, Z.; Sun, Y.; Wang, Y.; Sun, K.; Gao, Z.; Xu, Q.; Zhang, S.; Hu, G.; Xu, L.; Hu, X. Steam reforming of acetic acid and guaiacol over Ni/Attapulgite catalyst: Tailoring pore structure of the catalyst with KOH activation for enhancing the resistivity towards coking. Mol. Catal. 2020, 493, 111051. [Google Scholar] [CrossRef]
  94. Trane, R.; Dahl, S.; Skjøth-Rasmussen, M.S.; Jensen, A.D. Catalytic steam reforming of bio-oil. Int. J. Hydrogen Energy 2012, 37, 6447–6472. [Google Scholar] [CrossRef]
  95. Xie, H.; Yu, Q.; Wei, M.; Duan, W.; Yao, X.; Qin, Q.; Zuo, Z. Hydrogen production from steam reforming of simulated bio-oil over Ce-Ni/Co catalyst with in continuous CO2 capture. Int. J. Hydrogen Energy 2015, 40, 1420–1428. [Google Scholar] [CrossRef]
  96. Soria, M.A.; Barros, D.; Madeira, L.M. Hydrogen production through steam reforming of bio-oils derived from biomass pyrolysis: Thermodynamic analysis including in situ CO2 and/or H2 separation. Fuel 2019, 244, 184–195. [Google Scholar] [CrossRef]
  97. Kechagiopoulos, P.N.; Voutetakis, S.S.; Lemonidou, A.A.; Vasalos, I.A. Hydrogen Production via Steam Reforming of the Aqueous Phase of Bio-Oil in a Fixed Bed Reactor. Energy Fuels 2006, 20, 2155–2163. [Google Scholar] [CrossRef]
  98. Remiro, A.; Valle, B.; Aguayo, A.T.; Bilbao, J.; Gayubo, A.G. Operating conditions for attenuating Ni/La2O3-αAl2O3 catalyst deactivation in the steam reforming of bio-oil aqueous fraction. Fuel Processing Technol. 2013, 115, 222–232. [Google Scholar] [CrossRef]
  99. Remiro, A.; Arandia, A.; Oar-Arteta, L.; Bilbao, J.; Gayubo, A.G. Regeneration of NiAl2O4 spinel type catalysts used in the reforming of raw bio-oil. Appl. Catal. B: Environ. 2018, 237, 353–365. [Google Scholar] [CrossRef]
  100. Valle, B.; Aramburu, B.; Olazar, M.; Bilbao, J.; Gayubo, A.G. Steam reforming of raw bio-oil over Ni/La2O3-αAl2O3: Influence of temperature on product yields and catalyst deactivation. Fuel 2018, 216, 463–474. [Google Scholar] [CrossRef]
  101. Yu, N.; Rahman, M.M.; Chen, J.; Sun, J.; Engelhard, M.; Hernandez, X.I.P.; Wang, Y. Steam reforming of simulated bio-oil on K-Ni-Cu-Mg-Ce-O/Al2O3: The effect of K. Catal. Today 2019, 323, 183–190. [Google Scholar] [CrossRef]
  102. Cheng, Y.W.; Khan, M.R.; Ng, K.H.; Wongsakulphasatch, S.; Cheng, C.K. Harnessing renewable hydrogen-rich syngas from valorization of palm oil mill effluent (POME) using steam reforming technique. Renew. Energy 2019, 138, 1114–1126. [Google Scholar] [CrossRef]
  103. Cheng, Y.W.; Chong, C.C.; Lee, S.P.; Lim, J.W.; Wu, T.Y.; Cheng, C.K. Syngas from palm oil mill effluent (POME) steam reforming over lanthanum cobaltite: Effects of net-basicity. Renew. Energy 2019, 148, 349–362. [Google Scholar] [CrossRef]
  104. Medrano, J.A.; Oliva, M.; Ruiz, J.; Garcia, L.; Arauzo, J. Catalytic steam reforming of acetic acid in a fluidized bed reactor with oxygen addition. Int. J. Hydrogen Energy 2008, 33, 4387–4396. [Google Scholar] [CrossRef]
  105. Matas Güell, B.; ISilva, M.T.d.; Seshan, K.; Lefferts, L. Sustainable route to hydrogen—Design of stable catalysts for the steam gasification of biomass related oxygenates. Appl. Catal. B: Environ. 2009, 88, 59–65. [Google Scholar] [CrossRef]
  106. Hoang, T.M.C.; Geerdink, B.; Sturm, J.M.; Lefferts, L.; Seshan, K. Steam reforming of acetic acid—A major component in the volatiles formed during gasification of humin. Appl. Catal. B Environ. 2015, 163 (Suppl. C), 74–82. [Google Scholar] [CrossRef]
  107. Chen, G.; Tao, J.; Liu, C.; Yan, B.; Li, W.; Li, X. Steam reforming of acetic acid using Ni/Al2O3 catalyst: Influence of crystalline phase of Al2O3 support. Int. J. Hydrogen Energy 2017, 42, 20729–20738. [Google Scholar] [CrossRef]
  108. Savuto, E.; Navarro, R.M.; Mota, N.; di Carlo, A.; Bocci, E.; Carlini, M.; Fierro, J.L.G. Steam reforming of tar model compounds over Ni/Mayenite catalysts: Effect of Ce addition. Fuel 2018, 224, 676–686. [Google Scholar] [CrossRef]
  109. Silva, J.M.; Soria, M.A.; Madeira, L.M. Challenges and strategies for optimization of glycerol steam reforming process. Renew. Sustain. Energy Rev. 2015, 42, 1187–1213. [Google Scholar] [CrossRef] [Green Version]
  110. Zhang, Z.; Hu, X.; Li, J.; Gao, G.; Dong, D.; Westerhof, R.; Hu, S.; Xiang, J.; Wang, Y. Steam reforming of acetic acid over Ni/Al2O3 catalysts: Correlation of nickel loading with properties and catalytic behaviors of the catalysts. Fuel 2018, 217, 389–403. [Google Scholar] [CrossRef]
  111. Borges, R.P.; Ferreira, R.A.R.; Rabelo-Neto, R.C.; Noronha, F.B.; Hori, C.E. Hydrogen production by steam reforming of acetic acid using hydrotalcite type precursors. Int. J. Hydrogen Energy 2018, 43, 7881–7892. [Google Scholar] [CrossRef]
  112. Yu, Z.; Hu, X.; Jia, P.; Zhang, Z.; Dong, D.; Hu, G.; Hu, S.; Wang, Y.; Xiang, J. Steam reforming of acetic acid over nickel-based catalysts: The intrinsic effects of nickel precursors on behaviors of nickel catalysts. Appl. Catal. B Environ. 2018, 237, 538–553. [Google Scholar] [CrossRef]
  113. Matas Güell, B.; Babich, I.; Nichols, K.P.; Gardeniers, J.G.E.; Lefferts, L.; Seshan, K. Design of a stable steam reforming catalyst—A promising route to sustainable hydrogen from biomass oxygenates. Appl. Catal. B Environ. 2009, 90, 38–44. [Google Scholar] [CrossRef]
  114. de Castro, T.P.; Silveira, E.B.; Rabelo-Neto, R.C.; Borges, L.E.P.; Noronha, F.B. Study of the performance of Pt/Al2O3 and Pt/CeO2/Al2O3 catalysts for steam reforming of toluene, methane and mixtures. Catal. Today 2018, 299 (Suppl. C), 251–262. [Google Scholar] [CrossRef]
  115. Pant, K.K.; Mohanty, P.; Agarwal, S.; Dalai, A.K. Steam reforming of acetic acid for hydrogen production over bifunctional Ni-Co catalysts. Catal. Today 2013, 207 (Suppl. C), 36–43. [Google Scholar] [CrossRef]
  116. Polychronopoulou, K.; Fierro, J.L.G.; Efstathiou, A.M. The phenol steam reforming reaction over MgO-based supported Rh catalysts. J. Catal. 2004, 228, 417–432. [Google Scholar] [CrossRef]
  117. Constantinou, D.A.; Efstathiou, A.M. Low-temperature purification of gas streams from phenol by steam reforming over novel supported-Rh catalysts. Appl. Catal. B Environ. 2010, 96, 276–289. [Google Scholar] [CrossRef]
  118. Iida, H.; Fujiyama, A.; Igarashi, A.; Okumura, K. Steam reforming of toluene over Ru/SrCO3-Al2O3 catalysts. Fuel Processing Technol. 2017, 168 (Suppl. C), 50–57. [Google Scholar] [CrossRef]
  119. Liang, T.; Wang, Y.; Chen, M.; Yang, Z.; Liu, S.; Zhou, Z.; Li, X. Steam reforming of phenol-ethanol to produce hydrogen over bimetallic NiCu catalysts supported on sepiolite. Int. J. Hydrogen Energy 2017, 42, 28233–28246. [Google Scholar] [CrossRef]
  120. Adnan, M.A.; Muraza, O.; Razzak, S.A.; Hossain, M.M.; de Lasa, H.I. Iron Oxide over Silica-Doped Alumina Catalyst for Catalytic Steam Reforming of Toluene as a Surrogate Tar Biomass Species. Energy Fuels 2017, 31, 7471–7481. [Google Scholar] [CrossRef]
  121. Wang, S.; Li, X.; Guo, L.; Luo, Z. Experimental research on acetic acid steam reforming over Co-Fe catalysts and subsequent density functional theory studies. Int. J. Hydrogen Energy 2012, 37, 11122–11131. [Google Scholar] [CrossRef]
  122. Furusawa, T.; Saito, K.; Kori, Y.; Miura, Y.; Sato, M.; Suzuki, N. Steam reforming of naphthalene/benzene with various types of Pt- and Ni-based catalysts for hydrogen production. Fuel 2013, 103, 111–121. [Google Scholar] [CrossRef]
  123. Wang, S.; Zhang, F.; Cai, Q.; Zhu, L.; Luo, Z. Steam reforming of acetic acid over coal ash supported Fe and Ni catalysts. Int. J. Hydrogen Energy 2015, 40, 11406–11413. [Google Scholar] [CrossRef]
  124. Nabgan, W.; Abdullah, T.A.T.; Mat, R.; Nabgan, B.; Gambo, Y.; Triwahyono, S. Influence of Ni to Co ratio supported on ZrO2 catalysts in phenol steam reforming for hydrogen production. Int. J. Hydrogen Energy 2016, 41, 22922–22931. [Google Scholar] [CrossRef]
  125. Silva, O.C.V.; Silveira, E.B.; Rabelo-Neto, R.C.; Borges, L.E.P.; Noronha, F.B. Hydrogen Production Through Steam Reforming of Toluene Over Ni Supported on MgAl Mixed Oxides Derived from Hydrotalcite-Like Compounds. Catal. Lett. 2018, 148, 1622–1633. [Google Scholar] [CrossRef]
  126. Pu, J.; Nishikado, K.; Wang, N.; Nguyen, T.T.; Maki, T.; Qian, E.W. Core-shell nickel catalysts for the steam reforming of acetic acid. Appl. Catal. B: Environ. 2018, 224, 69–79. [Google Scholar] [CrossRef]
  127. Li, J.; Mei, X.; Zhang, L.; Yu, Z.; Liu, Q.; Wei, T.; Wu, W.; Dong, D.; Xu, L.; Hu, X. A comparative study of catalytic behaviors of Mn, Fe, Co, Ni, Cu and Zn-Based catalysts in steam reforming of methanol, acetic acid and acetone. Int. J. Hydrogen Energy 2020, 45, 3815–3832. [Google Scholar] [CrossRef]
  128. Zhang, C.; Hu, X.; Yu, Z.; Zhang, Z.; Chen, G.; Li, C.; Liu, Q.; Xiang, J.; Wang, Y.; Hu, S. Steam reforming of acetic acid for hydrogen production over attapulgite and alumina supported Ni catalysts: Impacts of properties of supports on catalytic behaviors. Int. J. Hydrogen Energy 2019, 44, 5230–5244. [Google Scholar] [CrossRef]
  129. Chen, J.; Wang, M.; Wang, S.; Li, X. Hydrogen production via steam reforming of acetic acid over biochar-supported nickel catalysts. Int. J. Hydrogen Energy 2018, 43, 18160–18168. [Google Scholar] [CrossRef]
  130. Wang, Y.; Zhang, Z.; Zhang, S.; Wang, Y.; Hu, S.; Xiang, J.; Hu, X. Steam reforming of acetic acid over Ni/biochar catalyst treated with HNO3: Impacts of the treatment on surface properties and catalytic behaviors. Fuel 2020, 278, 118341. [Google Scholar] [CrossRef]
  131. Yang, X.; Wang, Y.; Li, M.; Sun, B.; Li, Y.; Wang, Y. Enhanced Hydrogen Production by Steam Reforming of Acetic Acid over a Ni Catalyst Supported on Mesoporous MgO. Energy Fuels 2016, 30, 2198–2203. [Google Scholar] [CrossRef]
  132. Bangala, D.N.; Abatzoglou, N.; Chornet, E. Steam reforming of naphthalene on Ni-Cr/Al2O3 catalysts doped with MgO, TiO2, and La2O3. AIChE J. 1998, 44, 927–936. [Google Scholar] [CrossRef]
  133. Garcia, L.a.; French, R.; Czernik, S.; Chornet, E. Catalytic steam reforming of bio-oils for the production of hydrogen: Effects of catalyst composition. Appl. Catal. A Gen. 2000, 201, 225–239. [Google Scholar] [CrossRef]
  134. Wang, M.; Zhang, F.; Wang, S. Effect of La2O3 replacement on γ-Al2O3 supported nickel catalysts for acetic acid steam reforming. Int. J. Hydrogen Energy 2017, 42, 20540–20548. [Google Scholar] [CrossRef]
  135. Li, L.; Jiang, B.; Tang, D.; Zhang, Q.; Zheng, Z. Hydrogen generation by acetic acid steam reforming over Ni-based catalysts derived from La1−xCexNiO3 perovskite. Int. J. Hydrogen Energy 2018, 43, 6795–6803. [Google Scholar] [CrossRef]
  136. Matas Güell, B.; Babich, I.V.; Lefferts, L.; Seshan, K. Steam reforming of phenol over Ni-based catalysts—A comparative study. Appl. Catal. B Environ. 2011, 106, 280–286. [Google Scholar] [CrossRef]
  137. Frusteri, F.; Freni, S.; Chiodo, V.; Spadaro, L.; di Blasi, O.; Bonura, G.; Cavallaro, S. Steam reforming of bio-ethanol on alkali-doped Ni/MgO catalysts: Hydrogen production for MC fuel cell. Appl. Catal. A Gen. 2004, 270, 1–7. [Google Scholar] [CrossRef]
  138. Hu, X.; Dong, D.; Shao, X.; Zhang, L.; Lu, G. Steam reforming of acetic acid over cobalt catalysts: Effects of Zr, Mg and K addition. Int. J. Hydrogen Energy 2017, 42, 4793–4803. [Google Scholar] [CrossRef]
  139. Ahmed, T.; Xiu, S.; Wang, L.; Shahbazi, A. Investigation of Ni/Fe/Mg zeolite-supported catalysts in steam reforming of tar using simulated-toluene as model compound. Fuel 2018, 211 (Suppl. C), 566–571. [Google Scholar] [CrossRef]
  140. Pu, J.; Ikegami, F.; Nishikado, K.; Qian, E.W. Effect of ceria addition on NiRu/CeO2Al2O3 catalysts in steam reforming of acetic acid. Int. J. Hydrogen Energy 2017, 42, 19733–19743. [Google Scholar] [CrossRef]
  141. Wang, Y.; Chen, M.; Yang, J.; Liu, S.; Yang, Z.; Wang, J.; Liang, T. Hydrogen Production from Steam Reforming of Acetic Acid over Ni-Fe/Palygorskite Modified with Cerium. BioResources 2017, 12, 4830–4853. [Google Scholar] [CrossRef] [Green Version]
  142. Zhao, X.; Xue, Y.; Lu, Z.; Huang, Y.; Guo, C.; Yan, C. Encapsulating Ni/CeO2-ZrO2 with SiO2 layer to improve it catalytic activity for steam reforming of toluene. Catal. Commun. 2017, 101 (Suppl. C), 138–141. [Google Scholar] [CrossRef]
  143. Li, L.; Tang, D.; Song, Y.; Jiang, B.; Zhang, Q. Hydrogen production from ethanol steam reforming on Ni-Ce/MMT catalysts. Energy 2018, 149, 937–943. [Google Scholar] [CrossRef]
  144. Zhang, Z.; Hu, X.; Gao, G.; Wei, T.; Dong, D.; Wang, Y.; Hu, S.; Xiang, J.; Liu, Q.; Geng, D. Steam reforming of acetic acid over NiKOH/Al2O3 catalyst with low nickel loading: The remarkable promotional effects of KOH on activity. Int. J. Hydrogen Energy 2018, 44, 729–747. [Google Scholar] [CrossRef]
  145. Choi, I.-H.; Hwang, K.-R.; Lee, K.-Y.; Lee, I.-G. Catalytic steam reforming of biomass-derived acetic acid over modified Ni/γ-Al2O3 for sustainable hydrogen production. Int. J. Hydrogen Energy 2019, 44, 180–190. [Google Scholar] [CrossRef]
  146. Higo, T.; Saito, H.; Ogo, S.; Sugiura, Y.; Sekine, Y. Promotive effect of Ba addition on the catalytic performance of Ni/LaAlO3 catalysts for steam reforming of toluene. Appl. Catal. A Gen. 2017, 530, 125–131. [Google Scholar] [CrossRef]
  147. Fally, F.; Perrichon, V.; Vidal, H.; Kaspar, J.; Blanco, G.; Pintado, J.M.; Bernal, S.; Colon, G.; Daturi, M.; Lavalley, J.C. Modification of the oxygen storage capacity of CeO2-ZrO2 mixed oxides after redox cycling aging. Catal. Today 2000, 59, 373–386. [Google Scholar] [CrossRef]
  148. Vidal, H.; Kašpar, J.; Pijolat, M.; Colon, G.; Bernal, S.; Cordón, A.; Perrichon, V.; Fally, F. Redox behavior of CeO2-ZrO2 mixed oxides: II. Influence of redox treatments on low surface area catalysts. Appl. Catal. B Environ. 2001, 30, 75–85. [Google Scholar] [CrossRef]
  149. Menendez, R.B.; Graschinsky, C.; Amadeo, N.E. Sorption-Enhanced Ethanol Steam Reforming Process in a Fixed-Bed Reactor. Ind. Eng. Chem. Res. 2018, 57, 11547–11553. [Google Scholar] [CrossRef]
  150. Chen, M.; Wang, Y.; Yang, Z.; Liang, T.; Liu, S.; Zhou, Z.; Li, X. Effect of Mg-modified mesoporous Ni/Attapulgite catalysts on catalytic performance and resistance to carbon deposition for ethanol steam reforming. Fuel 2018, 220, 32–46. [Google Scholar] [CrossRef]
  151. Kumar, A.; Sinha, A.S.K. Hydrogen production from acetic acid steam reforming over nickel-based catalyst synthesized via MOF process. Int. J. Hydrogen Energy 2020, 45, 24397–24411. [Google Scholar] [CrossRef]
  152. Yu, Z.; Zhang, L.; Zhang, C.; Gao, G.; Ye, Z.; Zhang, S.; Liu, Q.; Hu, G.; Hu, X. Steam reforming of acetic acid over nickel catalysts: Impacts of fourteen additives on the catalytic behaviors. J. Energy Inst. 2020, 93, 1000–1019. [Google Scholar] [CrossRef]
  153. Galdámez, J.R.; García, L.; Bilbao, R. Hydrogen Production by Steam Reforming of Bio-Oil Using Coprecipitated Ni−Al Catalysts. Acetic Acid as a Model Compound. Energy Fuels 2005, 19, 1133–1142. [Google Scholar] [CrossRef]
  154. Rostrup-Nielsen, J.R.; Sehested, J.; Nørskov, J.K. Hydrogen and synthesis gas by steam- and CO2 reforming. In Advances in Catalysis; Academic Press: Cambridge, MA, USA, 2002; Volume 47, pp. 65–139. [Google Scholar]
  155. Bengaard, H.S.; Nørskov, J.K.; Sehested, J.; Clausen, B.S.; Nielsen, L.P.; Molenbroek, A.M.; Rostrup-Nielsen, J.R. Steam Reforming and Graphite Formation on Ni Catalysts. J. Catal. 2002, 209, 365–384. [Google Scholar] [CrossRef]
  156. Pekmezci Karaman, B.; Cakiryilmaz, N.; Arbag, H.; Oktar, N.; Dogu, G.; Dogu, T. Performance comparison of mesoporous alumina supported Cu & Ni based catalysts in acetic acid reforming. Int. J. Hydrogen Energy 2017, 42, 26257–26269. [Google Scholar]
  157. Goicoechea, S.; Kraleva, E.; Sokolov, S.; Schneider, M.; Pohl, M.-M.; Kockmann, N.; Ehrich, H. Support effect on structure and performance of Co and Ni catalysts for steam reforming of acetic acid. Appl. Catal. A Gen. 2016, 514 (Suppl. C), 182–191. [Google Scholar] [CrossRef]
  158. Meng, J.; Zhao, Z.; Wang, X.; Zheng, A.; Zhang, D.; Huang, Z.; Zhao, K.; Wei, G.; Li, H. Comparative study on phenol and naphthalene steam reforming over Ni-Fe alloy catalysts supported on olivine synthesized by different methods. Energy Convers. Manag. 2018, 168, 60–73. [Google Scholar] [CrossRef]
  159. Zhu, H.L.; Pastor-Pérez, L.; Millan, M. Catalytic Steam Reforming of Toluene: Understanding the Influence of the Main Reaction Parameters over a Reference Catalyst. Energies 2020, 13, 813. [Google Scholar] [CrossRef] [Green Version]
  160. Rodrigues, A.E.; Madeira, L.M.; Faria, R.; Wu, Y.J. Sorption Enhanced Reaction Processes; World Scientific Publishing Company Pte Limited: Singapore, 2017. [Google Scholar]
  161. Ni, F.; Caram, H.S. Sorption enhanced reaction for high purity products in reversible reactions. AIChE J. 2017, 63, 5452–5461. [Google Scholar] [CrossRef]
  162. Wu, Y.-J.; Li, P.; Yu, J.-G.; Cunha, A.F.; Rodrigues, A.E. Sorption-enhanced steam reforming of ethanol for continuous high-purity hydrogen production: 2D adsorptive reactor dynamics and process design. Chem. Eng. Sci. 2014, 118, 83–93. [Google Scholar] [CrossRef]
  163. Soria, M.A.; Tosti, S.; Mendes, A.; Madeira, L.M. Enhancing the low temperature water-gas shift reaction through a hybrid sorption-enhanced membrane reactor for high-purity hydrogen production. Fuel 2015, 159, 854–863. [Google Scholar] [CrossRef] [Green Version]
  164. Iliuta, I.; Radfarnia, H.R.; Iliuta, M.C. Hydrogen Production by Sorption-Enhanced Steam Glycerol Reforming: Sorption Kinetics and Reactor Simulation. AIChE J. 2013, 59, 2105–2118. [Google Scholar] [CrossRef]
  165. Clough, P.T.; Boot-Handford, M.E.; Zheng, L.; Zhang, Z.; Fennell, P.S. Hydrogen Production by Sorption Enhanced Steam Reforming (SESR) of Biomass in a Fluidised-Bed Reactor Using Combined Multifunctional Particles. Materials 2018, 11, 859. [Google Scholar] [CrossRef] [Green Version]
  166. Park, B.-G. A hybrid adsorbent-membrane reactor (HAMR) system for hydrogen production. Korean J. Chem. Eng. 2004, 21, 782–792. [Google Scholar] [CrossRef]
  167. Wu, Y.J.; Alvarado, F.D.; Santos, J.C.; Gracia, F.; Cunha, A.F.; Rodrigues, A.E. Sorption-Enhanced Steam Reforming of Ethanol: Thermodynamic Comparison of CO2 Sorbents. Chem. Eng. Technol. 2012, 35, 847–858. [Google Scholar] [CrossRef]
  168. Dou, B.; Dupont, V.; Rickett, G.; Blakeman, N.; Williams, P.T.; Chen, H.; Ding, Y.; Ghadiri, M. Hydrogen production by sorption-enhanced steam reforming of glycerol. Bioresour. Technol. 2009, 100, 3540–3547. [Google Scholar] [CrossRef] [Green Version]
  169. Martavaltzi, C.S.; Lemonidou, A.A. Hydrogen production via sorption enhanced reforming of methane: Development of a novel hybrid material—reforming catalyst and CO2 sorbent. Chem. Eng. Sci. 2010, 65, 4134–4140. [Google Scholar] [CrossRef]
  170. Aloisi, I.; Adi, G.; di Carlo, A.; Foscolo, P.U.; Courson, C.; Gallucci, K. Sorption enhanced catalytic Steam Methane Reforming: Experimental data and simulations describing the behaviour of bi-functional particles. Chem. Eng. J. 2017, 314, 570–582. [Google Scholar] [CrossRef]
  171. Reijers, H.T.J.; Valster-Schiermeier, S.E.A.; Cobden, P.D.; van den Brink, R.W. Hydrotalcite as CO2 Sorbent for Sorption-Enhanced Steam Reforming of Methane. Ind. Eng. Chem. Res. 2006, 45, 2522–2530. [Google Scholar] [CrossRef]
  172. Kim, C.-H.; Han, J.-Y.; Lim, H.; Lee, K.-Y.; Ryi, S.-K. Hydrogen production by steam methane reforming in membrane reactor equipped with Pd membrane deposited on NiO/YSZ/NiO multilayer-treated porous stainless steel. J. Membr. Sci. 2018, 563, 75–82. [Google Scholar] [CrossRef]
  173. Soria, M.A.; Mateos-Pedrero, C.; Rodríguez-Ramos, I.; Guerrero-Ruiz, A. Catalytic steam reforming of methane under conditions of applicability with Pd membranes over supported Ru catalysts. Catal. Today 2011, 171, 126–131. [Google Scholar] [CrossRef]
  174. Adiya, Z.I.S.G.; Dupont, V.; Mahmud, T. Steam reforming of shale gas with nickel and calcium looping. Fuel 2019, 237, 142–151. [Google Scholar] [CrossRef]
  175. Omoniyi, O.A.; Dupont, V. Optimised cycling stability of sorption enhanced chemical looping steam reforming of acetic acid in a packed bed reactor. Appl. Catal. B Environ. 2019, 242, 397–409. [Google Scholar] [CrossRef]
  176. Diglio, G.; Bareschino, P.; Mancusi, E.; Pepe, F.; Montagnaro, F.; Hanak, D.P.; Manovic, V. Feasibility of CaO/CuO/NiO sorption-enhanced steam methane reforming integrated with solid-oxide fuel cell for near-zero-CO2 emissions cogeneration system. Appl. Energy 2018, 230, 241–256. [Google Scholar] [CrossRef] [Green Version]
  177. Xie, H.; Zhang, W.; Zhao, X.; Chen, H.; Yu, Q.; Qin, Q. Sorption-enhanced reforming of tar: Influence of the preparation method of CO2 absorbent. Korean J. Chem. Eng. 2018, 35, 2191–2197. [Google Scholar] [CrossRef]
  178. Di Felice, L.; Kazi, S.S.; Sørby, M.H.; Martinez, I.; Grasa, G.; Maury, D.; Meyer, J. Combined sorbent and catalyst material for sorption enhanced reforming of methane under cyclic regeneration in presence of H2O and CO2. Fuel Processing Technol. 2019, 183, 35–47. [Google Scholar] [CrossRef]
  179. Di Giuliano, A.; Giancaterino, F.; Gallucci, K.; Foscolo, P.U.; Courson, C. Catalytic and sorbent materials based on mayenite for sorption enhanced steam methane reforming with different packed-bed configurations. Int. J. Hydrogen Energy 2018, 43, 21279–21289. [Google Scholar] [CrossRef]
  180. Ghungrud, S.A.; Dewoolkar, K.D.; Vaidya, P.D. Cerium-promoted bi-functional hybrid materials made of Ni, Co and hydrotalcite for sorption-enhanced steam methane reforming (SESMR). Int. J. Hydrogen Energy 2018, 44, 694–706. [Google Scholar] [CrossRef]
  181. Dang, C.; Wu, S.; Cao, Y.; Wang, H.; Peng, F.; Yu, H. Co-production of high quality hydrogen and synthesis gas via sorption-enhanced steam reforming of glycerol coupled with methane reforming of carbonates. Chem. Eng. J. 2019, 360, 47–53. [Google Scholar] [CrossRef]
  182. Cherbański, R.; Molga, E. Sorption-enhanced steam methane reforming (SE-SMR)—A review: Reactor types, catalysts and sorbents characterization, process modelling. Chem. Process Eng. 2018, 39, 427–448. [Google Scholar]
  183. Jang, H.M.; Lee, K.B.; Caram, H.S.; Sircar, S. High-purity hydrogen production through sorption enhanced water gas shift reaction using K2CO3-promoted hydrotalcite. Chem. Eng. Sci. 2012, 73, 431–438. [Google Scholar] [CrossRef]
  184. Xie, H.; Yu, Q.; Lu, H.; Ji, L.; Chen, H.; Qin, Q. Selection and preparation of CO2 sorbent for sorption-enhanced steam reforming process of raw coke oven gas. Environ. Prog. Sustain. Energy 2019, 38, 89–97. [Google Scholar] [CrossRef] [Green Version]
  185. Soria, M.A.; Rocha, C.; Tosti, S.; Mendes, A.; Madeira, L.M. COx free hydrogen production through water-gas shift reaction in different hybrid multifunctional reactors. Chem. Eng. J. 2019, 356, 727–736. [Google Scholar] [CrossRef]
  186. Vanga, G.; Gattia, D.M.; Stendardo, S.; Scaccia, S. Novel synthesis of combined CaO-Ca12Al14O33-Ni sorbent-catalyst material for sorption enhanced steam reforming processes. Ceram. Int. 2019, 45, 7594–7605. [Google Scholar] [CrossRef]
  187. Liu, L.; Hong, D.; Wang, N.; Guo, X. High purity H2 production from sorption enhanced bio-ethanol reforming via sol-gel-derived Ni-CaO-Al2O3 bi-functional materials. Int. J. Hydrogen Energy 2020, 45, 34449–34460. [Google Scholar] [CrossRef]
  188. Wang, X.; He, Y.; Xu, T.; Xiao, B.; Liu, S.; Hu, Z.; Li, J. CO2 sorption-enhanced steam reforming of phenol using Ni-M/CaO-Ca12Al14O33 (M = Cu, Co, and Ce) as catalytic sorbents. Chem. Eng. J. 2020, 393, 124769. [Google Scholar] [CrossRef]
  189. Dang, C.; Wu, S.; Yang, G.; Cao, Y.; Wang, H.; Peng, F.; Wang, S.; Yu, H. Hydrogen Production from Sorption-Enhanced Steam Reforming of Phenol over a Ni-Ca-Al-O Bifunctional Catalyst. ACS Sustain. Chem. Eng. 2020, 8, 7111–7120. [Google Scholar] [CrossRef]
  190. Nimmas, T.; Wongsakulphasatch, S.; Cheng, C.K.; Assabumrungrat, S. Bi-metallic CuO-NiO based multifunctional material for hydrogen production from sorption-enhanced chemical looping autothermal reforming of ethanol. Chem. Eng. J. 2020, 398, 125543. [Google Scholar] [CrossRef]
  191. Li, H.; Tian, H.; Chen, S.; Sun, Z.; Liu, T.; Liu, R.; Assabumrungrat, S.; Saupsor, J.; Mu, R.; Gong, C.P.J. Sorption enhanced steam reforming of methanol for high-purity hydrogen production over Cu-MgO/Al2O3 bifunctional catalysts. Appl. Catal. B Environ. 2020, 276, 119052. [Google Scholar] [CrossRef]
  192. Zhang, W.; Xie, H.; Yu, Z.; Wang, P.; Wang, Z.; Yu, Q. Steam reforming of tar from raw coke oven gas over bifunctional catalysts: Reforming performance for H2 production. Environ. Prog. Sustain. Energy 2021, 40, e13501. [Google Scholar] [CrossRef]
  193. Coenen, K.; Gallucci, F.; Cobden, P.; van Dijk, E.; Hensen, E.; Annaland, M.V. Influence of material composition on the CO2 and H2O adsorption capacities and kinetics of potassium-promoted sorbents. Chem. Eng. J. 2018, 334, 2115–2123. [Google Scholar] [CrossRef]
  194. Rocha, C.; Soria, M.A.; Madeira, L.M. Steam reforming of olive oil mill wastewater with in situ hydrogen and carbon dioxide separation—Thermodynamic analysis. Fuel 2017, 207, 449–460. [Google Scholar] [CrossRef]
  195. Lima da Silva, A.; Müller, I.L. Hydrogen production by sorption enhanced steam reforming of oxygenated hydrocarbons (ethanol, glycerol, n-butanol and methanol): Thermodynamic modelling. Int. J. Hydrogen Energy 2011, 36, 2057–2075. [Google Scholar] [CrossRef] [Green Version]
  196. Silva, J.M.; Soria, M.A.; Madeira, L.M. Thermodynamic analysis of Glycerol Steam Reforming for hydrogen production with in situ hydrogen and carbon dioxide separation. J. Power Sources 2015, 273, 423–430. [Google Scholar] [CrossRef] [Green Version]
  197. Rocha, C.; Soria, M.A.; Madeira, L.M. Olive mill wastewater valorization through steam reforming using hybrid multifunctional reactors for high-purity H2 production. Chem. Eng. J. 2022, 430, 132651. [Google Scholar] [CrossRef]
  198. Nations, U. COP26 Keeps 1.5C Alive and Finalises Paris Agreement. 2021. Available online: https://ukcop26.org/cop26-keeps-1-5c-alive-and-finalises-paris-agreement/ (accessed on 30 November 2021).
  199. Al-Mamoori, A.; Thakkar, H.; Li, X.; Rownaghi, A.A.; Rezaei, F. Development of Potassium- and Sodium-Promoted CaO Adsorbents for CO2 Capture at High Temperatures. Ind. Eng. Chem. Res. 2017, 56, 8292–8300. [Google Scholar] [CrossRef]
  200. Martavaltzi, C.S.; Lemonidou, A.A. Development of new CaO based sorbent materials for CO2 removal at high temperature. Microporous Mesoporous Mater. 2008, 110, 119–127. [Google Scholar] [CrossRef]
  201. Wang, S.; Shen, H.; Fan, S.; Zhao, Y.; Ma, X.; Gong, J. Enhanced CO2 adsorption capacity and stability using CaO-based adsorbents treated by hydration. AIChE J. 2013, 59, 3586–3593. [Google Scholar] [CrossRef]
  202. Westbye, A.; di Felice, L.; Aranda, A.; Dietzel, P.D.C. Efficient calcination of CaCO3 by means of combined Ca-Cu materials for Sorption Enhanced Reforming. In Proceedings of the 25th International Conference on Chemical Reaction Engineering, Florence, Italy, 20–23 May 2018. [Google Scholar]
  203. Zhang, Y.; Gong, X.; Chen, X.; Yin, L.; Zhang, J.; Liu, W. Performance of synthetic CaO-based sorbent pellets for CO2 capture and kinetic analysis. Fuel 2018, 232, 205–214. [Google Scholar] [CrossRef]
  204. Yoon, H.J.; Lee, K.B. Introduction of chemically bonded zirconium oxide in CaO-based high-temperature CO2 sorbents for enhanced cyclic sorption. Chem. Eng. J. 2019, 355, 850–857. [Google Scholar] [CrossRef]
  205. Guo, H.; Xu, Z.; Jiang, T.; Zhao, Y.; Ma, X.; Wang, S. The effect of incorporation Mg ions into the crystal lattice of CaO on the high temperature CO2 capture. J. CO2 Util. 2020, 37, 335–345. [Google Scholar] [CrossRef]
  206. Zhang, X.; Li, Z.; Peng, Y.; Su, W.; Sun, X.; Li, J. Investigation on a novel CaO-Y2O3 sorbent for efficient CO2 mitigation. Chem. Eng. J. 2014, 243 (Suppl. C), 297–304. [Google Scholar] [CrossRef]
  207. Lu, H.; Reddy, E.P.; Smirniotis, P.G. Calcium Oxide Based Sorbents for Capture of Carbon Dioxide at High Temperatures. Ind. Eng. Chem. Res. 2006, 45, 3944–3949. [Google Scholar] [CrossRef]
  208. Di Giuliano, A.; Gallucci, K. Sorption enhanced steam methane reforming based on nickel and calcium looping: A review. Chem. Eng. Processing Process Intensif. 2018, 130, 240–252. [Google Scholar] [CrossRef]
  209. Silva, J.M.; Trujillano, R.; Rives, V.; Soria, M.A.; Madeira, L.M. High temperature CO2 sorption over modified hydrotalcites. Chem. Eng. J. 2017, 325, 25–34. [Google Scholar] [CrossRef]
  210. Miguel, C.V.; Trujillano, R.; Rives, V.; Vicente, M.A.; Ferreira, A.F.P.; Rodrigues, A.E.; Mendes, A.; Madeira, L.M. High temperature CO2 sorption with gallium-substituted and promoted hydrotalcites. Sep. Purif. Technol. 2014, 127 (Suppl. C), 202–211. [Google Scholar] [CrossRef] [Green Version]
  211. Kim, S.; Jeon, S.G.; Lee, K.B. High-Temperature CO2 Sorption on Hydrotalcite Having a High Mg/Al Molar Ratio. ACS Appl. Mater. Interfaces 2016, 8, 5763–5767. [Google Scholar] [CrossRef] [PubMed]
  212. Yong, Z.; Mata, V.; Rodrigues, A.E. Adsorption of Carbon Dioxide onto Hydrotalcite-like Compounds (HTlcs) at High Temperatures. Ind. Eng. Chem. Res. 2001, 40, 204–209. [Google Scholar] [CrossRef]
  213. Wang, X.P.; Yu, J.J.; Cheng, J.; Hao, Z.P.; Xu, Z.P. High-Temperature Adsorption of Carbon Dioxide on Mixed Oxides Derived from Hydrotalcite-Like Compounds. Environ. Sci. Technol. 2008, 42, 614–618. [Google Scholar] [CrossRef] [PubMed]
  214. Ram Reddy, M.K.; ZXu, P.; Lu, G.Q.; da Costa, J.C.D. Layered Double Hydroxides for CO2 Capture: Structure Evolution and Regeneration. Ind. Eng. Chem. Res. 2006, 45, 7504–7509. [Google Scholar] [CrossRef]
  215. Maroño, M.; Torreiro, Y.; Gutierrez, L. Influence of steam partial pressures in the CO2 capture capacity of K-doped hydrotalcite-based sorbents for their application to SEWGS processes. Int. J. Greenh. Gas Control 2013, 14 (Suppl. C), 183–192. [Google Scholar] [CrossRef]
  216. Coenen, K.; Gallucci, F.; Cobden, P.; van Dijk, E.; Hensen, E.; Annaland, M.v. Chemisorption working capacity and kinetics of CO2 and H2O of hydrotalcite-based adsorbents for sorption-enhanced water-gas-shift applications. Chem. Eng. J. 2016, 293, 9–23. [Google Scholar] [CrossRef]
  217. Helwani, Z.; Wiheeb, A.D.; Kim, J.; Othman, M.R. Improved carbon dioxide capture using metal reinforced hydrotalcite under wet conditions. Int. J. Greenh. Gas Control 2012, 7, 127–136. [Google Scholar]
  218. Silva, J.M.; Trujillano, R.; Rives, V.; Soria, M.A.; Madeira, L.M. Dynamic behaviour of a K-doped Ga substituted and microwave aged hydrotalcite-derived mixed oxide during CO2 sorption experiments. J. Ind. Eng. Chem. 2019, 72, 491–503. [Google Scholar] [CrossRef]
  219. Boon, J.; Cobden, P.D.; van Dijk, H.A.J.; Hoogland, C.; van Selow, E.R.; Annaland, M.v. Isotherm model for high-temperature, high-pressure adsorption of CO2 and H2O on K-promoted hydrotalcite. Chem. Eng. J. 2014, 248, 406–414. [Google Scholar] [CrossRef]
  220. Halabi, M.H.; de Croon, M.H.J.M.; van der Schaaf, J.; Cobden, P.D.; Schouten, J.C. High capacity potassium-promoted hydrotalcite for CO2 capture in H2 production. Int. J. Hydrogen Energy 2012, 37, 4516–4525. [Google Scholar] [CrossRef]
  221. Wu, Y.-J.; Li, P.; Yu, J.-G.; Cunha, A.F.; Rodrigues, A.E. K-Promoted Hydrotalcites for CO2 Capture in Sorption Enhanced Reactions. Chem. Eng. Technol. 2013, 36, 567–574. [Google Scholar] [CrossRef]
  222. Wang, Q.; Wu, Z.; Tay, H.H.; Chen, L.; Liu, Y.; Chang, J.; Zhong, Z.; Luo, J.; Borgna, A. High temperature adsorption of CO2 on Mg-Al hydrotalcite: Effect of the charge compensating anions and the synthesis pH. Catal. Today 2011, 164, 198–203. [Google Scholar] [CrossRef]
  223. Hanif, A.; Dasgupta, S.; Divekar, S.; Arya, A.; Garg, M.O.; Nanoti, A. A study on high temperature CO2 capture by improved hydrotalcite sorbents. Chem. Eng. J. 2014, 236 (Suppl. C), 91–99. [Google Scholar] [CrossRef]
  224. Hutson, N.D.; Attwood, B.C. High temperature adsorption of CO2 on various hydrotalcite-like compounds. Adsorption 2008, 14, 781–789. [Google Scholar] [CrossRef]
  225. Sanna, A.; Thompson, S.; Whitty, K.J.; Maroto-Valer, M.M. Fly Ash Derived Lithium Silicate for in-situ Pre-combustion CO2 Capture. Energy Procedia 2017, 114 (Suppl. C), 2401–2404. [Google Scholar] [CrossRef]
  226. Seggiani, M.; Puccini, M.; Vitolo, S. Alkali promoted lithium orthosilicate for CO2 capture at high temperature and low concentration. Int. J. Greenh. Gas Control 2013, 17 (Suppl. C), 25–31. [Google Scholar] [CrossRef]
  227. Seggiani, M.; Stefanelli, E.; Puccini, M.; Vitolo, S. CO2 sorption/desorption performance study on K2CO3-doped Li4SiO4-based pellets. Chem. Eng. J. 2018, 339, 51–60. [Google Scholar] [CrossRef]
  228. Gao, N.; Ma, K.; Ding, T.; Cai, J.; Tian, Y.; Li, X. Enhanced carbon dioxide adsorption performance and kinetic study of K and Al co-doped Li4SiO4. Chin. Chem. Lett. 2018, 29, 482–484. [Google Scholar] [CrossRef]
  229. Hu, Y.; Liu, W.; Zhou, Z.; Yang, Y. Preparation of Li4SiO4 Sorbents for Carbon Dioxide Capture via a Spray-Drying Technique. Energy Fuels 2018, 32, 4521–4527. [Google Scholar] [CrossRef]
  230. Wang, H.; Zhang, J.; Wang, G.; Wang, Q.; Song, T. High-temperature capture of CO2 by Li4SiO4 prepared with blast furnace slag and kinetic analysis. J. Therm. Anal. Calorim. 2018, 133, 981–989. [Google Scholar] [CrossRef]
  231. Hu, Y.; Qu, M.; Li, H.; Yang, Y.; Yang, J.; Qu, W.; Liu, W. Porous extruded-spheronized Li4SiO4 pellets for cyclic CO2 capture. Fuel 2019, 236, 1043–1049. [Google Scholar] [CrossRef]
  232. Izquierdo, M.T.; Saleh, A.; Sánchez-Fernández, E.; Maroto-Valer, M.M.; García, S. High-Temperature CO2 Capture by Li4SiO4 Sorbents: Effect of CO2 Concentration and Cyclic Performance under Representative Conditions. Ind. Eng. Chem. Res. 2018, 57, 13802–13810. [Google Scholar] [CrossRef]
  233. Hu, Y.; Liu, W.; Yang, Y.; Qu, M.; Li, H. CO2 capture by Li4SiO4 sorbents and their applications: Current developments and new trends. Chem. Eng. J. 2019, 359, 604–625. [Google Scholar] [CrossRef]
  234. Kwon, Y.M.; Chae, H.J.; Cho, M.S.; Park, Y.K.; Seo, H.M.; Lee, S.C.; Kim, J.C. Effect of a Li2SiO3 phase in lithium silicate-based sorbents for CO2 capture at high temperatures. Sep. Purif. Technol. 2019, 214, 104–110. [Google Scholar] [CrossRef]
  235. Zhang, T.; Li, M.; Ning, P.; Jia, Q.; Wang, Q.; Wang, J. K2CO3 promoted novel Li4SiO4-based sorbents from sepiolite with high CO2 capture capacity under different CO2 partial pressures. Chem. Eng. J. 2020, 380, 122515. [Google Scholar] [CrossRef]
  236. Li, H.; Qu, M.; Hu, Y. Preparation of spherical Li4SiO4 pellets by novel agar method for high-temperature CO2 capture. Chem. Eng. J. 2020, 380, 122538. [Google Scholar] [CrossRef]
  237. Fang, Y.; Zou, R.; Chen, X. High-temperature CO2 adsorption over Li4SiO4 sorbents derived from different lithium sources. Can. J. Chem. Eng. 2020, 98, 1495–1500. [Google Scholar] [CrossRef]
  238. Kim, H.; Jang, H.D.; Choi, M. Facile synthesis of macroporous Li4SiO4 with remarkably enhanced CO2 adsorption kinetics. Chem. Eng. Process. 2015, 280, 132–137. [Google Scholar] [CrossRef]
  239. Palacios-Romero, L.M.; Pfeiffer, H. Lithium Cuprate (Li2CuO2): A New Possible Ceramic Material for CO2 Chemisorption. Chem. Lett. 2008, 37, 862–863. [Google Scholar] [CrossRef] [Green Version]
  240. Iwan, A.; Stephenson, H.; Ketchie, W.C.; Lapkin, A.A. High temperature sequestration of CO2 using lithium zirconates. Chem. Eng. J. 2009, 146, 249–258. [Google Scholar] [CrossRef]
  241. Peltzer, D.; Mùnera, J.; Cornaglia, L.; Strumendo, M. Characterization of potassium doped Li2ZrO3 based CO2 sorbents: Stability properties and CO2 desorption kinetics. Chem. Eng. J. 2018, 336, 1–11. [Google Scholar] [CrossRef]
  242. Halabi, M.H.; de Croon, M.H.J.M.; van der Schaaf, J.; Cobden, P.D.; Schouten, J.C. Reactor modeling of sorption-enhanced autothermal reforming of methane. Part I: Performance study of hydrotalcite and lithium zirconate-based processes. Chem. Eng. J. 2011, 168, 872–882. [Google Scholar] [CrossRef]
  243. Lee, K.B.; Verdooren, A.; Caram, H.S.; Sircar, S. Chemisorption of carbon dioxide on potassium-carbonate-promoted hydrotalcite. J. Colloid Interface Sci. 2007, 308, 30–39. [Google Scholar] [CrossRef] [PubMed]
  244. Iruretagoyena Ferrer, D. Supported Layered Double Hydroxides as CO2 Adsorbents for Sorption-Enhanced H2 Production. Ph.D. Thesis, Chemical Engineering Department, Imperial College London, London, UK, 2014. [Google Scholar]
  245. Akgsornpeak, A.; Witoon, T.; Mungcharoen, T.; Limtrakul, J. Development of synthetic CaO sorbents via CTAB-assisted sol-gel method for CO2 capture at high temperature. Chem. Eng. J. 2014, 237 (Suppl. C), 189–198. [Google Scholar] [CrossRef]
  246. Qin, Q.; Wang, J.; Zhou, T.; Zheng, Q.; Huang, L.; Zhang, Y.; Lu, P.; Umar, A.; Louis, B.; Wang, Q. Impact of organic interlayer anions on the CO2 adsorption performance of Mg-Al layered double hydroxides derived mixed oxides. J. Energy Chem. 2017, 26, 346–353. [Google Scholar] [CrossRef] [Green Version]
  247. Li, S.; Shi, Y.; Yang, Y.; Zheng, Y.; Cai, N. High-Performance CO2 Adsorbent from Interlayer Potassium-Promoted Stearate-Pillared Hydrotalcite Precursors. Energy Fuels 2013, 27, 5352–5358. [Google Scholar] [CrossRef]
  248. De Roy, A.; Forano, C.; Besse, J.P. Layered Double Hydroxides: Synthesis and post synthesis modification. In Layered Double Hydroxides: Present and Future; Nova Science Publishers: New York, NY, USA, 2001; pp. 1–39. [Google Scholar]
  249. Tronto, J.; Bordonal, A.C.; Naal, Z.; Valim, J.B. Conducting Polymers/Layered Double Hydroxides Intercalated Nanocomposites. In Materials Science—Advanced Topics; Mastai, Y., Ed.; IntechOpen: London, UK, 2013. [Google Scholar]
  250. Bhatta, L.K.G.; Subramanyam, S.; Chengala, M.D.; Olivera, S.; Venkatesh, K. Progress in hydrotalcite like compounds and metal-based oxides for CO2 capture: A review. J. Clean. Prod. 2015, 103 (Suppl. C), 171–196. [Google Scholar]
  251. Khan, A.I.; O’Hare, D. Intercalation chemistry of layered double hydroxides: Recent developments and applications. J. Mater. Chem. 2002, 12, 3191–3198. [Google Scholar] [CrossRef]
  252. Brindley, G.; Kikkawa, S. A crystal-chemical study of Mg, Al and Ni, N hydroxy-perchlorates and hydroxycarbonates. Am. Mineral. 1979, 64, 836–843. [Google Scholar]
  253. Meyer, O.; Roessner, F.; Rakoczy, R.A.; Fischer, R.W. Impact of Organic Interlayer Anions in Hydrotalcite Precursor on the Catalytic Activity of Hydrotalcite-Derived Mixed Oxides. ChemCatChem 2010, 2, 314–321. [Google Scholar] [CrossRef]
  254. Velu, S.; Ramjumar, V.; Narayana, A.; Swamy, C.S. Effect of interlayer anions on the physicochemical properties of zinc-aluminium hydrotalcite-like compounds. J. Mater. Sci. 1997, 32, 957–964. [Google Scholar] [CrossRef]
  255. Yamamoto, T.; Kodama, T.; Hasegawa, N.; Tsuji, M.; Tamaura, Y. Synthesis of hydrotalcite with high layer charge for CO2 adsorbent. Energy Convers. Manag. 1995, 36, 637–640. [Google Scholar] [CrossRef]
  256. Tanaka, R.; Ogino, I.; Mukai, S.R. Synthesis of Mg-Al Mixed Oxides with Markedly High Surface Areas from Layered Double Hydroxides with Organic Sulfonates. ACS Omega 2018, 3, 16916–16923. [Google Scholar] [CrossRef] [PubMed]
  257. Wang, Q.; Tay, H.H.; Zhong, Z.; Luo, J.; Borgna, A. Synthesis of high-temperature CO2 adsorbents from organo-layered double hydroxides with markedly improved CO2 capture capacity. Energy Environ. Sci. 2012, 5, 7526–7530. [Google Scholar] [CrossRef]
  258. Othman, M.R.; Helwani, Z.; Martunus; Fernando, W.J.N. Synthetic hydrotalcites from different routes and their application as catalysts and gas adsorbents: A review. Appl. Organomet. Chem. 2009, 23, 335–346. [Google Scholar] [CrossRef]
  259. Zeng, S.; Xu, X.; Wang, S.; Gong, Q.; Liu, R.; Yu, Y. Sand flower layered double hydroxides synthesized by co-precipitation for CO2 capture: Morphology evolution mechanism, agitation effect and stability. Mater. Chem. Phys. 2013, 140, 159–167. [Google Scholar] [CrossRef]
  260. Ezeh, C.I.; Huang, X.; Yang, X.; Sun, C.-G.; Wang, J. Sonochemical surface functionalization of exfoliated LDH: Effect on textural properties, CO2 adsorption, cyclic regeneration capacities and subsequent gas uptake for simultaneous methanol synthesis. Ultrason. Sonochem. 2017, 39, 330–343. [Google Scholar] [CrossRef] [Green Version]
  261. León, M.; Díaz, E.; Bennici, S.; Vega, A.; Ordóñez, S.; Auroux, A. Adsorption of CO2 on Hydrotalcite-Derived Mixed Oxides: Sorption Mechanisms and Consequences for Adsorption Irreversibility. Ind. Eng. Chem. Res. 2010, 49, 3663–3671. [Google Scholar] [CrossRef]
  262. Yang, Z.-z.; Wei, J.-j.; Zeng, G.-m.; Zhang, H.-q.; Tan, X.-f.; Ma, C.; Li, X.-c.; Li, Z.-h.; Zhang, C. A review on strategies to LDH-based materials to improve adsorption capacity and photoreduction efficiency for CO2. Coord. Chem. Rev. 2019, 386, 154–182. [Google Scholar]
  263. Wang, Q.; Tay, H.H.; Ng, D.J.W.; Chen, L.; Liu, Y.; Chang, J.; Zhong, Z.; Luo, J.; Borgna, A. The Effect of Trivalent Cations on the Performance of Mg-M-CO3 Layered Double Hydroxides for High-Temperature CO2 Capture. ChemSusChem 2010, 3, 965–973. [Google Scholar] [CrossRef]
  264. Goh, K.-H.; Lim, T.-T.; Dong, Z. Application of layered double hydroxides for removal of oxyanions: A review. Water Res. 2008, 42, 1343–1368. [Google Scholar] [PubMed]
  265. Hutson, N.D.; Speakman, S.A.; Payzant, E.A. Structural Effects on the High Temperature Adsorption of CO2 on a Synthetic Hydrotalcite. Chem. Mater. 2004, 16, 4135–4143. [Google Scholar] [CrossRef]
  266. Othman, M.R.; Fernando, W.J. Elevated temperature carbon dioxide capture via reinforced metal hydrotalcite. Microporous Mesoporous Mater. 2011, 138, 110–117. [Google Scholar]
  267. Tichit, D.; Rolland, A.; Prinetto, F.; Fetter, G.; Martinez-Ortiz, M.d.; Valenzuela, M.A.; Bosch, P. Comparison of the structural and acid-base properties of Ga- and Al-containing layered double hydroxides obtained by microwave irradiation and conventional ageing of synthesis gels. J. Mater. Chem. 2002, 12, 3832–3838. [Google Scholar] [CrossRef]
  268. Walspurger, S.; Cobden, P.D.; Safonova, O.V.; Wu, Y.; Anthony, E.J. High CO2 Storage Capacity in Alkali-Promoted Hydrotalcite-Based Material: In Situ Detection of Reversible Formation of Magnesium Carbonate. Chem. A Eur. J. 2010, 16, 12694–12700. [Google Scholar] [CrossRef]
  269. Othman, M.R.; Rasid, N.M.; Fernando, W.J.N. Mg-Al hydrotalcite coating on zeolites for improved carbon dioxide adsorption. Chem. Eng. Sci. 2006, 61, 1555–1560. [Google Scholar] [CrossRef]
  270. Yang, W.; Kim, Y.; Liu, P.K.T.; Sahimi, M.; Tsotsis, T.T. A study by in situ techniques of the thermal evolution of the structure of a Mg-Al-CO3 layered double hydroxide. Chem. Eng. Sci. 2002, 57, 2945–2953. [Google Scholar] [CrossRef]
  271. Reichle, W.T.; Kang, S.Y.; Everhardt, D.S. The nature of the thermal decomposition of a catalytically active anionic clay mineral. J. Catal. 1986, 101, 352–359. [Google Scholar] [CrossRef]
  272. Mascolo, G.; Mascolo, M.C. On the synthesis of layered double hydroxides (LDHs) by reconstruction method based on the “memory effect”. Microporous Mesoporous Mater. 2015, 214, 246–248. [Google Scholar] [CrossRef]
  273. Yuan, X.; Wang, Y.; Wang, J.; Zhou, C.; Tang, Q.; Rao, X. Calcined graphene/MgAl-layered double hydroxides for enhanced Cr(VI) removal. Chem. Eng. J. 2013, 221, 204–213. [Google Scholar] [CrossRef]
  274. Guo, Y.; Zhu, Z.; Qiu, Y.; Zhao, J. Enhanced adsorption of acid brown 14 dye on calcined Mg/Fe layered double hydroxide with memory effect. Chem. Eng. J. 2013, 219, 69–77. [Google Scholar] [CrossRef]
  275. Wang, Q.; Luo, J.; Zhong, Z.; Borgna, A. CO2 capture by solid adsorbents and their applications: Current status and new trends. Energy Environ. Sci. 2011, 4, 42–55. [Google Scholar]
  276. Yang, J.-I.; Kim, J.-N. Hydrotalcites for adsorption of CO2 at high temperature. Korean J. Chem. Eng. 2006, 23, 77–80. [Google Scholar] [CrossRef]
  277. Aschenbrenner, O.; McGuire, P.; Alsamaq, S.; Wang, J.; Supasitmongkol, S.; Al-Duri, B.; Styring, P.; Wood, J. Adsorption of carbon dioxide on hydrotalcite-like compounds of different compositions. Chem. Eng. Res. Des. 2011, 89, 1711–1721. [Google Scholar] [CrossRef] [Green Version]
  278. van Selow, E.R.; Cobden, P.D.; Wright, A.D.; van den Brink, R.W.; Jansen, D. Improved sorbent for the sorption-enhanced water-gas shift process. Energy Procedia 2011, 4, 1090–1095. [Google Scholar] [CrossRef] [Green Version]
  279. Coenen, K.; Gallucci, F.; Pio, G.; Cobden, P.; van Dijk, E.; Hensen, E.; Annaland, M.v. On the influence of steam on the CO2 chemisorption capacity of a hydrotalcite-based adsorbent for SEWGS applications. Chem. Eng. J. 2017, 314 (Suppl. C), 554–569. [Google Scholar] [CrossRef]
  280. Drits, V.A.; Bookin, A.S. Crystal Structure and Xray Identification of Layered Double Hydroxides. In Layered Double Hydroxides: Present and Future; Nova Science Publishers: New York, NY, USA, 2001; pp. 41–100. [Google Scholar]
  281. Oliveira, E.L.G.; Grande, C.A.; Rodrigues, A.E. CO2 sorption on hydrotalcite and alkali-modified (K and Cs) hydrotalcites at high temperatures. Sep. Purif. Technol. 2008, 62, 137–147. [Google Scholar] [CrossRef]
  282. Sun, L.; Yang, Y.; Ni, H.; Liu, D.; Sun, Z.; Li, P.; Yu, J. Enhancement of CO2 Adsorption Performance on Hydrotalcites Impregnated with Alkali Metal Nitrate Salts and Carbonate Salts. Ind. Eng. Chem. Res. 2020, 59, 6043–6052. [Google Scholar] [CrossRef]
  283. Zhu, X.; Chen, C.; Wang, Q.; Shi, Y.; O’Hare, D.; Cai, N. Roles for K2CO3 doping on elevated temperature CO2 adsorption of potassium promoted layered double oxides. Chem. Eng. J. 2019, 366, 181–191. [Google Scholar] [CrossRef]
  284. Zheng, Y.; Shi, Y.; Li, S.; Yang, Y.; Cai, N. Elevated temperature hydrogen/carbon dioxide separation process simulation by integrating elementary reaction model of hydrotalcite adsorbent. Int. J. Hydrogen Energy 2014, 39, 3771–3779. [Google Scholar] [CrossRef]
  285. Ebner, A.D.; Reynolds, S.P.; Ritter, J.A. Nonequilibrium Kinetic Model That Describes the Reversible Adsorption and Desorption Behavior of CO2 in a K-Promoted Hydrotalcite-like Compound. Ind. Eng. Chem. Res. 2007, 46, 1737–1744. [Google Scholar] [CrossRef]
  286. Ding, Y.; Alpay, E. Equilibria and kinetics of CO2 adsorption on hydrotalcite adsorbent. Chem. Eng. Sci. 2000, 55, 3461–3474. [Google Scholar] [CrossRef]
  287. Yavuz, C.T.; Shinall, B.D.; Iretskii, A.V.; White, M.G.; Golden, T.; Atilhan, M.; Ford, P.C.; Stucky, G.D. Markedly Improved CO2 Capture Efficiency and Stability of Gallium Substituted Hydrotalcites at Elevated Temperatures. Chem. Mater. 2009, 21, 3473–3475. [Google Scholar] [CrossRef]
  288. Yong, Z.; Rodrigues, A.r.E. Hydrotalcite-like compounds as adsorbents for carbon dioxide. Energy Convers. Manag. 2002, 43, 1865–1876. [Google Scholar] [CrossRef]
  289. Lee, C.H.; Choi, S.W.; Yoon, H.J.; Kwon, H.J.; Lee, H.C.; Jeon, S.G.; Lee, K.B. Na2CO3-doped CaO-based high-temperature CO2 sorbent and its sorption kinetics. Chem. Eng. J. 2018, 352, 103–109. [Google Scholar] [CrossRef]
  290. Coenen, K.; Gallucci, F.; Hensen, E.; Annaland, M.v. CO2 and H2O chemisorption mechanism on different potassium-promoted sorbents for SEWGS processes. J. CO2 Util. 2018, 25, 180–193. [Google Scholar] [CrossRef]
  291. Du, H.; Williams, C.T.; Ebner, A.D.; Ritter, J.A. In Situ FTIR Spectroscopic Analysis of Carbonate Transformations during Adsorption and Desorption of CO2 in K-Promoted HTlc. Chem. Mater. 2010, 22, 3519–3526. [Google Scholar] [CrossRef]
  292. Giorno, L. The Principle of Membrane Reactors. In Encyclopedia of Membranes; Drioli, E., Giorno, L., Eds.; Springer: Berlin/Heidelberg, Germany, 2015; pp. 1–3. [Google Scholar]
  293. Mendes, D.; Mendes, A.; Madeira, L.M.; Iulianelli, A.; Sousa, J.M.; Basile, A. The water-gas shift reaction: From conventional catalytic systems to Pd-based membrane reactors—A review. Asia-Pac. J. Chem. Eng. 2010, 5, 111–137. [Google Scholar]
  294. Arratibel, A.; Tanaka, A.P.; Laso, I.; Annaland, M.v.; Gallucci, F. Development of Pd-based double-skinned membranes for hydrogen production in fluidized bed membrane reactors. J. Membr. Sci. 2018, 550, 536–544. [Google Scholar] [CrossRef]
  295. Rahimpour, M.R.; Samimi, F.; Babapoor, A.; Tohidian, T.; Mohebi, S. Palladium membranes applications in reaction systems for hydrogen separation and purification: A review. Chem. Eng. Processing: Process Intensif. 2017, 121, 24–49. [Google Scholar]
  296. Fedotov, A.; Antonov, D.; Uvarov, V.; Tsodikov, M. Syngas and ultrapure hydrogen co-production from biomass products and synthetic fuel using hybrid membrane-catalytic reactor. In Proceedings of the 25th International Conference on Chemical Reaction Engineering, Florence, Italy, 20–23 May 2018. [Google Scholar]
  297. Lin, K.-H.; Lin, W.-H.; Hsiao, C.-H.; Chang, H.-F.; Chang, A.C.C. Hydrogen production in steam reforming of glycerol by conventional and membrane reactors. Int. J. Hydrogen Energy 2012, 37, 13770–13776. [Google Scholar] [CrossRef]
  298. Basile, A.; Gallucci, F.; Paturzo, L. Hydrogen production from methanol by oxidative steam reforming carried out in a membrane reactor. Catal. Today 2005, 104, 251–259. [Google Scholar] [CrossRef]
  299. Mendes, D.; Sá, S.; Tosti, S.; Sousa, J.M.; Madeira, L.M.; Mendes, A. Experimental and modeling studies on the low-temperature water-gas shift reaction in a dense Pd-Ag packed-bed membrane reactor. Chem. Eng. Sci. 2011, 66, 2356–2367. [Google Scholar] [CrossRef]
  300. Tong, J.; Su, L.; Kashima, Y.; Shirai, R.; Suda, H.; Matsumura, Y. Simultaneously Depositing Pd−Ag Thin Membrane on Asymmetric Porous Stainless Steel Tube and Application To Produce Hydrogen from Steam Reforming of Methane. Ind. Eng. Chem. Res. 2006, 45, 648–655. [Google Scholar] [CrossRef]
  301. Chang, A.C.C.; Lin, W.-H.; Lin, K.-H.; Hsiao, C.-H.; Chen, H.-H.; Chang, H.-F. Reforming of glycerol for producing hydrogen in a Pd/Ag membrane reactor. Int. J. Hydrogen Energy 2012, 37, 13110–13117. [Google Scholar] [CrossRef]
  302. Didenko, L.P.; Babak, V.N.; Sementsova, L.A.; Chizhov, P.E.; Dorofeeva, T.V. Effect of Pd-Ru alloy membrane thickness on H2 flux from steam reforming products. Pet. Chem. 2017, 57, 935–946. [Google Scholar] [CrossRef]
  303. Coenen, K.; Gallucci, F.; Mezari, B.; Hensen, E.; Annaland, M.v. An in-situ IR study on the adsorption of CO2 and H2O on hydrotalcites. J. CO2 Util. 2018, 24, 228–239. [Google Scholar] [CrossRef]
  304. Lytkina, A.A.; Orekhova, N.V.; Yaroslavtsev, A.B. Methanol Steam Reforming in Membrane Reactors. Pet. Chem. 2018, 58, 911–922. [Google Scholar] [CrossRef]
  305. Gallucci, F.; Paturzo, L.; Basile, A. Hydrogen Recovery from Methanol Steam Reforming in a Dense Membrane Reactor:  Simulation Study. Ind. Eng. Chem. Res. 2004, 43, 2420–2432. [Google Scholar] [CrossRef]
  306. Macedo, M.S.; Soria, M.A.; Madeira, L.M. Glycerol steam reforming for hydrogen production: Traditional versus membrane reactor. Int. J. Hydrogen Energy 2019, 44, 24719–24732. [Google Scholar] [CrossRef]
  307. Parente, M.; Soria, M.A.; Madeira, L.M. Hydrogen and/or syngas production through combined dry and steam reforming of biogas in a membrane reactor: A thermodynamic study. Renew. Energy 2020, 157, 1254–1264. [Google Scholar] [CrossRef]
  308. Gallucci, F.; Fernandez, E.; Corengia, P.; Annaland, M.v. Recent advances on membranes and membrane reactors for hydrogen production. Chem. Eng. Sci. 2013, 92 (Suppl. C), 40–66. [Google Scholar]
  309. Adhikari, S.; Fernando, S. Hydrogen Membrane Separation Techniques. Ind. Eng. Chem. Res. 2006, 45, 875–881. [Google Scholar] [CrossRef]
  310. Rebollo, E.; Mortalò, C.; Escolástico, S.; Boldrini, S.; Barison, S.; Serra, J.M.; Fabrizio, M. Exceptional hydrogen permeation of all-ceramic composite robust membranes based on BaCe0.65Zr0.20Y0.15O3−δ and Y- or Gd-doped ceria. Energy Environ. Sci. 2015, 8, 3675–3686. [Google Scholar] [CrossRef]
  311. Montaleone, D.; Mercadelli, E.; Escolástico, S.; Gondolini, A.; Serra, J.M.; Sanson, A. All-ceramic asymmetric membranes with superior hydrogen permeation. J. Mater. Chem. A 2018, 6, 15718–15727. [Google Scholar] [CrossRef]
  312. Miguel, C.V.; Mendes, A.; Tosti, S.; Madeira, L.M. Effect of CO and CO2 on H2 permeation through finger-like Pd-Ag membranes. Int. J. Hydrogen Energy 2012, 37, 12680–12687. [Google Scholar] [CrossRef]
  313. O’Brien, C.P.; Lee, I.C. CO Poisoning and CO Hydrogenation on the Surface of Pd Hydrogen Separation Membranes. J. Phys. Chem. C 2017, 121, 16864–16871. [Google Scholar] [CrossRef]
  314. Pérez, P.; Cornaglia, C.A.; Mendes, A.; Madeira, L.M.; Tosti, S. Surface effects and CO/CO2 influence in the H2 permeation through a Pd-Ag membrane: A comprehensive model. Int. J. Hydrogen Energy 2015, 40, 6566–6572. [Google Scholar] [CrossRef] [Green Version]
  315. Shirasaki, Y.; Tsuneki, T.; Seki, T.; Yasuda, I.; Sato, T.; Itoh, N. Improvement in Hydrogen Permeability of Palladium Membrane by Alloying with Transition Metals. J. Chem. Eng. Jpn. 2018, 51, 123–125. [Google Scholar] [CrossRef]
  316. Dalla Fontana, A.; Sirini, N.; Cornaglia, L.M.; Tarditi, A.M. Hydrogen permeation and surface properties of PdAu and PdAgAu membranes in the presence of CO, CO2 and H2S. J. Membr. Sci. 2018, 563, 351–359. [Google Scholar] [CrossRef]
  317. Lewis, F.A. The palladium-hydrogen system: Structures near phase transition and critical points. Int. J. Hydrogen Energy 1995, 20, 587–592. [Google Scholar] [CrossRef]
  318. Nakajima, T.; Kume, T.; Ikeda, Y.; Shiraki, M.; Kurokawa, H.; Iseki, T.; Kajitani, M.; Tanaka, H.; Hikosaka, H.; Ito, Y.T.M. Effect of concentration polarization on hydrogen production performance of ceramic-supported Pd membrane module. Int. J. Hydrogen Energy 2015, 40, 11451–11456. [Google Scholar] [CrossRef]
  319. Al-Mufachi, N.A.; Rees, N.V.; Steinberger-Wilkens, R. Hydrogen selective membranes: A review of palladium-based dense metal membranes. Renew. Sustain. Energy Rev. 2015, 47, 540–551. [Google Scholar] [CrossRef]
  320. Peters, T.A.; Kaleta, T.; Stange, M.; Bredesen, R. Development of ternary Pd-Ag-TM alloy membranes with improved sulphur tolerance. J. Membr. Sci. 2013, 429, 448–458. [Google Scholar] [CrossRef]
  321. Mejdell, A.L.; Jøndahl, M.; Peters, T.A.; Bredesen, R.; Venvik, H.J. Experimental investigation of a microchannel membrane configuration with a 1.4μm Pd/Ag23wt.% membrane—Effects of flow and pressure. J. Membr. Sci. 2009, 327, 6–10. [Google Scholar] [CrossRef]
  322. Peters, T.A.; Stange, M.; Bredesen, R. On the high pressure performance of thin supported Pd-23%Ag membranes—Evidence of ultrahigh hydrogen flux after air treatment. J. Membr. Sci. 2011, 378, 28–34. [Google Scholar] [CrossRef]
  323. Pizzi, D.; Worth, R.; Baschetti, M.G.; Sarti, G.C.; Noda, K.-i. Hydrogen permeability of 2.5 μm palladium-silver membranes deposited on ceramic supports. J. Membr. Sci. 2008, 325, 446–453. [Google Scholar] [CrossRef]
  324. Maneerung, T.; Hidajat, K.; Kawi, S. Ultra-thin (<1 μm) internally-coated Pd-Ag alloy hollow fiber membrane with superior thermal stability and durability for high temperature H2 separation. J. Membr. Sci. 2014, 452 (Suppl. C), 127–142. [Google Scholar] [CrossRef]
  325. Babak, V.N.; Didenko, L.P.; Kvurt, Y.P.; Sementsova, L.A. Studying the Operation of a Membrane Module Based on Palladium Foil at High Temperatures. Theor. Found. Chem. Eng. 2018, 52, 181–194. [Google Scholar] [CrossRef]
  326. Mendes, D. Use of Pd-Ag Membrane Reactors in the Water-Gas Shift Reaction for Producing Ultra-Pure Hydrogen. Ph.D. Thesis, FEUP, Porto, Portugal, 2010. [Google Scholar]
  327. Nooijer, N.d.; Plazaola, A.A.; Rey, J.M.; Fernandez, E.; Tanaka, D.A.P.; Annaland, M.v.S.; Gallucci, F. Long-Term Stability of Thin-Film Pd-Based Supported Membranes. Processes 2019, 7, 106. [Google Scholar] [CrossRef] [Green Version]
  328. Gallucci, F.; Medrano, J.; Fernandez, E.; Melendez, J.; Annaland, M.v.; Pacheco, A. Advances on High Temperature Pd-Based Membranes and Membrane Reactors for Hydrogen Purification and Production. J. Membr. Sci. Res. 2017, 3, 142–156. [Google Scholar]
  329. Okazaki, J.; Ikeda, T.; Tanaka, D.A.P.; Sato, K.; Suzuki, T.M.; Mizukami, F. An investigation of thermal stability of thin palladium-silver alloy membranes for high temperature hydrogen separation. J. Membr. Sci. 2011, 366, 212–219. [Google Scholar] [CrossRef]
  330. Gallucci, F.; Gesalaga, E.F.; Jimenez, J.A.M.; Tanaka, D.A.P.; Annaland, M.v. Pd-based membranes for high temperature applications: Current status. Austin Chem. Eng. 2016, 3, 1025. [Google Scholar]
  331. Tosti, S.; Basile, A.; Bettinali, L.; Borgognoni, F.; Chiaravalloti, F.; Gallucci, F. Long-term tests of Pd-Ag thin wall permeator tube. J. Membr. Sci. 2006, 284, 393–397. [Google Scholar] [CrossRef]
  332. Jokar, S.; Rahimpour, M.; Shariati, A.; Iulianelli, A.; Bagnato, G.; Vita, A.; Dalena, F.; Basile, A. Pure Hydrogen Production in Membrane Reactor with Mixed Reforming Reaction by Utilizing Waste Gas: A Case Study. Processes 2016, 4, 33. [Google Scholar] [CrossRef]
  333. Pacheco Tanaka, D.A.; Medrano, J.A.; Sole, J.L.V.; Gallucci, F. 1—Metallic membranes for hydrogen separation. In Current Trends and Future Developments on (Bio-) Membranes; Basile, A., Gallucci, F., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 1–29. [Google Scholar]
  334. Patil, C. Membrane Reactor Technology for Ultrapure Hydrogen Production. Ph.D. Thesis, University of Twente, Enschede, The Netherlands, 2005. [Google Scholar]
  335. Brunetti, A.; Caravella, A.; Fernandez, E.; Tanaka, D.A.P.; Gallucci, F.; Drioli, E.; Curcio, E.; Viviente, J.L.; Barbieri, G. Syngas upgrading in a membrane reactor with thin Pd-alloy supported membrane. Int. J. Hydrogen Energy 2015, 40, 10883–10893. [Google Scholar] [CrossRef]
  336. Fernandez, E.; Helmi, A.; Coenen, K.; Melendez, J.; Viviente, J.L.; Tanaka, D.A.P.; Annaland, M.v.; Gallucci, F. Development of thin Pd-Ag supported membranes for fluidized bed membrane reactors including WGS related gases. Int. J. Hydrogen Energy 2015, 40, 3506–3519. [Google Scholar] [CrossRef] [Green Version]
  337. Li, H.; Caravella, A.; Xu, H.Y. Recent progress in Pd-based composite membranes. J. Mater. Chem. A 2016, 4, 14069–14094. [Google Scholar]
  338. Shah, M.R.; Noble, R.D.; Clough, D.E. Analysis of transient permeation as a technique for determination of sorption and diffusion in supported membranes. J. Membr. Sci. 2006, 280, 452–460. [Google Scholar] [CrossRef]
  339. Campo, M.; Tanaka, A.; Mendes, A.; Sousa, J.M. 3—Characterization of membranes for energy and environmental applications. In Advanced Membrane Science and Technology for Sustainable Energy and Environmental Applications; Basile, A., Nunes, S.P., Eds.; Woodhead Publishing: Sawston, UK, 2011; pp. 56–89. [Google Scholar]
  340. Basile, A.; Gallucci, F.; Tosti, S. Synthesis Characterization, and Applications of Palladium Membranes. In Membrane Science and Technology; Elsevier: Amsterdam, The Netherlands, 2008; pp. 255–323. [Google Scholar]
  341. Yun, S.; Ted Oyama, S. Correlations in palladium membranes for hydrogen separation: A review. J. Membr. Sci. 2011, 375, 28–45. [Google Scholar]
  342. Ward, T.L.; Dao, T. Model of hydrogen permeation behavior in palladium membranes. J. Membr. Sci. 1999, 153, 211–231. [Google Scholar] [CrossRef]
  343. Musket, R.G. Effects of contamination on the interaction of hydrogen gas with palladium: A review. J. Less Common Met. 1976, 45, 173–183. [Google Scholar] [CrossRef]
  344. Zhang, X.-L. 7—Effect of poisoning on metallic membranes. In Current Trends and Future Developments on (Bio) Membranes; Basile, A., Gallucci, F., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 169–187. [Google Scholar]
  345. Hara, S.; Sakaki, K.; Itoh, N. Decline in Hydrogen Permeation Due to Concentration Polarization and CO Hindrance in a Palladium Membrane Reactor. Ind. Eng. Chem. Res. 1999, 38, 4913–4918. [Google Scholar] [CrossRef]
  346. Hou, K.; Hughes, R. The effect of external mass transfer, competitive adsorption and coking on hydrogen permeation through thin Pd/Ag membranes. J. Membr. Sci. 2002, 206, 119–130. [Google Scholar] [CrossRef]
  347. Zhang, J.; Liu, D.; He, M.; Xu, H.; Li, W. Experimental and simulation studies on concentration polarization in H2 enrichment by highly permeable and selective Pd membranes. J. Membr. Sci. 2006, 274, 83–91. [Google Scholar] [CrossRef]
  348. Mori, N.; Nakamura, T.; Noda, K.-i.; Sakai, O.; Takahashi, A.; Ogawa, N.; Sakai, H.; Iwamoto, Y.; Hattori, T. Reactor Configuration and Concentration Polarization in Methane Steam Reforming by a Membrane Reactor with a Highly Hydrogen-Permeable Membrane. Ind. Eng. Chem. Res. 2007, 46, 1952–1958. [Google Scholar] [CrossRef]
  349. Nagy, E.; Kulcsár, E. The effect of the concentration polarization and the membrane layer mass transport on membrane separation. Desalination Water Treat. 2010, 14, 220–226. [Google Scholar] [CrossRef]
  350. Caravella, A.; Hara, S.; Drioli, E.; Barbieri, G. Sieverts law pressure exponent for hydrogen permeation through Pd-based membranes: Coupled influence of non-ideal diffusion and multicomponent external mass transfer. Int. J. Hydrogen Energy 2013, 38, 16229–16244. [Google Scholar] [CrossRef]
  351. Caravella, A.; Scura, F.; Barbieri, G.; Drioli, E. Sieverts Law Empirical Exponent for Pd-Based Membranes: Critical Analysis in Pure H2 Permeation. J. Phys. Chem. B 2010, 114, 6033–6047. [Google Scholar] [CrossRef]
  352. Gielens, F.C.; Tong, H.D.; van Rijn, C.J.M.; Vorstman, M.A.G.; Keurentjes, J.T.F. Microsystem technology for high-flux hydrogen separation membranes. J. Membr. Sci. 2004, 243, 203–213. [Google Scholar] [CrossRef]
  353. Keurentjes, J.T.F.; Gielens, F.C.; Tong, H.D.; van Rijn, C.J.M.; Vorstman, M.A.G. High-Flux Palladium Membranes Based on Microsystem Technology. Ind. Eng. Chem. Res. 2004, 43, 4768–4772. [Google Scholar] [CrossRef]
  354. Mendes, D.; Chibante, V.; Zheng, J.-M.; Tosti, S.; Borgognoni, F.; Mendes, A.; Madeira, L.M. Enhancing the production of hydrogen via water-gas shift reaction using Pd-based membrane reactors. Int. J. Hydrogen Energy 2010, 35, 12596–12608. [Google Scholar] [CrossRef]
  355. Santucci, A.; Borgognoni, F.; Vadrucci, M.; Tosti, S. Testing of dense Pd-Ag tubes: Effect of pressure and membrane thickness on the hydrogen permeability. J. Membr. Sci. 2013, 444 (Suppl. C), 378–383. [Google Scholar] [CrossRef]
  356. Lewis, A.E.; Kershner, D.C.; Paglieri, S.N.; Slepicka, M.J.; Way, J.D. Pd-Pt/YSZ composite membranes for hydrogen separation from synthetic water-gas shift streams. J. Membr. Sci. 2013, 437 (Suppl. C), 257–264. [Google Scholar] [CrossRef]
  357. Zhang, D.; Zhou, S.; Fan, Y.; Xu, N.; He, Y. Preparation of dense Pd composite membranes on porous Ti-Al alloy supports by electroless plating. J. Membr. Sci. 2012, 387–388 (Suppl. C), 24–29. [Google Scholar] [CrossRef]
  358. Pacheco Tanaka, D.A.; Tanco, M.A.L.; Niwa, S.-i.; Wakui, Y.; Mizukami, F.; Namba, T.; Suzuki, T.M. Preparation of palladium and silver alloy membrane on a porous α-alumina tube via simultaneous electroless plating. J. Membr. Sci. 2005, 247, 21–27. [Google Scholar] [CrossRef]
  359. Anzelmo, B.; Wilcox, J.; Liguori, S. Hydrogen production via natural gas steam reforming in a Pd-Au membrane reactor. Comparison between methane and natural gas steam reforming reactions. J. Membr. Sci. 2018, 568, 113–120. [Google Scholar] [CrossRef]
  360. Liu, J.; Bellini, S.; de Nooijer, N.C.A.; Sun, Y.; Tanaka, D.A.P.; Tang, C.; Li, H.; Gallucci, F.; Caravella, A. Hydrogen permeation and stability in ultra-thin PdRu supported membranes. Int. J. Hydrogen Energy 2020, 45, 7455–7467. [Google Scholar] [CrossRef]
  361. Catalano, J.; Baschetti, M.G.; Sarti, G.C. Influence of the gas phase resistance on hydrogen flux through thin palladium-silver membranes. J. Membr. Sci. 2009, 339, 57–67. [Google Scholar] [CrossRef]
  362. Melendez, J.; Gesalaga, E.F.; Farimani, A.H.S.; Gallucci, F.; Arias, P.L.; Tanaka, D.A.P. Preparation and characterization of ultra-thin (<1 micron) Pd-Ag membranes on porous alumina support (100 nm pore size). In Proceedings of the 12th International Conference on Catalysis in Membrane Reactors (ICCMR12), Szczecin, Poland, 22–25 June 2015. [Google Scholar]
  363. Helmi, A.; Fernandez, E.; Melendez, J.; Tanaka, D.A.P.; Gallucci, F.; Annaland, M.v. Fluidized Bed Membrane Reactors for Ultra Pure H2 Production—A Step forward towards Commercialization. Molecules 2016, 21, 376. [Google Scholar] [CrossRef] [Green Version]
  364. Fernandez, E.; Sanchez-Garcia, J.A.; Melendez, J.; Spallina, V.; Annaland, M.v.; Gallucci, F.; Tanaka, D.A.P.; Prema, R. Development of highly permeable ultra-thin Pd-based supported membranes. Chem. Eng. J. 2016, 305, 149–155. [Google Scholar] [CrossRef]
  365. Fernandez, E.; Helmi, A.; Medrano, J.A.; Coenen, K.; Arratibel, A.; Melendez, J.; de Nooijer, N.C.A.; Spallina, V.; Viviente, J.L.; Zuñiga, J.; et al. Palladium based membranes and membrane reactors for hydrogen production and purification: An overview of research activities at Tecnalia and TU/e. Int. J. Hydrogen Energy 2017, 42, 13763–13776. [Google Scholar] [CrossRef] [Green Version]
  366. Shi, L.; Goldbach, A.; Xu, H. High-flux H2 separation membranes from (Pd/Au)n nanolayers. Int. J. Hydrogen Energy 2011, 36, 2281–2284. [Google Scholar] [CrossRef]
Figure 1. World and Europe Union production of olive oil for the 2019/2020 crop year—data taken from the site of International Olive Council [2]. Figure adapted from the site of International Olive Council [2].
Figure 1. World and Europe Union production of olive oil for the 2019/2020 crop year—data taken from the site of International Olive Council [2]. Figure adapted from the site of International Olive Council [2].
Energies 15 00920 g001
Figure 2. World consumption/production of olive oil—data taken from the site of International Olive Council [2].
Figure 2. World consumption/production of olive oil—data taken from the site of International Olive Council [2].
Energies 15 00920 g002
Figure 3. Main processes of olive oil production—traditional (pressing) method, (a) three-phase centrifugation method and (b) two-phase centrifugation method [7,8]. Reprinted with permission from Ref. [7]. Copyright 2021 Elsevier.
Figure 3. Main processes of olive oil production—traditional (pressing) method, (a) three-phase centrifugation method and (b) two-phase centrifugation method [7,8]. Reprinted with permission from Ref. [7]. Copyright 2021 Elsevier.
Energies 15 00920 g003
Figure 4. A quantitative relationship between the OMW production and the remaining components present in the olive oil extraction. * The olive husk mass presented in the two-phase centrifugation process does not include the mass lost in secondary drying processes. Data taken from Caputo et al. [14].
Figure 4. A quantitative relationship between the OMW production and the remaining components present in the olive oil extraction. * The olive husk mass presented in the two-phase centrifugation process does not include the mass lost in secondary drying processes. Data taken from Caputo et al. [14].
Energies 15 00920 g004
Figure 6. Global H2 demand by sector in the Net Zero Emission Scenario, 2020–2030—data taken from the site of International Energy Agency [86,87].
Figure 6. Global H2 demand by sector in the Net Zero Emission Scenario, 2020–2030—data taken from the site of International Energy Agency [86,87].
Energies 15 00920 g006
Figure 7. Scheme of the operation of the TR.
Figure 7. Scheme of the operation of the TR.
Energies 15 00920 g007
Figure 8. H2 production at 700 °C over time on stream, FOMW = 0.6 mL·min−1, GHSV = 9000 h−1. Reprinted with permission from Ref. [28]. Copyright 2021 Elsevier.
Figure 8. H2 production at 700 °C over time on stream, FOMW = 0.6 mL·min−1, GHSV = 9000 h−1. Reprinted with permission from Ref. [28]. Copyright 2021 Elsevier.
Energies 15 00920 g008
Figure 9. (a) Gaseous species yield and TOC conversion during a stability test using a Rh-based catalyst. (b) H2 purity and H2 selectivity obtained using a Rh catalyst. For interpretation of the references to color in this figure legend, the reader is referred to the Web version of this article. Reprinted with permission from Ref. [85]. Copyright 2021 Elsevier.
Figure 9. (a) Gaseous species yield and TOC conversion during a stability test using a Rh-based catalyst. (b) H2 purity and H2 selectivity obtained using a Rh catalyst. For interpretation of the references to color in this figure legend, the reader is referred to the Web version of this article. Reprinted with permission from Ref. [85]. Copyright 2021 Elsevier.
Energies 15 00920 g009
Figure 10. Scheme of the operation of the SER.
Figure 10. Scheme of the operation of the SER.
Energies 15 00920 g010
Figure 11. Example of product distribution as a function of time in the outlet of an SER (steam reforming of ethanol for hydrogen production). Reprinted with permission from Ref. [162]. Copyright 2021 Elsevier.
Figure 11. Example of product distribution as a function of time in the outlet of an SER (steam reforming of ethanol for hydrogen production). Reprinted with permission from Ref. [162]. Copyright 2021 Elsevier.
Energies 15 00920 g011
Figure 12. Schematic representation of the hydrotalcite structure. Adapted from Tronto et al. [249].
Figure 12. Schematic representation of the hydrotalcite structure. Adapted from Tronto et al. [249].
Energies 15 00920 g012
Figure 13. The structural evolution of a typical HTC as a function of calcination temperature. Adapted with permission from Ref. [275]. Copyright 2021 Royal Society of Chemistry.
Figure 13. The structural evolution of a typical HTC as a function of calcination temperature. Adapted with permission from Ref. [275]. Copyright 2021 Royal Society of Chemistry.
Energies 15 00920 g013
Figure 14. CO2 sorption equilibrium isotherms at 300 °C on the calcined HTCs. The lines correspond to the fitting using the Freundlich equation (cHTC—calcined HTC; cK-HTC—calcined K-promoted HTC; cK-HTC MW—calcined K-promoted HTC aged under microwave irradiation; cK-HTCGa MW—calcined K-promoted Ga-substituted HTC aged under microwave irradiation). Reprinted with permission from Ref. [209]. Copyright 2021 Elsevier.
Figure 14. CO2 sorption equilibrium isotherms at 300 °C on the calcined HTCs. The lines correspond to the fitting using the Freundlich equation (cHTC—calcined HTC; cK-HTC—calcined K-promoted HTC; cK-HTC MW—calcined K-promoted HTC aged under microwave irradiation; cK-HTCGa MW—calcined K-promoted Ga-substituted HTC aged under microwave irradiation). Reprinted with permission from Ref. [209]. Copyright 2021 Elsevier.
Energies 15 00920 g014
Figure 15. Gaseous species yield and TOC conversion using an SER at (a) 350 °C and 1 bar or (b) 400 °C and 4 bar. * Thermodynamic equilibrium for the H2 yield in a TR at the operations conditions used in the respective experiments (dashed horizontal line). Dashed red vertical lines indicate instants of reactive regeneration. Reprinted with permission from Ref. [197]. Copyright 2021 Elsevier.
Figure 15. Gaseous species yield and TOC conversion using an SER at (a) 350 °C and 1 bar or (b) 400 °C and 4 bar. * Thermodynamic equilibrium for the H2 yield in a TR at the operations conditions used in the respective experiments (dashed horizontal line). Dashed red vertical lines indicate instants of reactive regeneration. Reprinted with permission from Ref. [197]. Copyright 2021 Elsevier.
Energies 15 00920 g015
Figure 16. Scheme of the operation of the MR.
Figure 16. Scheme of the operation of the MR.
Energies 15 00920 g016
Figure 17. Solution-diffusion mechanism of the permeation of H2 through a Pd-based membrane [337].
Figure 17. Solution-diffusion mechanism of the permeation of H2 through a Pd-based membrane [337].
Energies 15 00920 g017
Figure 18. Effect of pressure in OMW–ethanol co-reforming experimental tests (@ 450 °C, ethanol concentration 7.5 vol.%, feed: 15 g·h−1 mixture and 15 mL·min−1 carrier gas): (a) permeate, (b) retentate, and (c) total produced H2. Diamonds (◆) represent experiments with N2, while squares (■) are obtained with air as the carrier gas. Reprinted from Alique et al. [90].
Figure 18. Effect of pressure in OMW–ethanol co-reforming experimental tests (@ 450 °C, ethanol concentration 7.5 vol.%, feed: 15 g·h−1 mixture and 15 mL·min−1 carrier gas): (a) permeate, (b) retentate, and (c) total produced H2. Diamonds (◆) represent experiments with N2, while squares (■) are obtained with air as the carrier gas. Reprinted from Alique et al. [90].
Energies 15 00920 g018
Figure 19. Scheme of the operation of the SEMR.
Figure 19. Scheme of the operation of the SEMR.
Energies 15 00920 g019
Figure 20. Gaseous species yield and TOC conversion using a SEMR operation at (a) 350 °C and 1 bar and at (b) 400 °C and 4 bar. Sweep gas = 100 mLN2·min−1 and Ppermeate side = 1 bar. * Thermodynamic equilibrium for the H2 yield in a TR at the operations conditions used in the respective catalytic test (dashed horizontal line). Dashed red vertical lines indicate instants of reactive regeneration. Reprinted with permission from Ref. [197]. Copyright 2021 Elsevier.
Figure 20. Gaseous species yield and TOC conversion using a SEMR operation at (a) 350 °C and 1 bar and at (b) 400 °C and 4 bar. Sweep gas = 100 mLN2·min−1 and Ppermeate side = 1 bar. * Thermodynamic equilibrium for the H2 yield in a TR at the operations conditions used in the respective catalytic test (dashed horizontal line). Dashed red vertical lines indicate instants of reactive regeneration. Reprinted with permission from Ref. [197]. Copyright 2021 Elsevier.
Energies 15 00920 g020
Table 1. Typical physicochemical characterization of the OMW. Reprinted with permission from Ref. [23]. Copyright 2022 Elsevier.
Table 1. Typical physicochemical characterization of the OMW. Reprinted with permission from Ref. [23]. Copyright 2022 Elsevier.
Organic SpeciesInorganic Species
Solid Waste (g·L−1)11.5–102.5Pb (μg·L−1)6.7–10Ca (g·L−1)0.03–0.29
Organic Matter (g·L−1)16.7–81.6Cd (μg·L−1)0.03–10K (g·L−1)0.73–6.1
Lipids (g·L−1)1.64–9.8Fe (mg·L−1)0.45–20Cl (g·L−1)0.76–1
Polyphenols (g·L−1)0.002–11.5Zn (mg·L−1)1.7–4.98Na (g·L−1)0.03–0.13
Sugars (g·L−1)1.3–8.79Cu (mg·L−1)0.49–2.96
Organic Acids (g·L−1)0.78–1Mn (mg·L−1)0.46–20
Total Nitrogen (g·L−1)0.06–0.95Mg (mg·L−1)0.03–0.17
Table 2. Typical composition (wt.%) on a dry basis of the OMW. Reprinted with permission from Ref. [36]. Copyright 2022 John Wiley and Sons.
Table 2. Typical composition (wt.%) on a dry basis of the OMW. Reprinted with permission from Ref. [36]. Copyright 2022 John Wiley and Sons.
Ratio of Polyphenols/Sugars0.46
Species
Tyrosol16.68%
P-Coumaric acid14.83%
D-Arabinose22.83%
D-Galactose22.83%
D-Galacturonic acid22.83%
Table 3. Summary of the catalytic performance and operating conditions of several catalysts on OMWSR.
Table 3. Summary of the catalytic performance and operating conditions of several catalysts on OMWSR.
CatalystTemperature (°C)Pressure (Bar)Feed Flow RateFOMW (g·h−1)Conversion (%)H2 Yield (%)H2 Purity (%)Ref.
1 wt.% Pt/Al2O3 (commercial)4503WHSV (a) = 5.56 × 10−3 mol·h−1·gcat−110 n.d. (b)40 (c)38 (d)[6]
1 wt.% Pt/Al2O3 (commercial)4505WHSV (a) = 1.74 × 10−2 mol·h−1·gcat−115≈60 (e)29 (c)25 (d)[89]
<3 wt.% Pt-Rh-Pd/CeZr2-Al2O34505WHSV (a) = 1.74 × 10−2 mol·h−1·gcat−115>90 (e)35 (c)35 (d)[89]
1 wt.% Pt/CeLa7001GHSV (f) = 13,500 h−15490–96 (g)58 (i)72 (h)[28]
1 wt.% Ru/CeLa7001GHSV (f)= 13,500 h−15490–96 (g)52 (i)72 (h)[28]
1 wt.% Rh/CeLa7001GHSV (f) = 13,500 h−15490–96 (g)44 (i)73 (h)[28]
10 wt.% Ni/CeLa7001GHSV (f) = 13,500 h−15490–96 (g)42 (i)65 (h)[28]
2 wt.% Rh/M-Al2O34001WHSV (a) = 1.75 mol·h−1·gcat−130 (j)≈90 (g)10 (k)≈70 (h)[85]
(a) WHSV: weight hourly space velocity. (b) Not defined. (c) Yield (%) = H 2 perm TOC feed × 100 , in an MR with a H2 recovery of 80%. Recovery (%) = H 2 perm H 2 perm   + H 2 ret   × 100 . (d) H2 purity (%) = H 2 H 2 + CH 4 + CO 2 + CO × 100 , in an MR, on the retentate side, with a H2 recovery of 80%. Recovery (%) = H 2 perm H 2 perm   + H 2 ret   × 100 . (e) Conversion of phenols. (f) GHSV: Gas Hourly Space Velocity. (g) Conversion of organic carbon. (h) H2 purity (%) = H 2 H 2 + CH 4 + CO 2 + CO × 100 ,   in a TR. (i) Yield (%) = H 2 prod TOC feed × 100 , in a TR. (j) Volumetric flow rate of OMW in mL·h−1. (k) Yield (%) = H 2 prod F OS , where F OS is the molar flow rate of organic species present in the OMW fed to the reactor.
Table 4. Summary of the catalytic performance and operating conditions of catalysts on steam reforming reactions of oxygenated compounds.
Table 4. Summary of the catalytic performance and operating conditions of catalysts on steam reforming reactions of oxygenated compounds.
CatalystTemperature (°C)Pressure (Bar)Feed Flow RateSCFR (a)Conversion (%)H2 Yield (b) (%)H2 Selectivity (c)/Purity (d) (%)Ref.
3.5 wt.% Ni/5 wt.% La2O3-ZrO2 17001GHSV (e) = 240,000 h−151008789/64[105]
(2.5 + 2.5) wt.% Ni-Cu/Al2O3 17501WHSV (f) = 28 h−11.251006768/57[156]
15 wt.% Ni/α-Al2O3 16001WHSV (f) = 20 h−11≈1009090/66[107]
(15 + 2) wt.% Ni-Ru/10 wt.% CeO2-Al2O3 17001WHSV (f) = 21 h−13.2916779/61[140]
10 wt.% Ni/La2O3-Al2O3 1
(La2O3/Al2O3 = 1:3, weight ratio)
7001LHSV (g) = 10 h−111007271/59[134]
15 wt.% Ni/Al2O3 17001FAcetic Acid = 10 g·h−111005758/54[157]
3.2 wt.% Ni/CeO2-ZrO2 17001WHSV (f) = 25 h−114≈846888/64[106]
0.5 wt.% Pt/CeO2 1700n.s. (h)GHSV (e) = 80,000 h−159894≈100/67[113]
28 wt.% Co/7 wt.% K-Al2O3 16001LHSV (g) = 10 h−17.51008687/64[138]
(6.6 + 10) wt.% Ni-Fe/(CeO2)0.4-PG0.6 16001LHSV (g) = 14,427 h−13≈978884/63[141]
(35 wt.% NiO) Ni/Hydrotalcite 1 6001n.s. (h)1.5≈1004767/56[111]
Ni/BC4 17001LHSV (g) = 10 h−12.591.27179/61[129]
10 wt.% Ni/ATTP 16001LHSV (g) = 7.2 h−15≈1007291/65[128]
25 wt.% Ni/Atta 17001LHSV (g) = 10.4 h−15968990/64[93]
15 wt.% Ni/Al2O3-La2O3-CeO2 16501QT = 2.5 mL·h−16.5≈1008890/64[151]
Co (unsupported)6001LHSV (g) = 4 h−19.2100≈10099/66[121]
3 wt.% Ni/(0.4 + 5) wt.% K-LaO2-ZrO2 27001GHSV (e) = 950,000 h−120≈818398/70[136]
0.1 wt.% Rh/(50 + 25) wt.% MgO-CeO2-ZrO2 27001GHSV (e) = 80,000 h−1 13979086/67[116]
Ni-Fe/olivine 28501W/F (i) = 10.28 kgcat·h·m−31≈100n.d. (h)81/63[158]
1.5 wt.% Pt/CeO2-Al2O3 37001QT = 300 mL·min−16≈95n.d. (h)81/66[114]
16.4 wt.% Ni/Al2O3 38001GHSV (e) = 61,200 h−13>949175/66[159]
(a) Steam to carbon feed ratio. (b) Yield (%) = H 2 H 2 M á x . × 100 . (c) Selectivity (%) = H 2 RR × CO 2 + CH 4 + CO × 100 . RR is H2/CO2 reforming ratio. (d) Purity (%) = H 2 H 2 + CO 2 + CH 4 + CO × 100 . (e) GHSV: gas hourly space velocity. (f) WHSV: weight hourly space velocity. (g) LHSV: liquid hourly space velocity. (h) Not determined/specified. (i) Wcat/FPhenol. 1 Steam reforming of acetic acid. 2 Steam reforming of phenol. 3 Steam reforming of toluene.
Table 5. Stability and preparation methods for Ni-based catalysts used on steam reforming reactions of oxygenated species.
Table 5. Stability and preparation methods for Ni-based catalysts used on steam reforming reactions of oxygenated species.
CatalystMass of Catalyst (g)StabilityPreparation MethodRef.
3.5 wt.% Ni/5 wt.% La2O3-ZrO2 10.050Lost 7% H2 yield in 20 hImpregnation[105]
(2.5 + 2.5) wt.% Ni-Cu/Al2O3 10.1Result of 7.5 h
(no deactivation)
Impregnation over support prepared by following evaporation induced self-assembly[156]
15 wt.% Ni/α-Al2O3 11.5n.d. (a)Impregnation[107]
10 wt.% Ni/La2O3-Al2O3 1
(La2O3/Al2O3 = 1:3, weight ratio)
0.2Result of 30 h
(no deactivation)
Co-precipitation[134]
15 wt.% Ni/Al2O3 10.2n.d. (a)Incipient wetness impregnation[157]
6.6 wt.% Ni-10 wt.% Fe/(CeO2)0.4-PG0.6 1n.s. (a)Result of 20 h
(no deactivation)
Co-precipitation[141]
Ni/BC4 10.15n.d. (a)Impregnation[129]
10 wt.% Ni/ATTP 10.5n.d. (a)Impregnation[128]
16.4 wt.% Ni/Al2O3 2n.s. (a)Result of 5 h
(no deactivation)
Impregnation[159]
(a) Not determined/specified. 1 Steam reforming of acetic acid. 2 Steam reforming of toluene.
Table 6. Sorption capacities and sorption/regeneration temperatures of several CO2 sorbents.
Table 6. Sorption capacities and sorption/regeneration temperatures of several CO2 sorbents.
SorbentSorption Capacity (a) (mmol·g−1)Temperature (°C)Regeneration Temperature (°C)pCO2 (Bar)pH2O (Bar)Ref.
CaO-Y2O3 (20 wt.% Y2O3)12.95850850n.s. (b)-[206]
CaO17.306007000.30-[207]
CuO/CaO/Ca12Al14O33 (19/43/38 wt.%)3.146508700.150.25[202]
Na2CO3-CaO14.78008001.00-[208]
CaO-Mg10.226007000.50-[205]
Li2ZrO35.00500n.s. (b)n.s. (b)n.s. (b)[242]
Li4SiO47.296007501.00-[236]
Li4SiO46.245507000.15-[229]
K-Li4SiO4 (30 wt.% K2CO3)5.235807000.04-[226]
K-Li4SiO45.005757001.00-[241]
Li4SiO45.005507000.15-[229]
CaNd7510.9650n.s. (b)0.15-[203]
cHTC (Mg/Al = 2)0.81300n.s. (b)1.03-[209]
HTC (Mg/Al = 2)0.54300n.s. (b)0.14-[210]
K-HTC0.79400n.s. (b)1.00-[243]
K-HTC (20 wt.% K2CO2, Mg/Al = 0.6)0.284004000.50-[219]
K-HTC (22 wt.% K2CO2, Mg/Al ≈ 2)0.584004000.47-[220]
HTC-20K (20 wt.% K, Mg/Al = 2)2.14300n.s. (b)1.10-[210]
cK-HTC (20 wt.% K, Mg/Al = 21.42300n.s. (b)1.02-[209]
MG30-KN (20 wt.% K, Mg/Al ≈ 0.5)1.08 3354350.500.5[221]
K-HTC (20 wt.% K2CO2 and Mg/Al = 2)9.403006000.344.5[215]
K-HTC MW (20 wt.% K2CO2, Mg/Al = 2)1.46300n.s. (b)1.00-[223]
cK-HTC MW (20 wt.% K, Mg/Al = 2)1.51300n.s. (b)1.00-[209]
HTC-10Ga(10 wt.% Ga, Mg/Al = 2)0.58300n.s. (b)1.12-[210]
HTC-10Ga-20K (10 wt.% Ga, 20 wt.% K, Mg/Al = 2)1.823003001.08-[210]
cK-HTCGa MW (20 wt.% K, Mg/(Al+Ga) = 2, Al/Ga = 9)1.70300n.s. (b)1.05-[209]
Li/Na/K-MgAl-C16 (55 wt.% Li-Na-K, Mg/Al = 20)3.212004001.00-[246]
Mg3Al1-CO30.53200n.s. (b)1.00-[222]
Mg3Al1-NO30.21200n.s. (b)1.00-[222]
Mg3Al1-HNO30.18200n.s. (b)1.00-[222]
Mg3Al1-SO40.10200n.s. (b)1.00-[222]
Mg3Al1-Cl0.18200n.s. (b)1.00-[222]
(Mg,Al)(Cl) (Mg/Al = 3)0.44330n.s. (b)1.00-[224]
(Mg,Al)(ClO4) (Mg/Al = 3)3.55330n.s. (b)1.00-[224]
(Mg,Al)(Fe(CN)6) (Mg/Al = 3)0.75330n.s. (b)1.00-[224]
MgAl-C16 (Mg/Al = 3)0.912004001.00-[246]
Mg3Al-stearate1.25300n.s. (b)1.00-[247]
(a) Sorption capacity of the fresh sorbent. (b) Not specified (or regeneration step not performed).
Table 7. Summary of Pd-based membranes’ performances and respective characteristics (H2 diffusion through the membrane controls the permeation process).
Table 7. Summary of Pd-based membranes’ performances and respective characteristics (H2 diffusion through the membrane controls the permeation process).
Membrane/Support Temperature (°C)δ (µm)H2 Permeance (mol·m−2·s−1·Pa−0.5) (a)Selectivity (H2/N2)Ref.
Pd-25 wt.% Ag/Al2O3300501.15 × 10−4 [354]
Pd-23-25 wt.% Ag350842.26 × 10−4 [355]
Pd-5 wt.% Pt/YSZ4006.61.18 × 10−3 994[356]
Pd/TiO2/Ti-Al500141.07 × 10−3[357]
Pd-23 wt.% Ag/stainless steel4001501.27 × 10−4[6]
Pd-15 wt.% Ag/Al2O330052.13 × 10−3528[358]
Pd-23 wt.% Ag400611.46 × 10−4 [331]
Pd-20 wt.% Ag/Al2O34002.52.70 × 10−3 [323]
Pd-8 wt.% Ho6001007.89 × 10−4 n.d. (b)[315]
Pd-Au/PSS450121.30 × 10−3[359]
Pd-1.3 wt.% Ru/Al2O35001.84.12 × 10−32611[360]
Pd-20 wt.% Ag/Al2O34002.53.90 × 10−3≥10,000[361]
Pd-23 wt.% Ag/PSS4002.83.25 × 10−3≥2900[322]
(a) H2 Permeance obtained by Equation (15). (b) Not determined.
Table 8. Summary of Pd-based thin membranes’ performances and respective characteristics (H2 diffusion through the membrane does not control the permeation process).
Table 8. Summary of Pd-based thin membranes’ performances and respective characteristics (H2 diffusion through the membrane does not control the permeation process).
Membrane/SupportTemperature (°C)δ (µm)H2 Permeance (mol·m−2·s−1·Pa−1) (a)Selectivity (H2/N2)Ref.
Pd-7 wt.% Ag/Al2O34000.781.14 × 10−5640[362]
Pd-15 wt.% Ag/Al2O34004.04.2 × 10−620,000[363]
Pd-5 wt.% Ag/Al2O34001.04.6 × 10−625,938[294]
Pd-23 wt.% Ag/ZrO24001.08.0 × 10−6500[364]
Pd-Ag/Al2O34001.39.0–9.4 × 10−61900[365]
Pd-7-8 wt.% Au/Al2O35002-36.2 × 10−61400[366]
(a) Permeance values have been calculated for a H2 partial pressure of 1 bar.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Rocha, C.; Soria, M.A.; Madeira, L.M. Olive Mill Wastewater Valorization through Steam Reforming Using Multifunctional Reactors: Challenges of the Process Intensification. Energies 2022, 15, 920. https://doi.org/10.3390/en15030920

AMA Style

Rocha C, Soria MA, Madeira LM. Olive Mill Wastewater Valorization through Steam Reforming Using Multifunctional Reactors: Challenges of the Process Intensification. Energies. 2022; 15(3):920. https://doi.org/10.3390/en15030920

Chicago/Turabian Style

Rocha, Cláudio, Miguel Angel Soria, and Luís M. Madeira. 2022. "Olive Mill Wastewater Valorization through Steam Reforming Using Multifunctional Reactors: Challenges of the Process Intensification" Energies 15, no. 3: 920. https://doi.org/10.3390/en15030920

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