Elsevier

Ultramicroscopy

Volume 230, November 2021, 113369
Ultramicroscopy

In-situ characterization of discharge products of lithium-oxygen battery using Flow Electrochemical Atomic Force Microscopy

https://doi.org/10.1016/j.ultramic.2021.113369Get rights and content

Highlights

  • A novel Flow Electrochemical Atomic Force Microscopy (FE-AFM) technique is described.

  • The technique was tested for studying the discharge products of a Lithium-Oxygen Battery (LOB).

  • The FE-AFM cell allows us to visualize the discharge of thin Li2O2 deposits on HOPG.

  • The formation of larger nanoparticles, with elongated and spherical globular shape, is attributed to the solution mechanism of Li2O2 formation.

  • The oxidation of the nanoparticles, follow a bottom-up oxidation mechanism, characterized by a significant reduction in the height of the particles.

Abstract

The increasing interest in lithium-oxygen batteries (LOB), having the highest theoretical energy densities among the advanced lithium batteries, has triggered the search for in-situ characterization techniques, including Electrochemical Atomic Force Microscopy (EC-AFM). In this work we addressed the characterization of the formation and decomposition of lithium peroxide (Li2O2) on a carbon cathode using a modified AFM technique, called Flow Electrochemical Atomic Force Microscopy (FE-AFM), where an oxygen-saturated solution of the non-aqueous lithium electrolyte is circulated through a liquid AFM cell. This novel technique does not require keeping the AFM equipment inside a glove-box, and it allows performing a number of experiments using the same substrate with different electrolytes without disassembling the cell. We study the morphology of Li2O2 on graphite carbon using lithium bis(trifluoromethane sulfonyl)imide (LiTFSI) in dimethyl sulphoxide (DMSO) as electrolyte under different operational conditions, in order to compare our results with those reported using other electrolytes and in-situ and ex-situ EC-AFM.

Introduction

Since it was described for the first time by Abraham and Jiang in 1996 [1], the Li-O2 battery (LOB) has attracted enormous research attention, aiming to approach its outstanding theoretical energy density [2], close to 3500 Wh kg−1, needed to improve the autonomy of electrical vehicles beyond the limit imposed by the Li-ion batteries. It is now well-known that there are important technical challenges for developing any practical LOB, namely, high charge potential, low discharge capacities, and poor electrolyte and cathode stability that strongly limit discharge−charge cycle efficiency and power density [3], [4], [5], [6].

The mechanisms of the oxygen reduction reaction (ORR) and the oxygen evolution reaction (OER) in a LOB operating with aprotic solvents have been clarified by several authors [7], [8], [9], [10], [11]. Thus, the ORR discharge reaction leading to the formation of Li2O2 on the cathode can occurs through surface or solution routes where, in both cases, the first step is the reduction of O2 to the superoxide radical (O2−•). The solution mechanism takes place in solvents with high donor number, where the superoxide ion forms ion pairs and higher aggregates with Li+ ions in the solution, which diffuse before chemical disproportion into Li2O2, leading to the formation of a thick layer. Solvents with low donor number promote the surface mechanism, leading to thin films (< 10 nm in thickness) due to the insulating nature of the deposit that self-limits its growth.

The dissociation degree of the lithium salt also seems to affect the morphology of the Li2O2 deposits [12]. Thus, a “bottom-up” Li2O2 precipitation mechanism operates with highly dissociated Li salts, such as LiTFSI in low donor number (DN) glymes, forming a thin layer of the product that cover the surface till the electrons can not penetrate further and the discharge process stops. On the contrary, for highly associated salts, such as LiNO3 in diglyme, a “top-down” precipitation mechanism that expands in all directions takes place, leading to large deposits. However, since the degree of the lithium salt association depends on a balance between the interaction of the Li+ ions with the solvent and the counteranion [13], it is not straightforward to predict the type of deposit that can be formed in a LOB under a given choice of salt and solvent.

For this reason, the direct observation of morphology of the cathodic product during the discharge, and its subsequent dissolution (oxygen evolution) during the charge cycle, is of crucial interest for the optimization of a LOB in order to improve the electrolyte performance. The use of Atomic Force Microscopy (AFM) is well known in the field of lithium batteries as a direct method to visualize the formation of the solid-electrolyte interphase (SEI) and the dendritric formation in lithium-ion batteries (LIB) [14,15]. A recent review by Zhao et al. [16] briefly mentioned a few AFM studies describing the use of AFM for the characterization of LOBs. Thus, it is worth to review here the studies reported during the last years, which analyze the morphology and conductivity of the discharge products of LOB using AFM and Electrochemical Atomic Force Microscopy (EC-AFM), performed in-situ [17], [18], [19], [20], [21], [22], and ex-situ [23], [24], [25].

Byon and coworkers carried out the first in-situ study of the formation[17] and decomposition[18] of Li2O2 on HOPG (highly oriented pyrolytic graphite) using the EC-AFM technique, with 0.5 M LiTFSI in TEGDME as electrolyte and metallic Li as counter-electrode. The AFM was located inside a glove-box and the contact mode AFM was realized with silicon nitride tips for the electrochemical scan and Si tips to observe the HOPG surface in tapping mode, once the cell was disassembled. The Li2O2 deposit on HOPG terrace edges exhibited nanoplate shape having several hundred of nm length and 5 nm height, and they disappear during the oxygen evolution reaction (OER). A similar study on a mesoporous gold electrode using the same electrolyte [19], allowed to visualize the formation of a film of Li2O2 nanoparticles with sizes around 10 nm, which grows up as the discharge proceeds, reaching a final size of 50 nm.

Liu and Ye [20] used in-situ EC-AFM to determine the morphology of the deposits on gold electrodes with 0.1 M LiClO4 in DMSO as electrolyte. They found that the shape of the Li2O2 nanoparticles depend on the water content of the electrolyte, and concluded that there is no evidence of a disproportion of LiO2 in solution for the formation of Li2O2 on the surface. The voltagram shows two cathodic peaks which were attributed to the formation of the superoxide ion followed by the formation of the peroxide. At a cathodic potential close to the first peak small particles were observed on the surface, which grow up when the second peak is reached. During the anodic scan, in an electrolyte containing 1000 ppm of H2O, three peaks of low intensity were observed, indicating that the species formed during the ORR are not fully oxidized during the OER. The two first anodic peaks were assigned to partial oxidation of the peroxide and superoxide (in solution) on the surface, but the process that originates the third peaks was not identified. If the electrolyte contains only 33 ppm of H2O, the cathodic voltagram only shows one peak, and the Li2O2 particles are smaller as the water content diminishes.

Shen et al. [21] studied by EC-AFM the ORR/OER processes on HOPG using 0.5 M LiTFSI in DMSO with the addition of DBBQ (2,5-di-tert-butyl-1,4-benzoquinone) as soluble catalyst. The voltagrams show two cathodic peaks, attributed to the superoxide and peroxide formation, and three anodic peaks, the first one assigned to the superoxide oxidation and the other two to the peroxide oxidation. The AFM images in absence of DBBQ show the initial formation of Li2O2 of size around 5 nm when the first peak is approached, while a toroidal morphology around 300–400 nm in diameter is formed when the second peak is reached. During the charge process a bottom-up mechanism of toroid decomposition is observed. In the presence of DBBQ the Li2O2 deposit adopts a flower-like shape that decomposes from the outside to the inside during the OER.

An AFM cell that allows the control of the atmosphere over the glassy carbon substrate was built by Virwani et al. [22]. They used 1 M LiNO31 M in TEGDME as electrolyte, with different water contents, and analyzed the topography of the deposits during the ORR and OER. A remarkable effect of the water content was observed on the deposit morphology and the discharge capacity of the cell. For instance, the cell capacity increases 11 times when the water content increase from 20 to 4600 ppm. The Li2O2 deposits formed with small water contents exhibit spherical or bar shape in the range between 200 and 400 nm, and they appear at the final of the discharge. With high water content, large toroidal particles (∼ 400 nm) are observed at the beginning of the discharge, while smaller ones are formed later.

Ex-situ AFM studies were performed by Calvo and coworkers using 0.1 M LiPF6 in DMSO on HOPG [23,24]. In this case the cathodic voltagram has only one peak, while the anodic peaks were assigned to the oxidation of the superoxide, the solvent, and other products of reduction of oxygen. The particles deposited on the HOPG edges and terraces never exceed 12 nm. The few particles deposited at low polarization are supposed to form by disproportion of the LiO2, while the massive deposit at higher cathodic potentials would be due to the superoxide reduction over the surface.

An ex-situ study on HOPG aimed to analyze the effect of the incorporation of chloride on the capacity of LOB [25], and it is the only AFM work where the electrical conductivity of the formed film was analyzed. The authors claim that an important increase in the film conductivity was detected by Conductive-AFM, when Cl ions are added to the electrolyte. However, some doubts on the conclusions arise due to the lack of care when the film deposited on the electrode is transferred to the AFM cell. It is probable that the Li2O2 was converted to LiOH before the conductivity was measured with the AFM and also, that the high conductivity reported be a consequence of the contact of the Pt AFM tip with the HOPG surface in some places. In fact, DFT studies on Li2O2 structures doped with Cl ions have probed that the conductivity of the systems is not altered by the presence of chloride or other halogen anions [26].

In this work we have developed a novel procedure for in-situ observation of the deposit in a LOB during discharge/charge cycles, which can be performed without resorting to the use of a glove-box, making the technique more simple that previously reported in the literature. The technique was used to analyze the morphology of the Li2O2 deposit on highly oriented pyrolytic graphite (HOPG), using lithium bis(trifluoromethane sulfonyl)imide (LiTFSI) in dimethyl sulphoxide (DMSO) as electrolyte, under different operational conditions.

Section snippets

Materials

The electrolyte was prepared using anhydrous dimethyl sulfoxide (DMSO, > 99.9%, Sigma-Aldrich) and lithium bis(trifluoromethane sulfonyl)imide (LiTFSI, Sigma-Aldrich, 99.95%). The DMSO, dried during several days on 3 Å molecular sieves (Sigma-Aldrich), had a water content < 50 ppm, measured by Karl Fisher titration. The solutions of LiTFSI in DMSO were prepared in a MBRAUN glove-box (H2O < 3 ppm and O2 < 2 ppm), and stored in 10 cm3 vessels sealed with rubber septum. Once removed from the

Results and discussion

Fig. 3 shows the CV of the cell in 0.5 M LiTFSI/DMSO saturated with O2 and Ar. The scan rate was 50 mV/s from the OCP (- 3 V) up to 2.0 V for the cathodic scan and from 2.0 V up to 4.8 V for the anodic scan. The CV for the O2 reduction shows a two cathodic peaks at 2.80 V and 2.57 V, attributed to the reduction of O2 to LiO2 and Li2O2, respectively [4,5,12].

The anodic branch of the CV exhibits three peaks at 3.64, 4.19, and 4.53 V of lower intensity, indicating that the species formed during

Conclusions

In summary, we developed an experimental technique, called Flow Electrochemical Atomic Force Microscopy, which allows the in-situ characterization of Li2O2 morphology during its formation and electrochemical oxidation. This experimental setup does not require placing the AFM inside a glove box, representing an advantage regarding the EC-AFM technique used in the previous works aforementioned. The system was tested with a model LOB in which a HOPG was used as cathode and a solution 0.5 M LiTFSI

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Acknowledgment

This work was supported by CONICET (PIP 112 201301 00808). HRC is a research fellow of CONICET. HAC thank doctoral fellowships by CONICET.

References (26)

  • K.M. Abraham et al.

    Polymer electrolyte-based rechargeable lithium/oxygen battery

    J. Electrochem. Soc.

    (1996)
  • X. Zhang et al.

    Recent progress in rechargeable alkali metal–air batteries

    Green Energy Environ.

    (2016)
  • G. Girishkumar et al.

    Lithium-air battery: promise and challenges

    J. Phys. Chem. Lett.

    (2010)
  • P.G. Bruce et al.

    Li−O2 and Li-S batteries with high energy storage

    Nat. Mater.

    (2012)
  • A.C. Luntz et al.

    Nonaqueous Li–air batteries: a status report

    Chem. Rev.

    (2014)
  • P.A.N. Imanishi et al.

    The Lithium Air Battery

    (2014)
  • C.O. Laoire et al.

    Elucidating the mechanism of oxygen reduction for lithium-air battery applications

    J. Phys. Chem. C

    (2009)
  • D. Aurbach et al.

    Advances in understanding mechanisms underpinning lithium-air batteries

    Nat. Energy

    (2016)
  • L. Johnson et al.

    The role of LiO2 solubility in O2 reduction in aprotic solvents and its consequences for Li−O2 batteries

    Nat. Chem.

    (2014)
  • B.D. McCloskey et al.

    Mechanistic insights for the development of Li–O2 battery materials: addressing Li2O2 conductivity limitations and electrolyte and cathode instabilities

    Chem. Commun.

    (2015)
  • Y. Wang et al.

    A solvent-controlled oxidation mechanism of Li2O2 in lithium-oxygen batteries

    Joule

    (2018)
  • D. Sharon et al.

    Mechanistic role of Li+ dissociation level in aprotic Li-O2 battery

    ACS Appl. Mater. Interfaces

    (2016)
  • G. Horwitz et al.

    Ionic transport and speciation of lithium salts in glymes: experimental and theoretical results for electrolytes of interest for lithium−air batteries

    ACS Omega

    (2018)
  • Cited by (4)

    • Effects of sodium salts on compatibility between Na<inf>2</inf>Ti<inf>3</inf>O<inf>7</inf>@C anode and electrolyte for sodium-ion batteries

      2023, Journal of Alloys and Compounds
      Citation Excerpt :

      In recent years, new energy secondary batteries, such as lithium-oxygen battery [1–4], lithium-sulfur battery [5–8], lithium-ion battery [9–15], have already been brought into public concern due to their advantages such as high energy density, high operating voltage, no memory effect and long life.

    • Metal-air batteries: progress and perspective

      2022, Science Bulletin
      Citation Excerpt :

      Limits of in-situ characterizations. Massive in-situ characterization such as SERS, atomic force microscope (AFM), TEM, scanning electron microscope (SEM), Fourier transform infrared spectroscopy (FTIR), nuclear magnetic resonance (NMR), differential electrochemical mass spectrometry (DEMS), etc., has been carried out to unveil the reaction mechanisms and identify the active sites of catalysis [11,18,57–63]. However, the discharge mechanism is still not fully understood yet.

    • A novel efficient electrocatalyst for oxygen reduction and oxygen evolution reaction in Li-O<inf>2</inf> batteries: Co/CoSe embedded N, Se co-doped carbon

      2022, Applied Catalysis B: Environmental
      Citation Excerpt :

      From Fig. 6c, it is clear that the Co-CoSe@NSeC/bioC shows higher ORR and OER peak currents compared to Se-bioC, implying that Co-CoSe@NSeC/bioC can generate more Li2O2 during the ORR process and more discharge products are decomposed in the OER process. During the positive scanning, two cathodic (reduction) peaks at 2.8 V and 2.2 V are clearly observed and correspond to the reduction of O2 into LiO2 and further to Li2O2, respectively [54]. Accordingly, there are two oxidation peaks at 3.31 V and 4.15 V.

    View full text