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

Amorphous V2O5–P2O5 as high-voltage cathodes for magnesium batteries

Timothy S. Arthur *a, Keiko Kato a, Jason Germain a, Jinghua Guo b, Per-Anders Glans b, Yi-Sheng Liu b, Daniel Holmes c, Xudong Fan d and Fuminori Mizuno a
aMaterials Research Department, Toyota Research Institute of North America, 1555 Woodridge Ave., Ann Arbor, Michigan 48105, USA. E-mail: tim.arthur@tema.toyota.com
bAdvanced Light Source, Lawrence Berkeley National Lab, One Cyclotron Rd, Berkeley, California 94720, USA
cDepartment of Chemistry, Michigan State University, 578 Shaw Lane, East Lansing, Michigan 48824, USA
dCenter for Advanced Microscopy, Michigan State University, 578 Wilson Rd, East Lansing, Michigan 48824, USA

Received 25th August 2015 , Accepted 4th September 2015

First published on 4th September 2015


Abstract

A deep investigation of amorphous V2O5–P2O5 powders for magnesium batteries communicates the vital properties to achieving the superior electrochemical performance at a 75[thin space (1/6-em)]:[thin space (1/6-em)]25 V2O5[thin space (1/6-em)]:[thin space (1/6-em)]P2O5 molar ratio. The manipulation of the inter-layer spacing and amorphization of V2O5 can enhance Mg2+ diffusion and afford a cathode with high-voltage reversibility.


Multivalent magnesium batteries are viable candidates for post Li-ion technology due to the high volumetric capacity of the Mg metal anode (3833 mA h cm−3), high natural abundance, good safety characteristics and low equilibrium potential (−2.31 V vs. hydrogen).1 The promise of magnesium batteries can only be realized by overcoming two significant hurdles. First, in contrast to lithium metal systems, the chemical/electrochemical reactions at the interface between the Mg metal anode and common electrolytes, such as magnesium perchlorate or magnesium bis(trifluoromethanesulfonyl)imide in acetonitrile, inhibit the magnesium deposition and dissolution because the surface of the anode is passivated. Gregory and Winter,2 followed by Aurbach et al.,3 found that the passivating layer on the magnesium metal surfaces are non-existent or negligible when a magnesium organoborates and organohaloaluminates are used as the electrolytes. Since, significant improvements in voltage stability, purity, corrosion inhibition and current density have opened the door to search for high voltage, magnesium systems.4–7 The second major obstacle to realizing a complete Mg battery is the lack of high-voltage cathodes. Due to the divalent nature of Mg2+, high diffusion barriers and slow insertion/de-insertion kinetics have hindered the emergence of viable intercalation cathodes.8,9 Chevrel-phase MxMo6T8 (M = metal and T = S or Se) cathodes were first introduced as excellent prototype systems by Aurbach et al. and recent nanoparticle synthesis of Mo6S8 by Cheng et al.10 has shown superior capacity, rate and cycle-life. However, the low operating voltage (∼1.1 V) of the battery has encouraged further research. A magnesium battery with significant energy density demands high-voltage cathodes to compete with current Li-ion battery technology.

Although years of research have been dedicated to Mg battery cathodes, few impactful advances have been reported in the field. Notably, the emergence of layered V2O5 and MoO3, MoS2 nanosheets, polyanionic compounds and organic cathodes has renewed confidence in Mg battery cathodes.11–14 Recently, we have also reported on the importance of determining the magnesiation mechanism of cathodes which exhibit positive electrochemical results.15 Although the full potential of MnO2 polymorphs are still being explored, we were encouraged by the electrochemical reversibility exhibited by the amorphous mixture and thus, we considered the rational design of amorphous materials as a possible route to achieve high-voltage Mg battery cathodes.

Amorphous cathodes can provide several advantages for hosting the divalent Mg2+ ion. Recent calculations have shown that sluggish diffusion of Mg2+ is difficult to overcome in typical transition metal oxide (TMO) cathodes, and the loss of structural integrity of the TMO would hinder continuous and reversible divalent ion intercalation.9 Amorphous cathodes may enhance reversibility by providing a network capable of absorbing considerable deformations caused by magnesium insertion. In addition, the energy density of the cathode may be vastly improved if a TMO capable of multiple electron transfer is incorporated into the amorphous matrix. Divanadium pentoxide (V2O5) is touted as a potential high energy density cathode for battery applications. The capability for vanadium to access multiple oxidation states (V2+ ↔ V5+) in addition to a high theoretical capacity and voltage, could also provide benefits for polyvalent ion insertion.

For example, V2O5·nH2O xerogels, nano-crystalline V2O5, and sputtered V2O5 thin-films have shown promising results as Mg battery cathodes,11,16,17 however, the capability of Mg2+ to insert into an amorphous form of V2O5 is unexplored. A seminal result was first reported by Novàk et al.18 and confirmed by Lee et al.,19 concerning the improvement into layered cathodes, such as V2O5, by shielding of the Mg2+ with H2O. However, magnesium batteries containing an aqueous electrolyte media risks decomposition of the Mg metal anode surface. Therefore, our work here will concentrate on the modification of cathodes to improve the electrochemical properties in non-aqueous electrolytes. Previously, the amorphization of V2O5 through high-temperature melt-quenching with P2O5 has been studied as cathodes for Li-ion batteries,20 and more recently as a ternary Li2O–V2O5–P2O5 complex.21 Through a thorough analytical investigation, we identify the key elements to form the active component in the amorphous cathode.

The amorphous cathodes were synthesized through planetary ball-milling of V2O5 and P2O5 powders for 20 h. The resulting dark-to-light brown powders were analyzed with X-ray diffraction (XRD – Fig. 1a). The ratios of V2O5[thin space (1/6-em)]:[thin space (1/6-em)]P2O5 (mol[thin space (1/6-em)]:[thin space (1/6-em)]mol) 85[thin space (1/6-em)]:[thin space (1/6-em)]15 to 60[thin space (1/6-em)]:[thin space (1/6-em)]40 did not display the presence of any crystalline reflections, indicating the as-made powders are amorphous to XRD. In the high resolution transmission electron microscopy (HR-TEM) image of the 75[thin space (1/6-em)]:[thin space (1/6-em)]25 ratio, we did not observe ordering of the lattice (Fig. 1b) and we found similar results for all ratios, supporting the above XRD results. Instead, a halo pattern was observed for all ratios in the selected area electron diffraction. This synthetic method provides a scalable, safe route to form amorphous powders, however it must be noted that complete vitrification of the TMO is best achieved through high temperature melt-quenching.


image file: c5cc07161e-f1.tif
Fig. 1 (a) XRD patterns of representative V2O5[thin space (1/6-em)]:[thin space (1/6-em)]P2O5 powders compared to the commercial V2O5 powder. (b) TEM image and SAED pattern (inset) of a V2O5[thin space (1/6-em)]:[thin space (1/6-em)]P2O5 (75[thin space (1/6-em)]:[thin space (1/6-em)]25) powder. (c) CVs of a SiO2, a V2O5 powder cathode, and a V2O5[thin space (1/6-em)]:[thin space (1/6-em)]P2O5 (75[thin space (1/6-em)]:[thin space (1/6-em)]25) amorphous cathode. Comparison of the (d) current densities (mA g−1) and (e) 1st cycle galvanic curves of V2O5[thin space (1/6-em)]:[thin space (1/6-em)]P2O5 powders.

The new structure of V2O5[thin space (1/6-em)]:[thin space (1/6-em)]P2O5 powders exhibited significant improvements in the electrochemical properties when compared to the parent orthorhombic-V2O5. Zhou et al. has shown that the insertion of Mg2+ into crystalline V2O5 is accompanied with a large diffusion barrier,22 which indicates that manipulation of the structure is essential for V2O5-based cathodes. Fig. 1c is cyclic voltammograms (CVs) comparing a milled V2O5 cathode and a V2O5[thin space (1/6-em)]:[thin space (1/6-em)] P2O5 (75[thin space (1/6-em)]:[thin space (1/6-em)]25) cathode. A CV of electrochemically inert SiO2 is shown to confirm the width of the electrochemical window of the electrolyte and the electrochemical inertness of the conductive carbon and binder. At this point, we cannot completely discount the kinetic influence of the amorphous materials for decreasing the anodic stability of the electrolyte. However, with thorough XPS analysis, we have not been able to observe the decomposition of the ClO4 ion on the cathode surface. The CV of the bulk, crystalline V2O5 exhibits small cathodic current at low potentials (<1.1 V vs. Mg) and two anodic peaks at 2.6 V and 3.1 V vs. Mg. The addition of 25 mol% P2O5 revealed two notable electrochemical improvements: first, the current density (mA g−1) is significantly increased through the addition of P2O5, and second, the presence of an anodic peak at 2.9 V vs. Mg demonstrates improved reversibility in the high voltage region. We assign the cathodic current beginning at 1.1 V vs. Mg to a reduction of V2O5 observed in both the crystalline and amorphous form. The high-voltage reduction/oxidation couple precludes the accurate calculation of the charge through integration of the anodic/cathodic waves. Thus, the electrochemical properties of the cathodes were compared through calculation of the current density (mA g−1) of the oxidation peak potential. Fig. 1d displays the current density of the electrodes as a function of V2O5[thin space (1/6-em)]:[thin space (1/6-em)]P2O5 ratio. Interestingly, the current density of the cathodes finds a maximum value at 75[thin space (1/6-em)]:[thin space (1/6-em)]25 signifying that the electrochemical properties are not solely due to the quantity of P2O5. Galvanic discharge/charge curves taken at 5 mA g−1 rate of the first cycle are shown in Fig. 1e. The improved capacity within the high-voltage region is encouraging, and the initial capacity of 121 mA h g−1 with 57 mA h g−1 capacity retention up to 5 cycles for the 75[thin space (1/6-em)]:[thin space (1/6-em)]25 ratio (Fig. S1a, ESI). However, extended cycling is hindered by passivation of Mg metal in Mg(ClO4)2/CH3CN electrolytes.17 Since this is the first report on the magnesiation of amorphous V2O5–P2O5, we can only compare the initial discharge capacity to sputtered-V2O5 (180 mA h g−1),11 nanocrystalline V2O5 (180 mA h g−1),17 and the lithiation V2O5–P2O5 (>200 mA h g−1).20 We are encouraged that through the appropriate nanoscaling of our amorphous materials, we may obtain higher discharge capacities. Through V 2p and Mg 2s XPS (Fig. S1b, ESI), we have found that magnesium is indeed introduced into the amorphous matrix. The reduction of vanadium is evidenced through the appearance of a shoulder at 514.2 eV of the V 2p3 peak coinciding with appearance of a Mg 2s signal at the end of the discharge. The process is partially reversible after the first charge of the cathode. A deep investigation of the reaction mechanism of the amorphous material is required; however, the focus of this work is a thorough analysis of the structure of the amorphous cathode and the implications towards improved electrochemical activity.

We undertook a variety of analytical techniques to isolate and understand the role of vanadium and phosphorous in the cathodes. Fig. 2a is the V L-edge Near-Edge X-ray Absorption Fine Structure (NEXAFS) comparing the electronic structure of the V 3d levels in V2O5 and V2O5[thin space (1/6-em)]:[thin space (1/6-em)]P2O5 amorphous cathodes. Milled V2O5 exhibits three peaks (V1–V3) in the L3 region (2p3/2 → 3d) between 514–520 eV in addition to a broad peak in the L2 region (2p1/2 → 3d) between 522–528 eV. In accordance with the NEXAFS and XRD pattern (Fig. S2, ESI), the milling of V2O5 did not alter the structure of the material, but only served to reduce the grain size. On the other hand, with the addition of the P2O5, the appearance of a new peak (V4) in addition to the characteristic peaks for V2O5 is observed. The presence of V4 resonates well with the V 2p X-ray photoelectron spectroscopy (XPS) comparing the powdered, milled and V2O5[thin space (1/6-em)]:[thin space (1/6-em)]P2O5 (75[thin space (1/6-em)]:[thin space (1/6-em)]25) electrodes (Fig. S3, ESI). The addition of the network-former serves to reduce the vanadium to a mix of V4+/5+ oxidation state, which coincides with the loss of oxygen we observe with energy dispersive X-ray spectroscopy (Fig. S4, ESI). Previous researchers have hypothesized that amorphization results in re-orientation of the V2O5 sheets to accommodate the inclusion of PO4 units.23 Additionally, calculations show that atomic vacancies are most favorably created at the terminally-bonded, apexial oxygen,24 and are likely replaced by the PO4 units. As the percentage of P2O5 (and presumably the PO4) increases up to 75[thin space (1/6-em)]:[thin space (1/6-em)]25 mol%, there is a convergence of the V L3 peaks. Our hypothesis is the loss of peak structure is reflected by the higher quantity of V4+ sites and PO4 units resulting in improved vitrification. However beyond a 25 mol% of P2O5, there is a divergence of the L3 region into separate peaks, and the presence of a new peak, V5, at 514.5 eV. It is clear that the changes to the parent V2O5 are linked to the role(s) of P2O5.


image file: c5cc07161e-f2.tif
Fig. 2 (a) V L-edge NEXAFS spectra of V2O5[thin space (1/6-em)]:[thin space (1/6-em)]P2O5 powders and milled V2O5. (b) 31P MAS NMR of V2O5[thin space (1/6-em)]:[thin space (1/6-em)]P2O5 powders.

To observe the changes in the environment around the phosphorous, we performed 31P MAS NMR (Fig. 2b). From 15% ≤ P2O5 ≤ 25%, there is a single peak located at −12 ppm (Peak 1, the additional peaks are spinning side-bands). The position of the peak relative to the standard H3PO4 evidences the formation of the PO4 subunits during the mechanical milling synthesis, and the local environment around the phosphorous remains unchanged up to the 75[thin space (1/6-em)]:[thin space (1/6-em)]25 ratio in V2O5[thin space (1/6-em)]:[thin space (1/6-em)]P2O5. Upon further addition of P2O5 up to 40 mol%, a shoulder at −5 ppm (Peak 2) reveals a separate environment of the phosphorous in the amorphous powder, which coincides with a small upfield shift for Peak 1. Interestingly, the current density and capacity of the cathodes increases to a maximum at 25% P2O5 and decreases thereafter, which implies that the addition of P2O5 beyond 30 mol% has a detrimental effect on the magnesiation of the cathode.

Crystallization of the amorphous powders provides crucial links to electrochemical performance.25Fig. 3a is the heating profile from differential scanning calorimetry (DSC) comparing V2O5 and the V2O5[thin space (1/6-em)]:[thin space (1/6-em)]P2O5 powders. We would like to note that we did not observe any additional heat flow in the reverse scans for the DSC, and the structural changes due to heating are irreversible. A complete schedule for DSC measurements can be found in the ESI. All of the amorphous materials exhibit a glass-transition temperature (Tg), suggesting that the milling of the reactants generates a glassy state. The stability of the amorphous material is defined as the difference in temperatures (ΔT) between the 1st crystallization peak (Tc1) and Tg, ΔT = Tc1Tg. The value of Tg increases with P2O5 content, and at the same time, the value of ΔT tends to increase. Since P2O5 is a network former, it is reasonable that high content of P2O5 can easily stabilize a glassy state. For 75 ≤ V2O5 ≤ 85 mol%, we observe the presence of two exothermic peaks. The presence of two crystallization (Tc1 and Tc2) events after the Tg implies that metastable crystallites can be accessed from the pristine amorphous material by annealing at the first crystallization temperature, Tc1. On the other hand, for 60 ≤ V2O5 ≤ 70 mol%, the calorimetry displays only the presence of one peak after the glass transition temperature. The values of Tg, Tc1, Tc2 and ΔT are listed in Table S1 of the ESI. The calorimetry suggests that the first crystallization, Tc1, differs with P2O5 content, suggesting that a more metastable phase can be thermally isolated. Representative SEM images of the annealed powders are shown in Fig. S5a–d (ESI) for 75[thin space (1/6-em)]:[thin space (1/6-em)]25 and 60[thin space (1/6-em)]:[thin space (1/6-em)]40, respectively. For 75 ≤ V2O5 ≤ 85 mol%, annealing the powders at Tc1 triggers the formation of crater-like features due to the densification of the amorphous material and the subsequent crystallization at the base of the crater. When annealed at 450 °C, the formation of needle-like crystallites is observed in addition to the craters. Conversely, the 60 ≤ V2O5 ≤ 70 mol% V2O5 shows a reversed crystallization order. XRD (Fig. S6, ESI) shows how the crystallization of an apparent orthorhombic V2O5 phase coincides with the formation of the craters; while the needle-like structures are the formation of the various VOPO4 phases. Fig. 3b displays the XRD patterns of the (001) crystallographic plane for the reactant V2O5 powder and a 75[thin space (1/6-em)]:[thin space (1/6-em)]25 V2O5[thin space (1/6-em)]:[thin space (1/6-em)]P2O5 amorphous powder annealed at T = 313 °C (Tc1) and 450 °C. When precipitated from the amorphous powder at the first crystallization temperature, Tc1, the inter-layer spacing has increased as evidenced by the shift in peak position to 2θ = 20.10°.


image file: c5cc07161e-f3.tif
Fig. 3 (a) DSC scans of V2O5[thin space (1/6-em)]:[thin space (1/6-em)] P2O5 as-made powders and milled V2O5. (b) XRD patterns of 75[thin space (1/6-em)]:[thin space (1/6-em)]25 V2O5[thin space (1/6-em)]:[thin space (1/6-em)]P2O5 annealed at 313 °C and 450 °C. (c) Current densities (mA g−1) before and after annealing at Tc1 and 450 °C.

Crystallization has afforded insights into the active structural unit in of V2O5–P2O5 cathodes, however, precipitation from the amorphous phase does not yield a cathode with superior performance. Crystallization of amorphous Li2S[thin space (1/6-em)]:[thin space (1/6-em)]P2S5 solid-state ionic conductors helps the precipitation of superionic crystals, resulting in lowering of the activation energy and increasing of the conductivity for Li+ ion diffusion.26 The electrochemical results for the precipitated phases show that the formation of VOPO4, at 450 °C for P2O5 ≤ 25 mol% and at Tc1 for P2O5 ≥ 25 mol%, consistently reduces the electrochemical activity of the cathode (Fig. 3c). Therefore, the 75[thin space (1/6-em)]:[thin space (1/6-em)]25 ratio strikes a balance between obtaining the maximum amount of the active, metastable phase, without producing the inactive VOPO4 phase. The improved electrochemical activity due to the amorphization of the parent structure is further highlighted by the electrical conductivity of the 75[thin space (1/6-em)]:[thin space (1/6-em)]25 V2O5–P2O5 (1.7 × 10−8 S cm−1 @ 25 °C) in comparison to crystalline V2O5 (1.1 × 10−5 S cm−1 @ 25 °C). In accordance with previous studies, amorphous materials have much lower electronic conductivities than crystalline materials.27,28 Thus the improved electrochemical activity is a consequence of the increased interlayer spacing. Additionally, when a V–O–P bond is formed in place of an apexial oxygen, charge compensation dictates that there will be two V4+ for each PO4 unit as well as increasing the ionic character of the V–O–P bond. Scanlon et al.24 recently showed that the oxygen vacancy at the apexial position can displace the V4+ in the (001) direction towards the underlying layer and localize the charge on the surrounding vanadium atoms. The “clustering” of charges has been to shown to promote Mg2+ diffusion, as shown for Mo6 units in Chevrel Mo6S8 and recently C60.8,29

In conclusion, when compared to poly-crystalline V2O5, the amorphous V2O5[thin space (1/6-em)]:[thin space (1/6-em)]P2O5 cathodes show improved high-voltage electrochemical magnesiation. However, the reversibility, stability and battery voltage must still be explored through full-battery cell testing, therefore non-corrosive, high-voltage electrolytes compatible with magnesium metal remains an essential piece of the puzzle. Importantly, the results provide an exciting path to achieve high energy magnesium batteries.

Notes and references

  1. D. Aurbach, Z. Lu, A. Schechter, Y. Gofer, H. Gizbar, R. Turgeman, Y. Cohen, M. Moshkovich and E. Levi, Nature, 2000, 407, 724 CrossRef CAS PubMed.
  2. T. D. Gregory, R. J. Hoffman and R. C. Winterton, J. Electrochem. Soc., 1990, 137, 775 CrossRef.
  3. D. Aurbach, I. Weissman, Y. Gofer and E. Levi, Chem. Rec., 2003, 3, 61 CrossRef CAS PubMed.
  4. J. Muldoon, C. B. Buccur and T. Gregory, Chem. Rev., 2014, 114, 1683 CrossRef PubMed.
  5. R. E. Doe, R. Han, J. Hwang, A. J. Gmitter, I. Shterenberg, H. D. Yoo, N. Pour and D. Aurbach, Chem. Commun., 2014, 50, 243 RSC.
  6. T. J. Carter, R. Mohtadi, T. S. Arthur, F. Mizuno, R. Zhang, S. Shirai and J. W. Kampf, Angew. Chem., Int. Ed., 2014, 53, 1 CrossRef PubMed.
  7. Y. Cheng, T. Liu, Y. Shao, M. H. Engelhard, J. Liu and G. Li, J. Mater. Chem. A, 2014, 2, 2473 CAS.
  8. E. Levi, Y. Gofer and D. Aurbach, Chem. Mater., 2010, 22, 860 CrossRef CAS.
  9. C. Ling and F. Mizuno, Chem. Mater., 2013, 25, 3062 CrossRef CAS.
  10. Y. Cheng, L. R. Parent, Y. Shao, C. Wang, V. L. Sprenkle, G. Li and J. Liu, Chem. Mater., 2014, 26, 4904 CrossRef CAS.
  11. G. Gershinsky, H. D. Yoo, Y. Gofer and D. Aurbach, Langmuir, 2013, 29, 10964 CrossRef CAS PubMed.
  12. Y. Liang, R. Feng, S. Yang, H. Ma, J. Liang and J. Chen, Adv. Mater., 2011, 23, 640 CrossRef CAS PubMed.
  13. Y. NuLi, J. Yang, J. Wang and Y. Li, J. Phys. Chem. C, 2009, 113, 12594 CAS.
  14. Y. J. Kim, W. Wu, S.-E. Chun, J. F. Whitcare and C. J. Bettinger, Adv. Mater., 2014, 26, 6572 CrossRef CAS PubMed.
  15. T. A. Arthur, R. Zhang, C. Ling, P.-A. Glans, X. Fan, J. Guo and F. Mizuno, ACS Appl. Mater. Interfaces, 2014, 6, 7004 CAS.
  16. V. Petkov, P. Y. Zavalij, S. Lutta, M. S. Whittingham, V. Parvanov and S. Shastri, Phys. Rev. B: Condens. Matter Mater. Phys., 2004, 69, 85410 CrossRef.
  17. G. G. Amatucci, F. Badway, A. Singhal, B. Beaudoin, G. Skandan, T. Bowmer, I. Plitza, N. Pereira, T. Chapman and R. Jaworski, J. Electrochem. Soc., 2001, 148, A940 CrossRef CAS PubMed.
  18. P. Novàk and J. Desilvestro, J. Electrochem. Soc., 1993, 140, 140 CrossRef PubMed.
  19. J. Song, M. Noked, E. Gillette, J. Duay, G. Rubloff and S. B. Lee, Phys. Chem. Chem. Phys., 2015, 17, 5256 RSC.
  20. Y. Sakurai and J. Yamaki, J. Electrochem. Soc., 1985, 132, 512 CrossRef CAS PubMed.
  21. G. Delaizir, V. Seznec, P. Rozier, C. Surcin, P. Salles and M. Dolle, Solid State Ionics, 2013, 237, 22 CrossRef CAS PubMed.
  22. B. Zhou, H. Shi, R. Cao, X. Zhang and Z. Jiang, Phys. Chem. Chem. Phys., 2014, 16, 18578 RSC.
  23. F. R. Landsberger and P. J. Bray, J. Chem. Phys., 1970, 53, 2757 CrossRef CAS PubMed.
  24. D. O. Scanlon, A. Walsh, B. J. Morgan and G. W. Watson, J. Phys. Chem. C, 2008, 112, 9903 CAS.
  25. C. B. Carter and M. G. Norton, Ceramic Materials and Engineering, Springer, New York, NY, 2007 Search PubMed.
  26. F. Mizuno, A. Hayashi, K. Tadanaga and M. Tatsumisago, Adv. Mater., 2005, 17, 918 CrossRef CAS PubMed.
  27. F. Coustier, J. Hill, B. B. Owens, S. Passerini and W. H. Smyrl, J. Electrochem. Soc., 1999, 146, 1355 CrossRef CAS PubMed.
  28. L. Murawski, C. Gledel, C. Sanchez, J. Livage and J. P. Audieres, J. Non-Cryst. Solids, 1987, 89, 98 CrossRef CAS.
  29. R. Zhang, C. Ling and F. Mizuno, Chem. Commun., 2015, 51, 1108 RSC.

Footnote

Electronic supplementary information (ESI) available: Experimental details and additional results. See DOI: 10.1039/c5cc07161e

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.