Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Synthesis and characterization of a new monometallic layered double hydroxide using manganese

Damien Cornu *a, Romain Coustel a, Guillaume Renaudin b, Guillaume Rogez c, Aurélien Renard a, Pierrick Durand d, Cédric Carteret a and Christian Ruby *a
aUniversité de Lorraine, CNRS, LCPME, F-54000 Nancy, France. E-mail: damien.cornu@univ-lorraine.fr; christian.ruby@univ-lorraine.fr
bUniversité Clermont Auvergne, Clermont Auvergne INP, CNRS, ICCF, F-63000 Clermont-Ferrand, France
cUniversité de Strasbourg, CNRS, IPCMS, F-67000 Strasbourg, France
dUniversité de Lorraine, CNRS, CRM2, F-54000 Nancy, France

Received 10th June 2022 , Accepted 8th July 2022

First published on 12th July 2022


Abstract

This article reports for the first time the synthesis of an LDH using only manganese as the divalent and trivalent metallic ion. Analysis of the pH, redox potential, and chemical composition during the oxidation of a manganese basic salt using persulfate indicates the oxidation of 1/3 of the initial MnII ions, in agreement with the paramagnetic structure and XPS analysis. Infrared, Raman spectra and thermogravimetric analysis results were similar to the ones obtained with Fe-LDH also known as green rust. X-Ray diffractograms and Rietveld refinement were used to determine the structure of this solid. Thermodynamic considerations predict that this solid could reduce nitrate into gaseous nitrogen without further reduction to ammonium or ammonia unlike what is observed for Fe-LDH.


Introduction

Layered double hydroxides (LDHs) or anionic clays are a class of ionic solids with a layered structure and a general formula [MII1−xMIIIx(OH)2]x+[(An)x/n mH2O]x· in which MII and MIII are metallic cations and An is an intercalated anion. These materials are useful structures for many applications, e.g., catalysis (oxygen evolution reaction1 and photodegradation of pollutants2), anionic exchange3 or biomedical applications.4 They are usually synthesized using a trivalent cation distinct from the divalent cations e.g. Al3+ as MIII and Mg2+ as MII for the common hydrotalcite Al–Mg-LDH.5 But it is possible to synthesize such a structure using only one metallic element. This is well known for iron, Fe-LDH, also known as “green rust”6,7 or “fougerite”.8,9 Sulfate Fe-LDH is very active in reducing nitrate10 or other pollutants.11 This mono-metallic layered structure is also observed for cobalt,12,13 forming single transition metal hydroxides with mixed valences. The monometallic Co-LDH can be used as capacitors with high pseudocapacitive performance14 or for the oxygen evolution reaction. In addition, cerium monometallic LDHs exist with interesting photocatalytic properties15 but because of the high oxidation state of Ce (Ce3+–Ce4+), sulfate ions are not simply in the interlayer space but also directly connected to the cerium. Other lanthanide-LDHs have also been synthesized.16

Manganese is frequently incorporated into LDH structures as a divalent cation with Al3+ (ref. 17–19) or as a trivalent cation with Mg2+, Zn2+ or Co2+ (ref. 20–23) because of its photocatalytic potential.24 These LDHs can be used for phototherapy against cancers,25 for the electrocatalytic detection of hydrogen peroxide23 or as an electrode for the oxygen evolution reaction.26 However, to our knowledge, it was never used as a single metallic constituent of LDH. One of the main limitations is the impossibility to obtain a stable Mn3+ solution excluding the classical coprecipitation method using an initial solution prepared by the dissolution of MnII and MnIII salts. The MnIII cations contained in the LDH were obtained either by introducing an oxidant such as H2O2 in the initial solution27 or probably due to the contact of the LDH suspension with air.28,29

Besides coprecipitation, another pathway for the synthesis of LDH is the oxidation of MII hydroxide. Mn(OH)2 oxidation was thoroughly studied half a century ago30 and it appeared that the formation of oxide (Mn3O4), birnessite31 or oxide-hydroxide MnO(OH) was the sole fate. However, a new MnII salt was discovered recently with sulfate in its structure.32 In this article, this salt was successfully transformed into Mn-LDH with sulfate anions in the interlayer space.

The possibility to stabilise a mixed MnII–MnIII species in a hydroxide structure opens new perspectives because of their redox properties. Indeed, soluble Mn2+(aq) species are well known to be very slowly oxidised by soluble oxygen in acidic or neutral aqueous solutions. MnII species present in more compact solid structures such as spinels (e.g. hausmannite) are well known to be stable in contact with air. In contrast, hydroxylated MnII species present in an open solid structure such as LDH are expected to be much more reactive. Such differences in reactivity were clearly demonstrated in the case of FeII containing compounds, i.e., the oxidation of Fe(OH)2 or Fe3O4 by NO3 is extremely slow in comparison to the nitrate oxidation of Fe-LDH.33 Finally, the existence of a mixed MnII–MnIII LDH structure has to be considered when analysing the oxidation of MnII in geology. Indeed, the source of MnIII in aqueous media is still debated.34

Results

Oxidation of manganese hydroxide sulfate solution

Firstly, Mn4(OH)6SO4 or Mn(OH)2 was produced by adding 1 mol L−1 NaOH solution into 0.4 mol L−1 MnSO4 solution. The ratio image file: d2dt01835g-t1.tif indicates the proportion of the NaOH amount in the solutions. For R values below 1.5, Mn4(OH)6SO4 is produced,32 while Mn(OH)2 was produced for higher R values. As long as the solids obtained for various R values with MnSO4 are kept under nitrogen bubbling, the suspensions remain white while their pH values are stable for hours.

For a given R value, Na2S2O8 solution is then added drop by drop. Subsequently, the solids darken upon the addition of the oxidant (ESI Fig. S1) except for R = 0 (no addition of OH). This is in agreement with the fact that Mn2+ is not oxidized because of the large free energy barrier to overcome.35 The pH value decreases upon S2O82− addition (ESI Fig. S2).

The R value is now fixed at 1.5 in order to maximise the amount of initial Mn4(OH)6SO4. The progress of the addition of S2O82− in the solution is defined by the parameter image file: d2dt01835g-t2.tif. According to the stoichiometry of the oxidation reaction, if the reaction is complete, then x is also equal to the fraction of the MnIII valence in the MnII and MnIII mixture as one persulfate accepts two electrons to produce sulfate ions.

Fig. 1 indicates the pH and the potential Eh/SCE for increasing x values. Three different zones can be observed: the A zone ends with the inflection for x = 0.266 (inflection point for other R values are discussed in ESI section 2), the B zone, in which one can observe a pseudo-plateau for Eh and pH, and the C zone with a final drop.


image file: d2dt01835g-f1.tif
Fig. 1 pH and Eh/silver chloride electrode values for a R = 1.5 MnSO4 solution as a function of x.

For various x values, we analysed the remaining quantities of aqueous cations and anions in solution. Quantities are indicated in Fig. 2 and were measured by inductively coupled plasma mass spectrometry (aqueous Mn species) and ion chromatography (SO42−).


image file: d2dt01835g-f2.tif
Fig. 2 Soluble Mn (black circles) and SO42− (green triangles), from ICP-MS measurement and ion chromatography, respectively, for R = 1.5 MnSO4 solution as a function of the x ratio. Error bars for Mn are smaller than the mark (0.1 mmol). Dotted lines show the expected quantities in solution for the formation of Mn-LDH through eqn (2) for x between 0 and 0.25 and then the transformation of this solid into MnO(OH) for x between 0.25 and 1 (For interpretation of the references to colour in this figure legend, the reader is referred to the Web version of this article.).

4 possible chemical reactions between MnII4(OH)6SO4 and the oxidant are listed below:

 
image file: d2dt01835g-t3.tif(1)
 
image file: d2dt01835g-t4.tif(2)
 
image file: d2dt01835g-t5.tif(3)
 
image file: d2dt01835g-t6.tif(4)

Eqn (1) and (2) describe the formation of a new type of LDH containing only Mn (Mn-LDH) equivalent to the green rust (Fe-LDH) which presents a FeII/FeIII ratio of 2. Water molecules within the Mn-LDH are omitted for the moment in the chemical formula that will be written as MnII4MnIII2(OH)12SO42−. For eqn (1), the entire Mn atoms are incorporated in the solid, which is not in agreement with what is observed in Fig. 2. This possibility is therefore rejected. For eqn (2), there is a simultaneous release of Mn2+ cations in the solution. Eqn (3) and (4) describe the formation of the well-known mixed oxide MnIIMnIII2O4 and oxyhydroxide MnIIIO(OH).

For x values lower than 0.25, the remaining quantities of Mn and sulfate in the solution are in clear agreement with the expected values from eqn (2). In addition, the stability of the pH during this phase indicates that there is little consumption of OH in this reaction. Besides, the inflections for other R values are also in agreement with this stoichiometry (ESI section 2). For x = 0.25, the maximum quantity of Mn-LDH is obtained.

For values between 0.25 and 0.75, the pseudo-plateau in the pH values can be attributed to the transformation of Mn-LDH into MnO(OH) (as determined by XRD in the next section) as in eqn (5). The oxidation reaction of Mn-LDH does not involve any change in the acid–base concentrations:

 
image file: d2dt01835g-t7.tif(5)

Again, the remaining quantity of Mn in solution is in agreement with this chemical reaction. However, the quantity of sulfate is much lower than what is predicted. It is therefore very likely that, along with MnO(OH) and Mn3O4, an additional solid precipitate was also formed. The large peak for the sulfate vibrations in Raman for R = 1.00 (Fig. 4) is in agreement with this hypothesis.

It is interesting to note that the production of the fully oxidized metal(III) oxyhydroxide species is also observed for further oxidation of Co14 and Fe36 monometallic LDHs.

Eqn (2) and (5) together indicate that half of the initial manganese is released as Mn2+ cations that cannot be oxidized under our conditions (ESI p.4). Oxidation to Mn3O4 is also possible and observed during the degradation of Mn-LDH in contact with air.

Characterization of the oxidation products

XRD. Mn-LDH formation in the previous section is proposed due to stoichiometric considerations. The X-ray diffractograms of the wet solids produced for R = 1.5 and various x values are presented in Fig. 3. For x = 0, pyrochroite (Mn(OH)2) and Mn4(OH)6SO4[thin space (1/6-em)]32 patterns are observed, which is no longer the case for x = 0.12 (Fig. 3). Instead, a similar basic MnII salt is observed: [Mn(OH)2]7[MnSO4]2·H2O. This structure was also observed upon drying Mn4(OH)6SO4.32 Alongside this structure, other diffraction peaks are present which are then the only ones to be observed for x = 0.24. The XRD pattern for x = 0.24 is characteristic of layered LDH compounds with the first (00l) diffraction peaks indicating an interlayer distance of about 11 Å. A Le Bail (profile matching) mode of refinement using the usual space group for LDH compounds, which is the R[3 with combining macron]m trigonal space group with the a0 planar lattice parameter close to 3.2 Å clearly indicated the absence of a rhombohedral network (ESI Fig. S4a). A second attempt based on the P[3 with combining macron]m1 space group corresponding to the brucite like Mn(OH)2 crystal structure37 led to a reliable refinement with the following hexagonal refined unit cell parameters: a0 = 3.2270(7) Å and c = 10.9390(23) Å (indexation is shown in ESI Fig. S4b). The refinement obtained was relevant although one low intensity diffraction peak (close to 18.3°, see # on Fig. 3) was not considered, suggesting the presence of a superstructure. Similar to the Fe-LDH phase intercalated with SO42−,37 a superstructure in the basal plane with image file: d2dt01835g-t8.tifa0 and the P[3 with combining macron]1m space group was tested. The obtained Le Bail simulation was improved by including all the diffraction signals (ESI Fig. S4c). The definitive refined unit cell parameters are as follows: a = 5.5917 (1) Å and c = 10.9492 (4) Å, very close to those reported for the Fe-LDH phase intercalated with SO42−:38a = 5.5524 (1) Å and c = 11.011 (3) Å.
image file: d2dt01835g-f3.tif
Fig. 3 XRD of the solids obtained from MnSO4 solution with R = 1.5 for various x ratios. Stars represent the main peaks for the [Mn(OH)2]7[MnSO4]2·H2O structure (JCPDS file 00-018-0788). # indicates the additional peak due to the Mn-LDH superstructure.

For x = 1, the diffractogram shows larger peaks that can be attributed to poorly crystalline manganite γ-MnO(OH) and groutite α-MnO(OH).

Vibrational spectroscopy. The evolution of Raman spectra of the wet structures for R = 1,5 and increasing the x amount of persulfate are shown in Fig. 4. For the x = 0 spectrum, the pyrochroite identified in XRD corresponds to the peak at 637 cm−1, as referenced in the RRUFF database R10004.39 The other main peak at 525 cm−1, as well as the sharp sulfate peak at 997 cm−1, indicates the presence of Mn4(OH)6SO4.32 For x = 0.24, two peaks at 431 and 523 cm−1 are observed, corresponding to the Eu(T) and the A1g(T) vibrations in LDH.40 Those values are close to what is measured with the sulfate Fe-LDH (427 and 518 cm−1),8 which suggest the presence of only one solid.
image file: d2dt01835g-f4.tif
Fig. 4 Raman spectra of the wet solids obtained using the R = 1.5 manganese sulfate solution for increasing x values.

As the oxidation continues, a broad peak centered at around 592 cm−1 emerges. This peak can be attributed to the vibrations of a mixture of α-MnO(OH) at 550 cm−1 (ref. 41) and γ-MnO(OH) (558 cm−1 for the symmetric stretching of the Mn–O–Mn bridge and 621 cm−1 for the asymmetric stretching of these bridges).42 It is noteworthy that the oxidation of the solid is not homogeneous and pure LDH as well as pure MnO(OH) spots can be found while translating the laser spot on the x = 0.6 solid.

The last spectrum for x = 1 shows no more vibration associated with the LDH, and a new peak emerges at 660 cm−1 which can be attributed to the Mn3O4,41 structure that was not observed in XRD.

Infrared spectroscopy confirms the transformation of the basic salt into LDH, with both compounds containing sulfate anions in their structure, followed by subsequent oxidation to oxide and oxide-hydroxide (ESI section 3.2).

XPS. XPS is performed on the dried product. Unfortunately, the attempts to wash the solid to remove sodium sulfate impurities were not successful as the solid degraded in pure water solution. Using XPS, we were able to measure the proportion of the dried solid that is sodium sulfate using the relative proportion of oxygen, manganese, sulfur, and sodium. In this dry unrinsed sample, we found that 25% of the mass is Na2SO4, which is consistent with the Na2SO4 concentration measured in the preparation medium. The remaining mass is in agreement with the formula Mn6(OH)12SO4·8H2O.

In Fig. 5, Mn 2p3/2 features the results from the contributions of various Mn valence states. Due to the coupling between the unpaired electrons of the outer shell and unpaired core electron resulting from photoionization, high spin Mn ions give rise to multiplet splitting. Overlapping of these multiplets poses a serious difficulty for qualitative and quantitative analysis of the Mn valence state.43–52Fig. 5 shows the Mn 2p3/2 peak at 642.0 eV with a shoulder at 643 eV, which suggests the coexistence of MnII and MnIII valence states. The shake-up satellite at 646 eV appears as the fingerprint of MnII in MnO,44–46 LDH,48 or Mn(OH)2 as shown in Fig. 5. The Mn 2p3/2 spectra of Mn(OH)2 and MnO(OH) (MnIII) as well as their linear combination with an area ratio of 38[thin space (1/6-em)]:[thin space (1/6-em)]62 were added as shown in Fig. 5. The latter fairly reproduces the spectrum of Mn-LDH confirming the mixed-valence MnII–MnIII nature of this compound, even if it fails to account for the expected MnII/MnII ratio of 2.


image file: d2dt01835g-f5.tif
Fig. 5 XPS in the Mn 2p3/2 region of the solids for R = 1.5 and x = 0.24 ratios (black dots), x = 1 (green open dots), and Mn(OH)2 (red cross), as well as their combination (grey line), are plotted.

O 1s, S 2p and valence band spectra are given in ESI section 3.4.

Thermogravimetric measurements. Thermogravimetric analysis (TGA) of the dried powder was performed (ESI Fig. S9). Two weight losses were observed, at 95 and 174 °C, which is again in the same range as that observed for Fe-LDH38 (117 and 153 °C), which can be attributed to the vaporization of slightly bonded water molecules present in the interlayer and to the condensation reaction between hydroxyl groups in the hydroxide sheets and their departure as water molecules, respectively. Those mass losses (13.5 and 8%, respectively, of the total mass) correspond respectively to 18% and 10.5% of the LDH mass in the dried product determined by XPS analysis (75% of the dried product is Mn-LDH). This indicates that there are 8 water molecules per unit slightly bonded in the dried structure (8.5 in the Fe-LDH) and 4.5 additional water molecules leave at higher temperatures because of the condensation of 9 hydroxyl groups (2 water molecules and 4 hydroxyl groups for the Fe-LDH) associated with the formation of MnO(OH) and Mn2O3 at higher temperatures. This indicates that the complete structure of the solid is Mn6(OH)12SO4·8 H2O.
Magnetic properties. The susceptibility was recorded under a 0.5 T dc field (ESI Fig. S10) as a function of temperature.

The Curie constant, determined from the fit of the 1/χ = f(T) curve using the Curie–Weiss law, is C = 18.7 emu K mol−1 (Fig. 6). This value is in the same order of magnitude as the one expected from the elemental formula, oxidation degrees, and composition of the solid phase (23.5 emu K mol−1 considering a high-spin configuration for MnII and MnIII with g = 2). The negative Weiss temperature, θ = −105 K, indicates the occurrence of predominant antiferromagnetic interactions.


image file: d2dt01835g-f6.tif
Fig. 6 1/χ = f(T) (black circles) for compound LDH-Mn, under a dc field of 0.5 T and the best fit of the high temperature region (>150 K) using the Curie–Weiss law (full red line).

The χT product (ESI Fig. S10) decreases when the temperature decreases slowly down to a minimum at 67 K (9.2 emu K mol−1). Then, the χT product increases sharply to two successive maxima, at 33.5 K and 16.2 K (49.3 and 44.7 emu K mol−1 respectively). This overall behavior points to the occurrence of ferrimagnetic ordering, with two magnetic phases. One hypothesis to explain the presence of two magnetic phases can be the existence of phases with different hydration levels, which may modify the interaction between the layers, and hence the magnetic coupling leading to different ordering temperatures.

The magnetic ordering is better evidenced by ac susceptibility measurements, performed under a zero dc field, and a 2 Oe oscillating field (ESI Fig. 11). These measurements reveal a rather complex ordering phenomenon, with five peaks in the out-of-phase susceptibility (at 39.5, 37.4 (shoulder), 14.2, 6.1 and 2.0 K).

None of these peaks are frequency dependent, which exclude the signature of potential small paramagnetic clusters with slow magnetic relaxation. Therefore, ac measurements evidence five ferro or ferrimagnetic ordering temperatures, likely corresponding to the presence of five magnetic phases, and not only two as evidenced by dc measurements. This distribution of ordering temperatures is not unusual for layered hydroxide based magnets;53–55

Finally, magnetization vs. field at a low temperature (1.8 K) is presented in Fig. 7. At a high field, the magnetization varies linearly with the field and is far from the expected value for a full spin alignment, as expected for a ferrimagnetically ordered material. The coercive field is 88 mT at 1.8 K.


image file: d2dt01835g-f7.tif
Fig. 7 M = f(H) at 1.8 K.
SEM. Scanning electron microscopy was also used to characterize Mn-LDH (Fig. 8). Some ribbon structures can be seen, with 0.5 μm width, a few μm length and 0.1 μm thickness. This kind of structure could not be seen in the Mn4(OH)6SO4 salt (ESI Fig. S5). This ribbon structure is different from what is observed for Fe56 and Co-LDH,13 in which a hexagonal morphology was reported in SEM. The small roundish shapes of residual crystallites could be attributed to the part of the LDH oxidized in Mn3O4.
image file: d2dt01835g-f8.tif
Fig. 8 SEM images of Mn-LDH obtained for x = 0.25.

The preferential direction for the growth of Mn-LDH could be related to the anisotropy created by the Jahn–Teller effect in the MnIII octahedron,57 even if this effect could not be seen in the Rietveld analysis.

Discussion

Structure of the obtained product

Chemical analysis of the solution, as well as the variations of the pH (ESI section 2) during the oxidation of Mn4(OH)6SO4, shows that the stoichiometry of the reaction is compatible with the formation of Mn-LDH with a chemical composition of MnIII2MnII4(OH)12SO4·xH2O. X-ray diffraction and Raman spectroscopy show similar results for the Fe- and Mn-LDH intercalated by sulfate. Thermogravimetric analysis indicates that 8 water molecules are present in the interlayer space for each Mn-LDH motif based on two Mn2+, leading to the formula: MnII2MnIII4(OH)12SO4·8H2O. The XRD pattern was therefore treated by Rietveld refinements using the structural data of Fe-LDH intercalated by sulfate as a starting point. This initial structural model was efficient, and refinement cycles were carried out. Careful work has been done on the refinement of the diffraction peak profile, using anisotropic size parameters linked to the platelet shaped LDH crystallites. In the course of refinement the dS–O interatomic distances for sulfate anions were fixed at the 1.55 (1) Å value with a regular tetrahedral shape. Fig. 9 shows the obtained Rietveld plot and Table 1 shows the refined structural parameters of the Mn-LDH compound compared with those of the Fe-LDH analog (ESI Table 2).
image file: d2dt01835g-f9.tif
Fig. 9 Rietveld plot for the refined structural model of the Mn-LDH compound intercalated with sulfate anions showing the diffraction measured pattern (red dots), the calculated pattern (black line), the difference curve (blue line) and sticks for the Bragg peak positions.
Table 1 Refined structure of Mn-LDH
Parameters Mn-LDH (SO42−)
Space group P[3 with combining macron]1m
Lattice parameters a = 5.58821 (9) Å
c = 10.9410 (4) Å
Atomic positions
Mn3+ (1a) (0 0 0)
Mn2+ (2c) (⅔ ⅓ 0)
OH (6k) (0.327(3) 0 0.0935(6))
H2O (12l) (−0.248(3) 0.564(2) 0.6624(5))
S6+ (2e) (0 0 0.6054(8))
OA2− (2e) (0 0 0.7481(7))
OB2− (6k) (0.275(2) 0 0.573(1))
Site occupancies
H2O (12l) 0.39(3)
Sulfate (2e) 0.25(−)
Reliability factors
R Bragg 0.031
R F 0.060
R p 0.027
R wp 0.036
R exp 0.026


The origin of this ordering is in the interlayer in which sulfate anions are located at specific positions apical to trivalent Mn3+ cations. Having a Mn3+/SO42− ratio of 2 explains the formation of double planes within the interlayer; each Mn3+ octahedron being bounded only to one sulfate anion by a single face of the main layer (Fig. 10). The rest of the interlayer region is occupied by relatively mobile water molecules, and the refined occupancy for water molecules on the Wyckoff site corresponds to the following chemical composition: MnIII2MnII4(OH)6·SO4·9H2O, close to the one deduced by thermogravimetric analysis (MnIII2MnII4(OH)6·SO4·8H2O) Fig. 10 presents the crystal structure of the Mn-LDH synthesized compound. The superstructure of this sulfate intercalated Mn-LDH phase along the [110] direction of the Mn(OH)2 basic structure37 allows the ordering of the divalent and trivalent manganese cations, respectively, in the 2c and 1a Wyckoff positions (Table 1) which impose a Mn2+/Mn3+ ratio of 2.


image file: d2dt01835g-f10.tif
Fig. 10 Representation of the Mn-LDH crystal structure. Main layers composed of Mn3+ (brown) and Mn2+ (purple) octahedra, intercalated by sulfate anions (yellow tetrahedra) and water molecules (blue spheres).

Redox properties

The redox potential Eh for the oxidation of {Fe(OH)2(s),Fe2+(aq)} solution by air was used to determine the standard Gibbs free energy of Fe-LDH.38 For example, a value ΔfG0(Fe-LDH) = −3790 ± 10 kJ mol−1 was determined by this method for Fe-LDH with sulfate in good agreement with other works.58 Unfortunately, as the basic salt Mn4(OH)6SO4 is not referenced, the first step of oxidation (A part of Fig. 1) transforming it into Mn-LDH cannot be used to determine its standard Gibbs free energy of formation: ΔfG0(Mn-LDH) value.

In contrast, the redox potential values recorded in zone B (0.4 < x < 0.7) were relatively constant and a pseudo Eh plateau was observed. Zone B corresponds to the transformation of the Mn-LDH into MnO(OH). In fact, the product is a mixture of manganite γ-MnO(OH) and groutite α-MnO(OH). The value E0(MnO(OH)/Mn-LDH) = +1.117 V/SHE was computed (ESI section 3.1) by using the experimental Eh/ESH value (+380 mV) and pH value (8.2) recorded in the middle of the Eh plateau of zone B and corrected with the relative silver chloride electrode potential (Fig. 1).

By using the ΔfG0 values reported in ESI Table S4, the standard Gibbs energy of formation of MnII4MnIII2(OH)12SO4 was estimated to be in between −4519 and −4436 kJ mol−1, the lower and the higher values being determined for the final oxidation products manganite or groutite, respectively. The Pourbaix diagrams of Fe and Mn species were built by using the standard chemical potential of ESI Table S4 and the Nernst equations of the relevant redox couples with the details of the calculations in section 4.1.

The reduction reaction of nitrate to either ammonium or dinitrogen by oxidizing Fe-LDH to γ-FeO(OH) is thermodynamically favorable. Previously performed experiments are in agreement with such a prediction and Fe-LDH was observed to reduce nitrate either in ammonium10 for sulfate and chloride containing Fe-LDH or into a mixture of ammonium and N-gaseous species for carbonate containing LDH.59

Interestingly, the limit between the domain of stability of nitrate and dinitrogen is still situated above the one separating MnO(OH) and Mn-LDH on the Mn-Pourbaix diagram (Fig. 11A), while it is not the case for the line separating nitrate and ammonium/ammonia. This means that Mn-LDH is not reactive enough to reduce nitrate into ammonium, but it may potentially reduce nitrate into N2. This is in line with what is observed in the manganese assisted denitrification process in a laboratory scale sequencing batch reactor by Swathi et al.60


image file: d2dt01835g-f11.tif
Fig. 11 Pourbaix diagrams for Mn (A) and Fe (B).

The latter reaction is of utmost importance if the goal is to find a material useful for the water denitrification process. Such properties could be unique among the monometallic LDH family. Indeed, Fe-LDH is a too strong reductant that transformed nitrate mainly into the more reduced form of nitrogen, i.e. NH4+. In contrast, Co-LDH and Ni-LDH could be too weak reductants to transform NH4+ into N2. Such a preliminary forecast can be proposed by comparing similar redox potentials, e.g. the standard potentials E0 of M2+(aq)/MIII2O3(s) redox couples are equal to 0.728, 1.443, 1.746 and 1.753 V/SHE when M = Fe, Mn, Co and Ni, respectively.61 However, further experiments dedicated to the determination of the standard chemical potentials of Co-LDH and Ni-LDH and to the reactivity of monometallic LDH towards nitrate species should be explored to confirm these assumptions.

Conclusions

A new monometallic manganese double layered hydroxide MnIII2MnII4(OH)6·SO4·8H2O was discovered using MnII4(OH)6SO4 oxidation with S2O82− and following the pH, Eh and composition of the liquid solution during the oxidation. The characterization (XRD, Raman, infrared, and thermogravimetric analysis) of the solid shows that the structure of Mn-LDH is very close to the one of Fe-LDH (green rust). XPS shows that the solid contains a mixture of MnII and MnIII, which is confirmed by magnetic experiments. SEM shows a ribbon shape organization of the solid, perhaps linked with a preferential direction due to the Jahn–Teller effect induced by MnIII ions. The crystal structure was therefore determined using Rietveld refinement. The solid was determined to be ferrimagnetic at low temperatures.

Pourbaix diagrams calculated from the experimental measurements during the synthesis indicate that this solid is theoretically able to reduce nitrate into dinitrogen without further reduction to ammonium or ammonia. This indicates that this solid could be used for water remediation. Finally, the eventual formation of Mn-LDH in the natural environment should also be considered. Indeed, fougerite, the mineral containing Fe-LDH, is already identified in hydromorphic soils and groundwater. MnII–MnIII LDH could be an intermediate species in the oxidation pathway of manganese.

Experimental

Synthesis

Mn-LDH were synthesized using 50 mL of MnSO4·H2O (Sigma Aldrich, >99%) solution (0.4 mol L−1) with a variable volume of 1 mol L−1 NaOH (VWR 31627.290) to reach the desired R value image file: d2dt01835g-t9.tif under stirring and nitrogen bubbling. Once the pH was stabilized, a 0.2 mol L−1 solution of Na2S2O8 (Sigma Aldrich, >99%) was added using a peristaltic pump with a flow of 0.167 mL min−1 and the pH and the redox potential Eh were registered with a Metrohm pH electrode Unitrode with Pt1000 and a Toledo Inlab redox electrode. The x value indicates the quantity of persulfate introduced, as image file: d2dt01835g-t10.tif. For XPS analysis, Mn(OH)2 was prepared from a mixture of 1.12 g of MnCl2·4H2O (Sigma Aldrich, >98%) and 0.57 g of NaOH (VWR, 99.1%) in 40 ml of water under nitrogen bubbling.

Characterization studies

ICP-MS. Released manganese concentrations in the solution were measured by Inductively Coupled Plasma Mass Spectroscopy (ICP-MS, Agilent-7800). 100 μL of the supernatant solution after centrifugation was diluted to 100 mL with water. The calibration curve in the range 50–2000 ppb has been established using Sigma-Aldrich 74128 samples. The detected isotope was 55Mn.
Ion chromatography. Sulfate concentration was determined by ion chromatography. Analysis was performed with a Metrohm 882 Compact IC plus instrument equipped with chemical (Metrohm suppressor MSM II) and sequential (Metrohm CO2 suppressor MCS) suppressors and a conductimetric detector. The eluent (1.8 mmol L−1 Na2CO3 and 1.7 mmol L−1 NaHCO3 solution) flowed continuously through a Metrosep A Supp 4–250/4.0 column associated with a guard column (Metrosep A supp 4/5 guard). Standard solutions were prepared from commercial 1000 μg mL−1 anions (Merck 1.11448.0500). The flow rate and the injected volume were respectively 1 mL min−1 and 20 μL.
XRD. The products were measured soon after their synthesis and centrifugation. XRD diffractograms were recorded either on a wet sample (x = 0, 0.12 and 0.24) or on the sample dried under nitrogen flow (x = 1). For the wet sample, excess water was removed by pressing the water soaked paste between tissues. The resulting paste was measured on an approximately 10 μm thick zero background X-ray holder covered with a minimal amount of glycerol to protect the paste from air oxidation, as the compounds were known to be air sensitive. Multiple (at least 15) 1 h X-ray scans were collected, and the diffractograms did not show any sign of evolution, the multiple scans were summed up for better X-ray statistics.

Powder X-ray diffraction patterns were recorded with a Panalytical X'Pert Pro MPD diffractometer in reflection geometry using a tube with Cu radiation (Kα1 = 1.5406 Å, a Ge(111) incident-beam monochromator, 0.02 rad Soller slits, programmable divergence and antiscatter slits (the irradiated area was fixed to 10 mm × 10 mm)), and an X'Celerator detector. Data were collected from finely ground samples with a sample holder spinner and continuous rotation of the sample to improve the statistical representation of the sample.

XPS. X-ray photoelectron spectra were recorded on a KRATOS Axis Ultra Xray photoelectron KRATOS Axis Ultra DLD spectrometer equipped with a monochromated Al Kα source ( = 1486.6 eV, spot size 0.7 mm × 0.3 mm). The detector was a hemispherical analyzer at an electron emission angle of 90° and a pass energy of 20 eV (high resolution spectra). For the high-resolution spectra, the overall energy resolution, resulting from the monochromator and electron analyzer bandwidths, was higher than 800 meV. The C 1s binding energy of adventitious carbon was taken as 284.6 eV. O 1s spectra were decomposed using Gaussian peaks after using the Shirley method of background subtraction.
Vibrational spectroscopy. Infrared IR measurements of the wet solids were recorded under nitrogen flow in an attenuated total reflectance mode on a Bruker Tensor 27 spectrometer equipped with a KBr beam splitter and a deuterated triglycine sulfate (DTGS) thermal detector. Spectra were recorded and processed using OPUS 7.5 software (Bruker, Karlsruhe, Germany).

10 μL samples for Raman spectra were collected and dried over an aluminium plate. Then Raman spectra were recorded on a Renishaw inVia Qontor microspectrometer equipped with a confocal microscope and an Olympus X50 objective (N.A = 0.55). It should be noted that a nitrogen atmosphere was used to avoid exposure to oxygen. A 532 nm exciting radiation was used with a laser power below 0.05 mW for all samples to prevent their degradation. The spot area was a few μm2. Several locations were probed on each sample. The spectral resolution was about 4 cm−1 and the precision on the wavenumber was lower than 1 cm−1.

SQUID. Magnetic measurements were performed at the Institut de Physique et Chimie des Matériaux de Strasbourg (UMR CNRS-Unistra 7504) using a Quantum Design MPMS3 magnetometer. Magnetization measurements at different fields at a given temperature confirm the absence of ferromagnetic impurities. Data were corrected for the sample holder and diamagnetism was estimated from Pascal constants.
Thermogravimetric measurements. TGA was performed on a TG/ATD 92-16.18 SETARAM instrument. About 50 mg were introduced into the cell under 2 L h−1Ar with 2° min−1 ramp from 20 to 600 °C. The mass loss was corrected with a reference analysis performed using a second heating ramp.
SEM. Scanning Electron Microscopy (SEM) analysis was performed on a JEOL JSM-IT500HR, with a Field Emission Gun (FEG). The powder sample was fixed on double face scotch tape. The analysis was made under high vacuum, and we settled on a 60 μm diaphragm aperture. The voltages for analysis were from 2 kV to 5 kV. We used mainly the Secondary Electron Detector (SED) for imaging the powder.

Author contributions

DC: Conceptualization, investigation, writing – original draft, and writing – review & editing; RC, GRo, and GRe: investigation and writing – review & editing; PD and AR: investigation; CC and CR: writing – review & editing.

Data availability

Most of the data and figures of this publication were deposited in the DOREL database: https://dorel.univ-lorraine.fr/dataset.xhtml?persistentId=doi:10.12763/TLLN2F

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

The authors thank Martine Mallet for her help with the XPS analysis. XPS, Raman, and FTIR spectroscopy measurements were performed at the spectroscopy and microscopy service facility of SMI LCPME (Université de Lorraine-CNRS). The authors thank the PMD2X X-ray diffraction facility of the Institut Jean Barriol, Université de Lorraine, for X-ray diffraction measurements, data processing and analysis, and for providing the reports for publication.

In addition, the authors thank Claire Genois and Christelle Despas from LCPME for the ICP-MS analysis and the ion chromatography analysis. The authors also thank Lionel Aranda from Institut Jean Lamour for the thermogravimetric analysis.

Notes and references

  1. Z. Cai, X. Bu, P. Wang, J. C. Ho, J. Yang and X. Wang, J. Mater. Chem. A, 2019, 7, 5069–5089 RSC.
  2. G. Zhang, X. Zhang, Y. Meng, G. Pan, Z. Ni and S. Xia, Chem. Eng. J., 2020, 392, 123684 CrossRef CAS.
  3. M. Zubair, M. Daud, G. McKay, F. Shehzad and M. A. Al-Harthi, Appl. Clay Sci., 2017, 143, 279–292 CrossRef CAS.
  4. Y. Kuthati, R. K. Kankala and C.-H. Lee, Appl. Clay Sci., 2015, 112–113, 100–116 CrossRef CAS.
  5. M. V. Bukhtiyarova, J. Solid State Chem., 2019, 269, 494–506 CrossRef CAS.
  6. W. Feitknecht and G. Keller, Z. Anorg. Allg. Chem., 1950, 262, 61–68 CrossRef CAS.
  7. J. D. Bernal, D. R. Dasgupta and A. L. Mackay, Clay Miner. Bull., 1959, 4, 15–30 CrossRef CAS.
  8. F. Trolard, J.-M. R. Génin, M. Abdelmoula, G. Bourrié, B. Humbert and A. Herbillon, Geochim. Cosmochim. Acta, 1997, 61, 1107–1111 CrossRef CAS.
  9. C. Ruby, A. Géhin, R. Aissa and J.-M. R. Génin, Corros. Sci., 2006, 48, 3824–3837 CrossRef CAS.
  10. H. Chr, B. Hansen, C. B. Koch, H. Nancke-Krogh, O. K. Borggaard and J. Sørensen, Environ. Sci. Technol., 1996, 30, 2053–2056 CrossRef.
  11. M. Usman, J. M. Byrne, A. Chaudhary, S. Orsetti, K. Hanna, C. Ruby, A. Kappler and S. B. Haderlein, Chem. Rev., 2018, 118, 3251–3304 CrossRef CAS PubMed.
  12. Z. P. Xu and H. C. Zeng, Int. J. Inorg. Mater., 2000, 2, 187–196 CrossRef CAS.
  13. R. Ma, K. Takada, K. Fukuda, N. Iyi, Y. Bando and T. Sasaki, Angew. Chem., 2008, 120, 92–95 CrossRef.
  14. P. Vialat, C. Mousty, C. Taviot-Gueho, G. Renaudin, H. Martinez, J.-C. Dupin, E. Elkaim and F. Leroux, Adv. Funct. Mater., 2014, 24, 4831–4842 CrossRef CAS.
  15. T. Ye, W. Huang, L. Zeng, M. Li and J. Shi, Appl. Catal., B, 2017, 210, 141–148 CrossRef CAS.
  16. J. Liang, R. Ma, F. Geng, Y. Ebina and T. Sasaki, Chem. Mater., 2010, 22, 6001–6007 CrossRef CAS.
  17. L.-M. Grand, S. J. Palmer and R. L. Frost, J. Therm. Anal. Calorim., 2010, 100, 981–985 CrossRef CAS.
  18. T. Coelho, R. Micha, S. Arias, Y. E. Licea, L. A. Palacio and A. C. Faro, Catal. Today, 2015, 250, 87–94 CrossRef CAS.
  19. C. Yang, L. Liao, G. Lv, L. Wu, L. Mei and Z. Li, J. Colloid Interface Sci., 2016, 479, 115–120 CrossRef CAS PubMed.
  20. S. Bhattacharjee and J. A. Anderson, Adv. Synth. Catal., 2006, 348, 151–158 CrossRef CAS.
  21. Y. Du, Q. Wang, X. Liang, Y. He, J. Feng and D. Li, J. Catal., 2015, 331, 154–161 CrossRef CAS.
  22. J. M. Fernandez, C. Barriga, M.-A. Ulibarri, F. M. Labajos and V. Rives, J. Mater. Chem., 1994, 4, 1117–1121 RSC.
  23. H. Farhat, C. Taviot-Gueho, G. Monier, V. Briois, C. Forano and C. Mousty, J. Phys. Chem. C, 2020, 124, 15585–15599 CrossRef CAS.
  24. Z. Timár, G. Varga, S. Muráth, Z. Kónya, Á. Kukovecz, V. Havasi, A. Oszkó, I. Pálinkó and P. Sipos, Catal. Today, 2017, 284, 195–201 CrossRef.
  25. Y. Ruan, X. Jia, C. Wang, W. Zhen and X. Jiang, Chem. Commun., 2018, 54, 11729–11732 RSC.
  26. F. Song and X. Hu, J. Am. Chem. Soc., 2014, 136, 16481–16484 CrossRef CAS PubMed.
  27. X. Zhao, C. Niu, L. Zhang, H. Guo, X. Wen, C. Liang and G. Zeng, Chemosphere, 2018, 204, 11–21 CrossRef CAS PubMed.
  28. R. Li, Y. Liu, H. Li, M. Zhang, Y. Lu, L. Zhang, J. Xiao, F. Boehm and K. Yan, Small Methods, 2019, 3, 1800344 CrossRef CAS.
  29. X. Wang, H. Yan, J. Zhang, X. Hong, S. Yang, C. Wang and Z. Li, J. Alloys Compd., 2019, 810, 151911 CrossRef CAS.
  30. R. Scholder and H. Kyri, Z. Anorg. Allg. Chem., 1952, 270, 56–68 CrossRef CAS.
  31. H. Boumaiza, R. Coustel, G. Medjahdi, C. Ruby and L. Bergaoui, J. Solid State Chem., 2017, 248, 18–25 CrossRef CAS.
  32. D. Cornu, R. Coustel, P. Durand, C. Carteret and C. Ruby, J. Solid State Chem., 2021, 305, 122631 Search PubMed.
  33. C. J. Matocha, P. Dhakal and S. M. Pyzola, in Advances in Agronomy, ed. D. L. Sparks, Academic Press, 2012, vol. 115, pp. 181–214 Search PubMed.
  34. C. M. Hansel, in Advances in Microbial Physiology, ed. R. K. Poole, Academic Press, 2017, vol. 70, pp. 37–83 Search PubMed.
  35. S. Namgung, C.-M. Chon and G. Lee, Geosci. J., 2018, 22, 373–381 CrossRef CAS.
  36. P. Refait, C. Bon, L. Simon, G. Bourrie, F. Trolard, J. Bessiere and J. M. R. Genin, Clay Miner., 1999, 34, 499–510 CrossRef CAS.
  37. A. N. Christensen and G. Ollivier, Solid State Commun., 1972, 10, 609–614 CrossRef CAS.
  38. L. Simon, M. François, P. Refait, G. Renaudin, M. Lelaurain and J.-M. R. Génin, Solid State Sci., 2003, 5, 327–334 CrossRef CAS.
  39. B. Lafuente, R. T. Downs, H. Yang and N. Stone, in Highlights in Mineralogical Crystallography, ed. T. Armbruster and R. M. Danisi, DE GRUYTER, Berlin, München, Boston, 2015, pp. 1–30 Search PubMed.
  40. A. Di Bitetto, Thèse: Etude structurale et dynamique d'hydroxydes doubles lamellaires: du matériau carbonaté aux hybrides organo-minéraux, 2017 Search PubMed.
  41. B. Hu, K. Lu, Q. Zhang, X. Ji and W. Lu, Comput. Mater. Sci., 2017, 136, 29–35 CrossRef CAS.
  42. P. F. Smith, B. J. Deibert, S. Kaushik, G. Gardner, S. Hwang, H. Wang, J. F. Al-Sharab, E. Garfunkel, L. Fabris, J. Li and G. C. Dismukes, ACS Catal., 2016, 6, 2089–2099 CrossRef CAS.
  43. A. A. Audi and P. M. A. Sherwood, Surf. Interface Anal., 2002, 33, 274–282 CrossRef CAS.
  44. E. S. Ilton, J. E. Post, P. J. Heaney, F. T. Ling and S. N. Kerisit, Appl. Surf. Sci., 2016, 366, 475–485 CrossRef CAS.
  45. M. C. Biesinger, B. P. Payne, A. P. Grosvenor, L. W. M. Lau, A. R. Gerson, R. St. and C. Smart, Appl. Surf. Sci., 2011, 257, 2717–2730 CrossRef CAS.
  46. H. W. Nesbitt and D. Banerjee, Am. Mineral., 1998, 83, 305–315 CrossRef CAS.
  47. H. Jung and Y.-S. Jun, Environ. Sci. Technol., 2016, 50, 105–113 CrossRef CAS PubMed.
  48. M. Asif, W. Haitao, D. Shuang, A. Aziz, G. Zhang, F. Xiao and H. Liu, Sens. Actuators, B, 2017, 239, 243–252 CrossRef CAS.
  49. J. O. Eniola, R. Kumar, A. A. Al-Rashdi, M. O. Ansari and M. A. Barakat, ACS Omega, 2019, 4, 18268–18278 CrossRef CAS PubMed.
  50. S. Werner, V. W. Lau, S. Hug, V. Duppel, H. Clausen-Schaumann and B. V. Lotsch, Langmuir, 2013, 29, 9199–9207 CrossRef CAS PubMed.
  51. T. Li, J. Wang, Y. Xu, Y. Cao, H. Lin and T. Zhang, ACS Appl. Energy Mater., 2018, 1, 2242–2253 CrossRef CAS.
  52. A. N. Simonov, R. K. Hocking, L. Tao, T. Gengenbach, T. Williams, X.-Y. Fang, H. J. King, S. A. Bonke, D. A. Hoogeveen, C. A. Romano, B. M. Tebo, L. L. Martin, W. H. Casey and L. Spiccia, Chem. – Eur. J., 2017, 23, 13482–13492 CrossRef CAS PubMed.
  53. F. Bellouard, M. Clemente-León, E. Coronado, J. R. Galán-Mascarós, C. J. Gómez-García, F. Romero and K. R. Dunbar, Eur. J. Inorg. Chem., 2002, 2002, 1603–1606 CrossRef.
  54. E. Coronado, J. R. Galán-Mascarós, C. J. Gómez-García and V. Laukhin, Nature, 2000, 408, 447–449 CrossRef CAS PubMed.
  55. Q. Evrard, Z. Chaker, M. Roger, C. M. Sevrain, E. Delahaye, M. Gallart, P. Gilliot, C. Leuvrey, J.-M. Rueff, P. Rabu, C. Massobrio, M. Boero, A. Pautrat, P.-A. Jaffrès, G. Ori and G. Rogez, Adv. Funct. Mater., 2017, 27, 1703576 CrossRef.
  56. I. A. M. Ahmed, L. G. Benning, G. Kakonyi, A. D. Sumoondur, N. J. Terrill and S. Shaw, Langmuir, 2010, 26, 6593–6603 CrossRef CAS PubMed.
  57. S. Hirai, S. Yagi, A. Seno, M. Fujioka, T. Ohno and T. Matsuda, RSC Adv., 2015, 6, 2019–2023 RSC.
  58. K. B. Ayala-Luis, C. B. Koch and H. C. B. Hansen, Clays Clay Miner., 2008, 56, 633–644 CrossRef CAS.
  59. M. Etique, A. Zegeye, B. Grégoire, C. Carteret and C. Ruby, Water Res., 2014, 62, 29–39 CrossRef CAS PubMed.
  60. D. Swathi, P. C. Sabumon and S. M. Maliyekkal, Int. Biodeterior. Biodegrad., 2017, 119, 499–510 CrossRef CAS.
  61. J. Van Muylder and M. Pourbaix, Atlas d'équilibres électrochimiques, Gauthier-Villars & Cie, Paris, 1963, p. 378 Search PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2dt01835g

This journal is © The Royal Society of Chemistry 2022
Click here to see how this site uses Cookies. View our privacy policy here.