Next Article in Journal
Modifying Plasmonic-Field Enhancement and Resonance Characteristics of Spherical Nanoparticles on Metallic Film: Effects of Faceting Spherical Nanoparticle Morphology
Next Article in Special Issue
In Vivo Assessment of Synthetic and Biological-Derived Calcium Phosphate-Based Coatings Fabricated by Pulsed Laser Deposition: A Review
Previous Article in Journal
Improvement of the Sphericity and the Thickness Uniformity of the Polystyrene (PS) Shell Microsphere during Curing Process
Previous Article in Special Issue
Relevance of the Preparation of the Target for PLD on the Magnetic Properties of Films of Iron-Doped Indium Oxide
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Pulsed Laser Deposited Films for Microbatteries

Institut de Minéralogie, de Physique des Matériaux et de Cosmochimie (IMPMC), Campus Pierre et Marie Curie, Sorbonne Université, CNRS UMR 7590, 4 Place Jussieu, 75005 Paris, France
*
Author to whom correspondence should be addressed.
Submission received: 10 May 2019 / Revised: 6 June 2019 / Accepted: 10 June 2019 / Published: 14 June 2019
(This article belongs to the Special Issue Current Research in Pulsed Laser Deposition)

Abstract

:
This review article presents a survey of the literature on pulsed laser deposited thin film materials used in devices for energy storage and conversion, i.e., lithium microbatteries, supercapacitors, and electrochromic displays. Three classes of materials are considered: Positive electrode materials (cathodes), solid electrolytes, and negative electrode materials (anodes). The growth conditions and electrochemical properties are presented for each material and state-of-the-art of lithium microbatteries are also reported.

Graphical Abstract

1. Introduction

It has been widely demonstrated that pulsed-laser deposition (PLD) based on the process of the transportation of a material (laser ablation) is a successful technique for the growth of stoichiometric multicomponent oxide films [1,2]. Indeed, PLD has shown unique advantages for the formation of dense films for energy storage and conversion, namely a high reproducibility, easy control of the growth rate, and a high film purity with a variety of substrates, such as amorphous glass, oriented silicon [3], stainless steel [4], (001)Al2O3 [5], indium tin oxide (ITO)- and ZnO-coated glass, and ITO-coated Upilex polymer [6]. Generally, the stoichiometry of the target phase is preserved in PLD films of oxides but a deviation is observed for lithiated material that implies an Li-enriched target. Consequently, the loss of volatile Li during deposition is compensated for by using about a 15 wt.% excess of Li2O [7,8]. Accurate stoichiometry can be obtained by controlling several parameters of the process. The typical set-up for the fabrication of PLD films consists of a stainless-steel vacuum chamber evacuated down to a residual pressure less than 1 × 10−4 Pa before material deposition. Energy (laser fluence) in the range of 1.0 to 3.0 J·cm−2 is generated by a pulsed-laser beam, which falls onto the target surface with an incidence angle of approximately 45° (see Table 1 for laser characteristics). Indeed, four PLD parameters are of prime importance for the growth of films: The laser fluence; type of substrate; orientation and lattice parameters, which must match with those of the film for an efficient epitaxy process; substrate temperature (Ts); and the oxygen partial pressure (PO₂). In addition, as the capacity of the microbattery depends on the electrode thickness, the duration of the deposition (tp) must also be considered. The activity of a thin-film electrode, i.e., specific discharge energy, is proportional to the thickness, thus an increase of the film thickness leads to a power limitation because of the slow transport kinetic of Li+ ions. Consequently, PLD is a popular technique due to the growth of a compact and dense film, which is replaced by a thick and porous film. Another advantage of the physical vacuum-like deposition techniques is the possibility of depositing a thin layer on top of the microbattery, which protects the device against a reactivity toward moisture. Moreover, due to the well-defined surface area of PLD films, a direct comparison of the electrochemical activity of materials can be done for different morphologies, from amorphous to single-crystalline materials [9].
Due to their high energy and power densities, lithium-ion batteries (LIBs) are initial power sources that are widely used in portable devices (laptops, mobile phones, cameras, etc.) and are now employed for sustainable transportation, such as full electric vehicles (EVs) and hybrid electric vehicles (HEVs). The fabrication of electrochemical cells with a thin-film architecture allows the development of microbatteries for powering micro-scaling devices, such as stand-alone sensor systems, medical implants and devices, labs-on-chip, credit cards, etc. In addition to technological applications, positive (cathode) and negative (anode) electrodes in the thin-film form are useful for studying the intrinsic properties of the material without the use of a polymeric binder and carbonaceous additive [15]. The use of thin-film technology may offer various advantages, such as: (i) Thin films are well suited for the design of devices; (ii) thinning of the layers provides a lower resistance in the transverse direction for weakly semiconducting materials; (iii) a reduction of the thickness of the solid electrolyte film allows the use of glassy materials with a low ionic conductivity; (iv) a reduction of the charge-transfer resistance of the electrolyte–electrode interface; (v) easy manufacture of microbatteries using the same technique that is currently used in the microelectronics industry; and (vi) the construction of microbatteries is realized in almost any two-dimensional shape. However, the fabrication of microbatteries also contains many difficulties, which are comprehensively discussed below [16].
In the present review paper, we present the properties of pulsed-laser deposited films used as components of energy storage devices (i.e., batteries, supercapacitors, electrochromics, etc.). The remainder of the article is organized as follows. First, the state-of-the-art of lithium microbatteries using PLD films are summarized in Section 2, providing the characteristics of the best lithium microbatteries fabricated so far. In Section 3, Section 4 and Section 5, the three classes of active materials constituting electrochemical microdevices, realized via the PLD technique, are considered: (i) Positive electrode materials (cathodes), (ii) electrolytes, and (iii) negative electrode materials (anodes). For each material, the growth conditions and electrochemical properties are presented. Finally, in Section 6, we compare and discuss the growth conditions that allow the best electrochemical performance of each electrochemically active component of microbatteries.

2. Lithium Microbatteries

The concept of a thin-film solid-state battery is quite old [17]. The subject of thin-film microbatteries has been discussed in the scientific literature for many years. The review by Kennedy is a good source for work prior to 1977 [18]. The concept and design of all-solid-state planar thin-film microbatteries have been patented by Bates et al. [19,20,21,22,23,24], who reported on micropower sources using lithium phosphate, lithium phosphorus oxynitride, and lithium phosphorus lithium oxide as a solid thin-film electrolyte. Julien investigated the electrochemical performance of individual layers in a microbattery in relation to the growth mechanism and thin-film structure [25]. Dudney addressed how to build a battery layer-by-layer by vapor deposition [26]. More recently, Oudenhoven et al. reviewed the concepts of three-dimensional (3D) microbatteries [27]. In 2015, Wang et al. discussed the choice of materials for lithium and lithium-ion microbatteries and reviewed the chemistry and electrochemistry for applications in microelectronic devices [28]. Ferrari et al. highlighted the importance of 3D microarchitecture electrodes to fabricate microgenerators for micro-electromechanical systems (MEMSs) [29]. In the presentation of in situ analytical microprobes, Meng et al. described PLD-produced thin-film lithium microbatteries using the PLD technique and showing the production of a multilayer structure with dense and smooth films [30].
There are many variations on the general scheme of microbatteries outlined in the literature. Two principal options are shown in Figure 1 [25]. Figure 1a shows a four-layer design on a conducting substrate (i.e., oriented silicon wafer) that can act as a current collector. Figure 1b shows a six-layer stack incorporating two metallic current collectors. There are two fast-ion conductor (FIC) layers in this design: A thick layer, which is the solid electrolyte itself, and a thin buffer film that acts as an electrolyte layer between the FIC and the Li metal film to prevent interface passivation. In 1992, a thin-film solid-state microbattery with an overall thickness of approximately 10 µm, including the TiS2 cathode, oxide-sulfide solid electrolyte, LiI buffer, and Li metal anode, was developed at the Technology Laboratory of Eveready Battery Company (EBC) [31]. Laïk et al. evaluated the performance of three 4-V commercial all-solid-state lithium microbatteries (200-µm thick) with a nominal capacity of 700 µAh based on a LiCoO2 cathode material [32]. Note that a typical 1-mWh battery weighs 2.5 mg and has a volume less than 1 µL, providing a specific energy and power of 400 Wh·kg−1 and 1 kWh·L−1, respectively [26].
Regarding the manufacture of thin-film batteries, several start-up companies have marketed micropower sources. Enfucell developed a thin, printable, and flexible SoftBattery® used in various wearable electronics products [33]. Cymbet Corporation fabricates the EnerChipTM battery, which is a battery bare die and can be embedded with other integrated circuits [34]. Excellatron announced a pilot production line (10,000 cells/month) of thin-film solid-state batteries (approximately 0.3 µm thick) made of cathode films of LiCoO2 or LiMn2O4, LiPON as the electrolyte, and Li metal or Sn3N4 as the anode based on the technology developed at Oak Ridge National Labs. Using a 2-µm thick positive electrode, these microbatteries have been cycled in excess of 2000 cycles [35]. Some industrial developments of thin film microbatteries are listed in Table 2.
A sequential PLD technique was applied for the fabrication of a rechargeable thin-film lithium battery (2-µm thick, area of 0.23 cm2) with partially crystallized LCO as the cathode, an Li6.1V0.61Si0.39O5.36 (LVSO) glassy electrolyte, and SnO film anode [41]. The ablation beam produced by a Q-switched Nd:YAG laser (λ = 266 nm, repetition rate of 10 Hz) was used at the fluence of 3.5 mJ·cm−2. A single phase LCO film was obtained by post annealing at 600 °C for 1 h in air and the amorphous LVSO film exhibited an ionic conductivity of ca. 10−7 S·cm−1 at room temperature. Such a Li microbattery cycled at 44 µA·cm−2 in the voltage range of 0.7 to 3.0 V delivered a capacity of 9.5 Ah·cm−2. After 100 cycles, the capacity retention was 45% of that of the first cycle. Characteristics of solid-state lithium microbatteries fabricated by the PLD technique reported in the literature are listed in Table 3. Most of the microcells use LiCoO2 as the positive electrode, providing a nominal voltage of ~3.9 V vs. Li+/Li. Thus, among the fabricated all-solid-state thin-film lithium batteries, the electrochemical chain of Li-In/80Li2S–20P2S5/LiCoO2 with an average potential of 3.5 V exhibits the best performance in terms of energy density.
Sakuda et al. reported the construction of an SSLMB based on an LCO positive electrode with an SE coating, a highly conductive 80Li2S–20P2S5 solid electrolyte, and an Li-In alloy as the anode [42]. Such a microbattery delivered a specific capacity of 95 mAh·g−1 at a current density of 0.13 mA·cm−2. By using LCO thin films prepared from a LiyCoOδ target containing 15% Li2O, Xia et al. fabricated thin-film microbatteries by the successive deposition of an LCO cathode on a Pt/Ti/SiO2 (amorphous)/Si composite substrate and amorphous Si anode [12]. Recently, the analysis by impedance spectroscopy of the microcell Li/LiPON/Li4Ti5O12 (nominal voltage of 1.5 V), in which LiPON is an amorphous lithium phosphorus oxynitride (i.e., nitrogen-modified Li3PO4), has clarified the debate on the interface stability with lithium; it was clearly shown that LiPON forms a well-conducting solid electrolyte interface (SEI) layer [43]. Despite the low ionic conductivity (1 µS·cm−1) and the rather large contribution to the internal cell resistance, LiPON can be used as a solid electrolyte film with a thickness of ~1 µm or less. Therefore, use of the LiPON-LiCoO2 system is very popular in the construction of thin-film lithium microbatteries [26].

3. Positive Electrode PLD Films

The main parameter for a microbattery is the delivered specific capacity. Rather than being expressed as the conventional unit of mAh·g−1, due to the uncertainty in the film density, technologists prefer the stored charge, Q (expressed in µAh or in coulomb), per film surface area and the film thickness, i.e., µAh·cm−2·µm−1 of mC·cm−2·µm−1.The relation between the gravimetric capacity, Qm, of the material and the volumetric capacity of a film, Qf, is given by:
Qf = 0.36 d Qm
where Qf is expressed in mC·cm−2·µm−1, Qm in mAh·g−1, and d is the density of the material in g·cm−3. Table 4 summarizes the energetic quantities for the studied cathode compounds.

3.1. LiCoO2 (LCO)

Having a lamellar structure, LiCoO2 (LCO) is the prototypal positive electrode material commonly used in Li-ion batteries that yields a practical specific capacity of 135 mAh·g−1 in the voltage range from ~3.8 V (fully lithiated state) to ~4.2 V (charge state at Li0.5CoO2) [46]. Since the early work in 1996 by Berkeley’s group [47], numerous studies have been devoted to the growth of LiCoO2 thin films prepared by the PLD technique due in large to their high electrochemical performances. Further investigations of dense and well-defined PLD films described the phase evolution during Li extraction and the kinetics of Li+ ions in the host lattice, which eventually found applications in the fabrication of the cathode element in microbattery stacks [48]. The two crystal forms, HT- and LT-LiCoO2 phases, with the rock-salt (rhombohedral, R-3m space group) and spinel (cubic, Fd3m space group) structure, respectively, have been synthesized by pulsed-laser deposition. It was pointed out that the crystallographic texture for LCO films differs from one deposit technique to another, i.e., PLD versus sputtering, which influences the electrochemical properties due to the diffusion plane orientation [49]. Julien et al. stated that well-crystallized PLD-grown LCO thin films with a single layered structure can be obtained at substrate temperatures (Ts) as low as 300 °C [3].
The first growth of single phase LCO films by the PLD method was realized by Antaya et al. [4]. Films deposited on unheated stainless-steel substrates were amorphous but crystallized readily with heat treatment in air above 500 °C. Later, Striebel et al. [47] demonstrated the promise of PLD-grown films as cathodes for rechargeable lithium cells. Crystalline (003)-textured LCO films with thicknesses ranging from 0.2 to 1.5 µm were prepared without postdeposition treatment, which displayed a specific capacity of films of 62 µAh·cm2·µm−1 and an Li diffusion coefficient of 1 × 10−10 cm2·s−1. Highly dense LCO films were first elaborated by the PLD process using a KrF laser under oxygen flow rates of 30 sccm and the pressure was maintained at 260 Pa on (200)-textured F-doped SnO2 on fa used silica substrate maintained at Ts = 700 °C [49]. As-prepared LCO thin films were (00l) textured and had a density of 85% of the single crystal. The charge–discharge profile of the films was typical of the LCO bulk and presented an ~18% capacity loss for a single cycle to 4.15 V. In the potential range of 4.14 to 4.19 V, the measured chemical diffusion coefficients ranged from 1.7 × 10−12 to 2.6 × 10−9 cm2·s−1 for as-deposited films and films annealed at 700 °C, respectively. Structural analysis of nanostructured LCO films prepared with PLD has been conducted by several research groups. Julien et al. [3,8,50] analyzed changes of the stoichiometry (i.e., the absence of the Co3O4 amorphous phase) as a function of the growth conditions using Raman spectroscopy. The inclusion of Co3O4 impurity is detected by analysis of the Raman intensity of the A1g modes. Impurity-free films exhibit a specific capacity as high as 195 mC·cm−2·µm−1 for polycrystalline films grown from an Li-rich target (i.e., excess of 15% Li2O). The work by Okada et al. revealed that a decrease of the amount of inclusions can be obtained by a lower laser fluence and lower Ts [51]. Figure 2 presents the relationship between the impurity inclusions and growth conditions of PLD-grown LCO films established from spectroscopic Raman data. In this figure, the Co3O4 phase is grown under the conditions of high PO₂, i.e., above the gray dashed line.
Ohnishi et al. [52,53] stated that suppression of the Co3O4 spinel phase can be ensured by the growth under a relatively low oxygen partial pressure. Zhang et al. [54] discussed the effect of the deposition conditions on the structure and morphology. The advantages of the preferential orientation of LCO films prepared by PLD has been discussed by numerous groups with the conclusion that dense uniaxial textured (003)-oriented films are obtained by a well-chosen substrate [6,53,55,56,57,58,59]. However, Xia et al. [59] stated that the fast transport of Li+ ions is obtained for LCO films with a random orientation, in contrast with the results obtained with films having (003)-preferred orientation. Contrastingly, Nishio et al. claimed an excellent electrochemical performance for epitaxially grown LCO (77-nm thick) with a (104)-orientation on a (100) Nb:SrTiO3 substrate that exhibited a discharge capacity of 26 mAh·g−1 even at high rates up to 100C [60]. Huo et al. stated that the film orientation is strongly dependent on the thickness and size of grains and demonstrated that films structured with parallel (003) planes are grown for thicknesses up to 300 nm [11]. Liu et al. showed that under certain PLD conditions, such as a high repetition rate of 35 Hz and low oxygen partial pressure of PO₂ = 1 Pa, LCO films tend to grow LCO films with a random orientation [61].
Epitaxial LCO thin films deposited on (001)-Al2O3 substrates remained in a single phase in a narrow range, 250 ≤ Ts ≤ 300 °C, whereas secondary phases appeared at Ts > 300 °C, i.e., Co2O3, Co3O4, and LiCo2O4 [5]. Xia et al. established that thin LCO films can be easily grown with a (003) orientation because of the lowest surface energy for the (003) plane, while the minimized strain energy in thick LCO films allows preferential (101) and (104) textures. It seems that this last type of orientation favors the electrochemical performance of the LCO cathode [12]. The reduction of the laser fluence results in a decrease of the surface roughness of LCO films. With post annealing at 400 °C and optimized deposition conditions, LCO films exhibit an initial discharge capacity of 36 µAh·cm−2·µm−1 and a cycleability of 94% [57]. Composition control was monitored to prepare stoichiometric LCO films using an Li-enriched target with a high-rate growth via an increase of the laser fluence to 0.29 J·cm−2 and an adjustment of the PO₂ to scatter the excess lithium. Ohnishi et al. showed that by using a Li1.1CoO2+δ target, the deposition of stoichiometric LCO with the highest crystallinity can be realized at the rate of 0.06 Å per pulse at the PO₂ of 0.1 Pa and Ts = 800 °C (Figure 3) [52,62].
Recently, Nishio et al. claimed that a high deposition rate of 1.2 Å·s−1 tends to form oxygen-deficient LCO films due to the destabilization of Co3+ cations and showed that post-annealing in air cancels the impurity phase [63]. Funayama et al. studied the effects of mechanical stress applied to LCO films (200 nm thick) deposited on Li-glass ceramic by the PLD method at 600 °C under a 20 Pa oxygen partial pressure for 1 h. Due to the lattice volume change, the generated electromotive force was 6.1 × 10−12 V·Pa−1 [64]. The electrode behavior shows an increase of the discharge capacity from 10 to 40 mAh·g−1 at a 2C rate with an increase of the Ts from 600 to 750 °C, whereas a Ts = 800 °C worsens the performance. Studies of the physico-chemistry of PLD-grown LCO thin films report the structural, surface morphology, optical, and electrical properties [65,66,67].
The electrochemical properties, i.e., thermodynamics and kinetics, of lithium intercalation in PLD LCO thin films have been widely investigated by the structure–electrochemistry relationship [59,68,69,70]; structural evolution upon charge-discharge cycles [71]; effect of doping by Ti, Al, and Mg [49,72,73,74]; characterization of the electrode/electrolyte interface [75]; and lithium-ion kinetics vs. basal plane orientation [76,77,78,79,80]. Highly (003)-oriented impurity-free LCO thin films grown by PLD on a stainless-steel substrate display an initial discharge capacity of 52.5 µAh·cm−2·µm−1 and a capacity loss of 0.18% per cycle at a moderate current density of 12.7 µA·cm−2. These films show a very small lattice expansion upon charge, i.e., 0.09 Å for a charge of 4.2 V [81]. Figure 4 shows the typical electrochemical features of PLD LCO thin films grown on an Si wafer maintained at different temperatures. Both the specific discharge capacity and the mid-voltage increased with increasing Ts. For a film deposited at Ts = 300 °C under PO₂ = 15 Pa, the discharge capacity reached a value ~140 mC·cm−2·µm−1.
The electrochemical behavior of doped LCO thin films show that the voltage plateau at 3.65 V disappears in the charge curve of LiTi0.05Co0.95O2 due to the doping effect, which cancels the semiconductor-metal-like transition of the LCO framework [81]. The Al-doped LCO film (LiCo0.5Al0.5O2) exhibits a steady increase in the voltage vs. Li extraction with the absence of a voltage plateau as observed in stoichiometric LCO films; however, such films suffered from a limited capability and an upper bound of the diffusion coefficient of Li (D* = 9 × 10−13 cm2·s−1) was observed [49]. Recent studies report on the improved electrochemical behavior of surface-modified LCO films using lithium tantalate (LTaO) and lithium niobite (LNbO). Coating with LNbO preserves the LCO surface and decreases the interfacial resistance, which indicates fast lithium transport [82]. LCO films modified by amorphous tungsten oxide (LWO) fabricated by PLD show a high capacity retention of Qr = 80% at a high rate of 20C, against Qr = 0% for bare LCO films cycled at the same C-rate. A slight increase of the superficial diffusion coefficient of Li+ ions from 2.2 × 10−13 and 3.0 × 10−13 cm2·s−1 was also observed, owing to the surface modification [83,84,85,86]. Note that LWO as well as LNBO are lithium ion conductors, which act as an efficient buffer between the electrolyte and LCO cathode. The structural degradation of cycled LCO films was investigated by Raman spectroscopy over 400 cycles, showing microstructural modification due to nanocrystallization and phase separation [87].
All-solid-state lithium microbatteries (SSLMB) using LiCoO2 films have been developed using various inorganic solid electrolyte (SE) films, i.e., LiPON, Li2S–P2S5, and amorphous Li3PO4. The thin-film battery with an electrochemical chain Li/amorphous Li3PO4/LCO/Pt shows a columnar-like LCO cathode (see the cross-sectional SEM image in Figure 5a) [88]. The excellent electrochemical performance is displayed in Figure 5b. This microcell exhibited an increase in capacity of up to 240 µAh·cm−2 when increasing the LCO thickness to 6.7 µm, which is 54% of the theoretical specific capacity of LCO (69 µAh·cm−2·µm−1). Shiraki et al. fabricated an SSLMB with an epitaxial LCO thin-film cathode (200 nm thick) by using PLD with a polycrystalline Li1.1CoO2 target ablated at a laser fluence of 1 J·cm−2, LiPON solid electrolyte (2 µm thick), and Li film as the anode (0.5 µm thick) [89]. The authors reported cyclic voltammograms with six redox peaks, which drastically changed upon cycling but did not display the galvanostatic charge–discharge profile of the SSLMB.

3.2. LiNiO2 (LNO)

PLD-grown thin films of lithium nickel oxide (LNO), i.e., LixNi1−xO and stoichiometric LiNiO2, are applied as electrochromic and/or battery electrodes. In an early work, Rubin et al. established the complex relationship of the surface morphology and chemical composition of LixNi1−xO thin films vs. the deposition oxygen partial pressure, substrate temperature, and substrate–target distance as well [90]. LNO films produced at Ts < 600 °C immediately absorb CO2 and H2O when exposed to air, whereas they show long-term stability for Ts = 600 °C. LNO film with a composition of Li0.5Ni0.5O (cubic rock-salt Fd-3m structure instead of the rhombohedral R-3m structure for LiNiO2) was obtained under a deposition atmosphere of PO₂ = 60 mTorr. This film (150-nm thick) showed excellent electrochemical reversibility as an electrochromic item in the range of 1.0 to 3.4 V vs. Li+/Li. An electrochromic device using WO3 as the opposite electrode and PEO/LiTFSI as the solid polymer electrolyte (250-µm thick) showed an optical transmission range of ≈70% at 550 nm. Bouessay et al. optimized the PLD conditions to prepare NiO films, i.e., PO₂ = 0.1 mbar and Ts = 25 °C, and analyzed the electrochromic reversibility associated with the Ni3+/Ni2+ redox couple [91]. Using a laser fluence of 1 to 2 J·cm−2, which corresponded to an ablation rate of 0.9 Å·s−1, NiO films with a cubic rock-salt structure (Fm-3m space group) were formed. Porous PLD NiO films were prepared using nickel foil as the target in a low oxygen atmosphere (PO₂ = 50 Pa) [92,93] and were applied as the electrode for supercapacitors, showing a high specific capacitance of 835 F·g−1 at a 1 A·g−1 current density.
Similarly to LCO, the PLD growth of stoichiometric LiNiO2 with an α-NaFeO2 layered structure requests an Li-enriched target, i.e., LiNiO2 + 15%Li2O. López-Iturbe and coworkers attempted to avoid Li loss by using an Ar atmosphere of PAr = 10 mTorr and laser fluence of 15 J·cm−2 [94], while Rao et al. introduced pure oxygen (PO₂ = 0.1 Torr) in the PLD chamber and ablated the target at a laser fluence of 10 J·cm−2 [95]. The LNO films prepared at Ts = 700 °C exhibited an initial discharge capacity of 175 mC·cm−2·µm−1. Yuki et al. used, as the oxygen evolution reaction (OER), electrocatalysts, which were prepared LNO films from a target composed of Li2O and NiO2 sintered at 1000 °C for 8.5 h in air ablated by a Nd:YAG laser operating at 532 nm [96]. Recently, PLD LixNi2−xO2 thin films with 0.15 ≤ x ≤ 0.45 deposited on a glass substrate under a pressure of 0.1 Pa and annealed at 350 °C were grown by PLD with an LiNiO2 structure. The films appeared to be entirely made of particles even in the cross-section (grain size of 95 nm for x = 0.45). The average surface roughness estimated from the AFM measurements decreased with an increasing x, reaching a value of 0.615 nm for x = 0.45 [97].

3.3. LiNi1−yCoyO2 (NCO)

The LiNi1−yCoyO2 (0 < y <1) system with a layered α-NaFeO2 structure belongs to a LiCoO2-LiNiO2 solid solution with a higher reversible capacity than LCO and better cycleability than LNO. Among these substituted oxides, Ni-rich LiNi0.8Co0.2O2 (NCO) has been identified as one the most attractive cathodes [98]. In this context, several works investigated the growth of NCO thin films using pulsed-laser deposition. Dense PLD NCO films grown at Ts > 400 °C exhibited a gravimetric density of 4.8 g·cm−3 [99]. Ramana et al. grew NCO films deposited on an Ni foil substrate at temperatures of 25 ≤ Ts ≤ 500 °C under PO₂ = 6 to 18 Pa from Li-rich ceramic (15 mol% Li2O excess to avoid NiO or Co3O4 impurity phases) [100]. At Ts ≤ 300 °C, the PLD film showed the highest intensity of the (00l) reflection, which indicates that the c-axis was normal to the film surface. The XRD (003) diffraction peak at 2θ = 18.5° corresponds to an interplanar spacing of 0.145 nm. Phase diagram mapping (Figure 6) was proposed to highlight the effect of the growth temperature on the microstructure of PLD LiNi0.8Co0.2O2 films. Galvanostatic titration carried out at a rate of C/30 in the potential range of 2.5 to 4.2 V showed a discharge capacity of 82 µAh·cm−2 µm−1, which compares with the theoretical value of 136 µAh·cm−2·µm−1 (490 mC·cm−2·µm−1) [101].
Hirayama et al. fabricated NCO films using the standard PLD conditions (Φ = 100–220 mJ, PO₂ = 3.3 Pa, target composition Li/Ni(Co) = 1.3, and Ts = 600–650 °C) at a deposition rate of 0.3 nm·min−1 on an oriented SrTiO3 (STO) substrate. Microstructural analysis shows a misfit of ca. 5% and roughness of 1 to 3 nm for the film grown with an in-plane orientation at Ts = 600 °C. AFM imaging revealed the surface modification for films cycled in the voltage range of 2 to 5 V [102]. PLD NCO films (0.62 µm thick) were electrochemically characterized by galvanostatic titration (GITT), cyclic voltammetry (CV), and electrochemical impedance spectroscopy (EIS). Films grown at Ts = 600 °C under PO₂ = 13 Pa with a laser fluence of Φ = 450 mJ per pulse exhibited an average specific capacity of ~60 µAh·cm−2·µm−1 and the Li+ diffusion coefficient varied from 3 × 10−13 to 2 × 10−10 cm2·s−1. After 100 cycles, the electrode showed a capacity retention of 85% [103]. The kinetics of the Li-ion intercalation in PLD NCO films grown on an Nb-doped STO substrate at Ts = 600 °C under PO₂ = 3.3 Pa were investigated by EIS [104]. Nyquist plots showed changes of the electrode impedance as a function of the Li extraction/insertion with a larger value at a potential of 4.2 V. Baskaran et al. prepared NCO films on Pt and Si substrates heated at Ts = 500 °C under a low oxygen partial pressure of 0.21 Pa [105]. The 40-min deposited films (120-nm thick) displayed a specific discharge capacity of 69.6 µAh cm−2·µm−1 (145 mAh·g−1) after 10 cycles. Based on these results, a Li-ion microbattery was fabricated with a LNCO/Li3.4V0.6Si0.4O4(LVSO)/SnO configuration with a thickness of ~1.5 μm. Such a microcell delivered a capacity of 16.1 µAh·cm−2·µm−1 after 20 cycles.

3.4. LiNi1−yMnyO2 (NMO)

The substitution of Mn for Ni has been demonstrated to be beneficial for LiNiO2 cathode materials. LiNi0.5Mn0.5O2 (NMO) thin films were prepared on stainless steel and gold substrates from an Li-enriched target with an Li/(Ni + Mn) ratio of 1.5 in the NiO + MnO2 mixture. Under standard conditions (Φ = 2 J·cm−2, Ts = 550 °C, and PO₂ = 266 Pa), impurity-free and (00l)-textured PLD films (300–500 nm thick) were obtained after an annealing process at 650 °C [106]. Galvanostatic charge–discharge tests showed that NMO films deposited on stainless steel displayed an electrochemical response, with a large voltage plateau between 2.5 and 3 V attributed to the presence of spinel phases (i.e., LiMn2O4 and LiNi0.5Mn1.5O4), while NMO films prepared on Au substrate showed the typical fingerprint of the LiMO2 layered compound with a single plateau at ca. 3.7 V. The analysis of the kinetics from CV measurements in the 2.5 to 4.5 voltage range led to a diffusion coefficient of the Li+-ions of D* = 3.13 × 10−13 cm2·s−1 for the Li insertion and 7.44 × 10−14 cm2·s−1 for the Li extraction. GITT results showed that D* is highly dependent on the electrode potential in the range of 10−12 to 10−16 cm2·s−1 [107]. Sakamoto et al. addressed the growth of epitaxial PLD NMO films with an orientation of the basal layered plane (BLP) that depends on the SrTiO3 (STO) substrate plane: The BLP is parallel to the STO(110) substrate, while the BLP is perpendicular to the STO(111) substrate. The relationship between the lattice parameters and applied voltage highlights the charge–discharge processes for both orientations [108]. In a second article, Sakamoto et al. examined the structural properties of the surface and bulk of LiNi0.5Mn0.5O2 epitaxial thin films during an electrochemical reaction using in situ X-ray scattering [109]. In normal conditions, the two-dimensional diffusion of Li during (de)intercalation proceeds for the (110) plane. However, 3D-diffusion activity can be observed for a high degree of cation mixing (Ni2+ in Li(3a) sites).

3.5. Li(Ni, Co, Al)O2 (NCA)

The growth of LiNi0.8Co0.15Al0.05O2 (NCA) has been envisaged because Al-doping provides excellent structural and thermal stability for the electrode with the suppression of phase transitions. NCA thin films were prepared on Ni substrates at Ts = 500 °C by PLD with an energy and laser-pulse repetition rate of 300 mJ and 10 Hz, respectively, under PO₂ = 18 Pa. The PLD target (i.e., pellet pressed at 1.5 to 5.0 × 103 kg·cm−2), optimized by using bulk NCA Li-enriched with 15 mol% Li2O as the precursor, was sintered at 800 °C for 24 h [110]. The Li//NCA microcells delivered an initial specific capacity of 92 µAh·cm−2·µm−1. The kinetics of Li+ ions in PLD films measured by the GITT method in the voltage range of 2.50 to 4.25 V vs. Li+/Li revealed a diffusion coefficient of 4 × 10−11 cm2·s−1 with a maximum of 1 × 10−10 cm2·s−1 for the composition of Li0.7Ni0.8Co0.15Al0.05O2. Figure 7 displays the specific capacity of Li//LiNi0.8Co0.15Al0.05O2 cells as a function of the substrate temperature. PLD films were grown onto various substrates at 25 ≤ Ts ≤ 500 °C under a controlled O2 atmosphere (PO₂ = 50 mTorr). NCA thin films prepared onto Ni foil at Ts = 500 °C exhibited the best specific discharge capacity of 100 μA·cm−2·μm−1.

3.6. Li(Ni, Mn, Co)O2 (NMC)

Lithiated nickel-manganese-cobalt oxides, LiNi1−yzMnyCozO2 (NMC), is a complex LiNiO2-LiMnO2-LiCoO2 solid solution, which displays the same structure as rock salt, α-NaFeO2, with a valence state of cations as Ni2+, Mn4+, and Co3+. LiNi1/3Co1/3Mn1/3O2 (NMC333) thin film electrodes were prepared by pulsed laser deposition on a Pt/Ti/SiO2/Si substrate (where the 150-nm thick Pt and the 100-nm thick Ti films act as the current collector and buffer layer, respectively) at room temperature under a PO₂ of 6.6 Pa. The deposition rate was about 4.4 nm·min−1. Impurity-free NMC333 films were obtained using an Li-enriched target (15% excess Li2O). The charge–discharge profiles strongly depended on the film morphology that was tuned by increasing the annealing temperature from 400 to 500 °C. The best electrochemical features were obtained for annealing at 450 °C, showing a discharge plateau of about 3.7 and 3.6 V [111]. Epitaxial and highly textured LiNi0.5Mn0.3Co0.2O2 (NMC532) thin film cathodes were deposited by a one-step PLD process [112]. Using a laser fluence of Φ = 6 J·cm−2, Ts = 750 °C, and an Li-enriched target (20% Li2CO3 excess), the films were deposited on silicon (111), stainless steel (SS), and c-cut sapphire (0001) substrates. It was stated that highly dense and epitaxial films with a strong (003) reflection peak intensity are achieved at high Ts. Films grown on an Au-SS substrate delivered 125 mAh·g−1 at 0.5C and demonstrated a 72% capacity retention after 100 cycles. Furthermore, 3D-electrode architectures were fabricated, and are applicable to power hearing aids. Structured NMC333 film electrodes (50–80 µm thick) were prepared by laser printing using a pulsed laser (λ = 355 nm) at a fluence of 50 to 100 mJ·cm−2. Such films exhibit a stable discharge capacity at a low C-rate (0.1C), but for a current density of 1C, the electrode reached 40% of the initial capacity [113]. Recently, Abe et al. elucidated the deterioration mechanism of pristine ZrO2-coated NMC333 thin films prepared by PLD on a (110)STO substrate maintained at two temperatures of 25 and 700 °C; the oxygen pressure ranged from 3.3 to 10 Pa [114]. Characterization was performed by cyclic voltammetry, in situ X-ray diffraction, and in situ neutron reflectometry. As a result, ZrO2 coating suppressed the low activity of the spinel phase formed at the surface and hence improved the cycleability of the thin film electrode.

3.7. Li-Rich Layered Oxides

The growth of Li-rich layered oxide Li1.2Mn0.55Ni0.15Co0.1O2 on SRO/STO(100) and SRO/STO(111) substrate (where SRO and STO are SrRuO3 and SrTiO3, respectively) heated at Ts = 600 °C was reported by Bendersky et al. [115]. The transmission electron microscopy (TEM) images recorded using a high-angle annular dark field (HAADF) mode showed a predominant Li2(Mn,Ni,Co)O3 monoclinic phase. Johnston-Peck et al. confirmed the growth of monoclinic Li-rich thin films (C2/m space group) using a target with the composition of Li1.2Mn0.55Ni0.15Co0.1O2 and displayed the CV response with a non-aqueous electrolyte at a scan rate of 0.1 mV·s−1 [116]. Yan et al. intended to develop a PLD thin film electrode using an Li-rich layer structured oxide with a composition of Li1.2Mn0.54Ni0.13Co0.13O2 (or written as 0.55Li2MnO3·0.45LiNi1/3Co1/3Mn1/3O2) [117]. Films deposited at Ts = 650 °C under PO₂ = 46 Pa and annealed at 800 °C showed the best electrochemical performance with an initial specific discharge capacity of 70 µAh·cm−2·µm−1. However, the differential capacity curves, dQ/dV, indicated that the layered structure gradually changed to the spinel phase during the charge–discharge cycling.

3.8. Li2MO3 (M = Mn, Ru)

Pulsed-laser deposited Li2MnO3 thin films at various thicknesses (12 < δ < 48 nm) were grown using a KrF excimer laser on Nb:SrTiO3(111) substrates from an Li-rich Li3.2MnO3 target. With synthetic conditions (Ts = 650 °C, PO₂ = 75 Pa, and Φ = 0.8–1.1 J·cm−2) and a laser frequency of f = 1 to 5 Hz, PLD Li2MnO3(111) films exhibited a single-phase with the C2/m symmetry. The results of ICP analysis gave a composition of Li1.90MnIVO2.95, which indicates lithium and oxygen vacancies. The highest discharge capacity of 300 mAh·g−1 was delivered after 50 cycles by a 12.6-nm thick film [118]. PLD epitaxial Li2RuO3 thin films were successfully prepared with a (010) and (001) orientation on STO(110) and (111) substrates heated at Ts = 500 to 550 °C under PO₂ = 3.3 Pa. The initial charge–discharge capacities calculated using a theoretical density of 5.15 g·cm−3 were 120 and 105 mAh·g−1 at a 0.5C rate for the (010) and (001) orientation, respectively [119,120].

3.9. LiMn2O4 (LMO)

Since the early report in 1996 [47], numerous studies have been devoted to LiMn2O4 (LMO) thin films grown by laser ablation. A spinel structure (Fd3m space group) was successfully prepared using different PLD conditions and applied as a positive electrode in thin-film lithium microbatteries (see, for example, [121]). Most prior works focused on the fundamental properties of PLD-grown LMO cathode films, aiming to deposit the highly structured and porous morphology required for a good operating electrode [7,122,123,124,125,126,127,128,129,130,131,132,133,134,135,136]. Morcrette et al. analyzed PLD film composition as a function of Ts and PO₂ using the Rutherford backscattering method and nuclear reaction analysis [122]. A stoichiometric LMO film was obtained at Ts = 500 °C and PO₂ = 20 Pa, while the Li/Mn ratio decreased with Ts and increased with PO₂. The film of Li0.6Mn2O3 (20% oxygen loss) was obtained under vacuum. The Mn3O4 + MnO and Mn3O4 + LiMn2O4 mixed phases were successively grown in the range of 0.01 ≤ PO₂ ≤ 1 Pa. Julien et al. defined the conditions of the disposition of LMO films grown on Si substrates. Impurity-free well-crystallized samples with a crystallite size of 300 nm were obtained at Ts as low as 300 °C and PO₂ = 10 Pa using an Li-enriched target (15% Li2O excess) to avoid Li deficiency in the film [123]. The electrochemical features of the Li cell showed a specific capacity as high as 120 mC·cm−2·µm−1 in the voltage range of 3.0 to 4.2 V vs. Li+/Li, which was attributed to the high degree of film crystallinity [7,123]. Simmen et al. studied the relationship between Li1+xMn2O4−δ thin films and Li excess in the target and concluded that a film deposited from a composite target of Li1.03Mn2O4-δ + 7.5 mol% Li2O was the best, exhibiting a discharge capacity of 42 µAh·cm−2·µm−1 [127]. Various substrates were successfully used for the epitaxial growth of LiMn2O4 (LMO) spinel thin films, such as Pt, Si, Au, MgO, Al2O3, and SrTiO3. Gao et al. [128] reported a detailed mechanism of the epitaxial LMO film/substrate (current collector) interface formation. A coherent hetero-interface was formed with the substrate but a tetragonal Jahn–Teller distortion was observed, induced by oxygens’ non-stoichiometry and the lattice misfit strain. PLD epitaxial LMO thin films were deposited on oriented Nb:SrTIO3 substrates maintained at 950 °C from a sintered target with 100 wt.% excess Li2O with different surface morphologies and orientations, such as (100)-oriented pyramidal, (110)-oriented rooftop, or (111)-oriented flat structure. The pyramidal-type LMO was cycled at a 3.3C rate, demonstrating a specific capacity of 90 mAh·g−1 after 1000 cycles [129]. Using oriented substrates, i.e., (111)Nb:SrTiO3 (STO) and (001)Al2O3 single crystal, LMO films were grown with the (111) orientation under the following synthesis conditions: Li-enriched target (Li/Mn = 0.6), Ts = 650 °C, and PO₂ = 30 Pa [130]. Electrochemical tests emphasized the interactions between the films and substrate, and showed a plateau voltage at 3.6 to 3.8 V for LMO/STO and 3.8 to 4.1 V for LMO/alumina. Canulescu et al. [131] investigated the mechanisms of laser-plume expansion during the PLD of LMO films under PO₂ ranging between 10−4 and 20 Pa. The Li deficiency occurs as a result of the different behavior of the species at elevated Ts. Hussain et al. [132,133] obtained highly oriented LiMn2O4 thin films on oriented Si substrates heated in the range of 100 ≤ Ts ≤ 600 °C under PO₂ = 10 Pa and with a laser fluence of Φ =10 J·cm−2. Grains with a spherical shape (around 230 nm in diameter) changed to a flake-like structure at Ts = 600 °C (Figure 8a). The grain size varied almost linearly with the substrate temperature (Figure 8b).
By applying an elevated-temperature PLD technique, Tang et al. [134] studied the influence of the substrate temperature (Ts) and the oxygen partial pressure (PO₂) on LMO film crystallinity. LMO thin films deposited on Si (001)/0.2 µm-SiO2 substrates at 575 °C under 13 Pa oxygen had a flat and smooth surface and exhibited mainly a (111) out-of-plane preferred texture (Figure 9a). Such films of the 300 nm thickness showed a very dense cross-section (density ~4.3 g·cm−3) (see Figure 9b) [135]. The effect of stoichiometric deviations on the electrochemical performance of an LMO thin-film cathode was investigated by Morcrette et al. [136], while the kinetics of Li+ ions in the LMO thin-film framework were documented by Yamada and coworkers [137]. A high-activation barrier of 50 kJ·mol−1 for Li-ion transfer was identified at the electrode/electrolyte interface for films deposited at Ts = 700 °C and PO₂ = 27 Pa. Albrecht et al. [138] reported the minimum crystallization temperature of spinel LMO thin films in a narrow annealing temperature range of around 700 °C. Electrochemical tests carried out with the galvanostatic cycling with the potential limits (GCPL) method proved that Li-ions are (de-)intercalated in different tetrahedral sites for which the processes occur at potentials that are slightly shifted by U ≈ 100 mV, which is similar to the previous results by Julien et al. [7].
Studies of the structure and electrochemical reactivity of heteroepitaxial LiMn2O4/La0.5Sr0.5CoO3 (LMO/LSCO) bilayer thin films deposited on crystalline SrTiO3 substrates show that LSCO reduced the lattice misfit strain with the substrate and favored a lower LMO surface roughness. However, a decrease of the electrical conductivity occurred during the electrochemical test (after first cycle) due to the lattice oxygen loss at the outermost layer (40 nm) [139]. Tang et al. reported a comparative investigation of the structures, morphologies, and properties of Li insertion for LMO films with different crystallizations. At Ts = 400 °C, LMO films consisted of nanocrystallites < 100 nm in size with rough surfaces that exhibited a discharge capacity of 61 µAh·cm2·µm−1 with a capacity loss of 0.032% per cycle up to 500 cycles, while for Ts = 600 °C and PO₂ = 10 Pa, highly crystallized films showed an initial discharge capacity of 54.3 µAh·cm2·µm−1 [134]. The intrinsic properties of PLD-grown LMO have been investigated by several techniques. Electrical measurements of LMO films showed that the conductivity is sensitive to Ts, as the activation energy that followed the Mott’s rule increased with Ts up to Ea = 0.64 eV at Ts = 600 °C [133]. Singh et al. characterized the crystallinity and texture of LMO films deposited at Ts = 650 °C. Here, (111)-oriented films were grown on a doped Si substrate, while films deposited on a stainless-steel substrate exhibited a (001) orientation [140]. The thermo-power (or Seebeck coefficient) of PLD LMO films was reported to be 70 µV·K−1 [141].
PLD LMO films were subjected to an overcharge (5 V vs. Li+/Li), which did not modify the structure and preserved the well-resolved voltage peaks at 4.1 and 4.2 V, while an overdischarge (2 V vs. Li+/Li) led to a loss of capacity due to the structural disorder associated with the tetragonal transition, i.e., Jahn–Teller distortion [142]. Singh patented the fabrication of PLD Li1−xMyMn2−2zO4 films, where M is a doping element (M = Al, Ni, Co, Cr, Mg, etc.) and x, y, and z vary from 0.0 to 0.5 [143]. These defective spinel structures enhanced the oxygen content as compared to LiMn2O4 crystal. In particular, the oxygen-rich Li1−δMn2−δO4 films were superior cathode films, leading to excellent rechargeable battery performances. It is claimed that a high discharge rate of 25C produces only a 25% capacity loss and a specific capacity >150 mAh·g−1 remains after 300 cycles. Rao et al. reported the preparation of well-crystallized LMO films at a high substrate temperature of Ts = 700 °C and PO₂ = 13 Pa that delivered a capacity of 133 mC·cm−2·µm−1 at a very slow C/100 rate [144]. Several workers reported the evolution of the thin-film electrode/electrolyte interface, as the planar form of the film is the ideal geometry for such investigations [145,146,147,148]. Room temperature impedance measurements were carried out to identify the formation of the solid electrolyte interface (SEI) layer on a PLD LMO film cathode and the degradation mechanism during cycling in an aprotic electrolyte containing LiPF6 salt. A reversible disproportionation reaction was suggested with the formation of the Li2Mn2O4 and λ-MnO2 phases at the surface [146]. Using epitaxial-film model electrodes, Hirayama studied the surface reaction and the formation of the SEI layer and the interfacial structural reconstruction during an initial battery process using in situ surface X-ray diffraction and reflectometry [147]. TEM images confirmed the surface reconstruction that occurred during the first charge, i.e., when a potential was applied. After 10 cycles, the SEI layer was observed on both the (111) and (110) surfaces and Mn dissolution appeared at the (110) surface [148]. Inaba et al. [149] investigated the surface morphology evolution of PLD LMO thin films grown on a PT substrate at Ts = 600 °C by electrochemical scanning tunneling microscopy (STM) with voltage cycling in the range of 3.5 to 4.25 V. The original LMO grains of 400 nm in size coexisted with small particles 120 to 250 nm in size, which appeared after 20 cycles and decreased to ~70 nm after 75 cycles through a kind of dissolution/precipitation process. LMO thin-film electrodes with a grain size of <100 nm deposited on a stainless steel substrate at Ts = 400 °C under a 26 Pa oxygen partial pressure displayed an excellent capacity of 62.4 µAh·cm−2·µm−1 when cycled at a 20 µA·cm−2 current density in the voltage range of 3.0 to 4.5 V. A very low capacity fading was recorded for up to 500 cycles at 55 °C. Li+-ion diffusion coefficients evaluated from EIS measurements were around 2.7 × 10−12 cm2·s−1 for an electrode charged at 4.0 V and 2.4 × 10−11 cm2·s−1 for 4.2 V [150]. Xie et al. [151] investigated the Li+-ion transport in LMO thin films (~100 nm thick) grown on Au substrates at 600 °C at a deposition rate of 0.14 nm·min−1. The chemical diffusion coefficients determined by the EIS, GITTm, and PITT methods were in the range of 10−14 to 10−11 cm2·s−1 in the voltage range of 3.9 to 4.2 V. Table 5 lists some typical results on the kinetics of Li+ ions in pulsed-laser deposited LMO thin films.
The electrochemical behavior of Li-rich spinel Li1.1Mn1.9O4 thin films grown by PLD on an Au substrate was reported by several workers [156,157]. The best performance was reported at a discharge current density of the 36C-rate for LMO films deposited for 30 min in PO₂ = 30 Pa and Ts = 600 °C with an Nd:YAG laser (266 nm) adjusted to an energy fluence of 1 J·cm−2 [156]. Nanocrystalline LMO films with grains less than 100 nm were deposited on a stainless-steel substrate at Ts = 400 °C and PO₂ = 26 Pa using a PLD pulse power of 100 mJ at the frequency of 10 Hz. The film cycled over 100 cycles delivered a specific capacity of 118 mAh·g−1 at a current density of 100 A·cm−2 [157]. Using reflectometry measurements, Hirayama et al. [158] studied the structural modifications at the electrode/electrolyte interface of a lithium cell, in which the LMO electrodes were prepared as epitaxial films by the PLD method with different orientations. The respective orientation of the LMO film corresponded to that of the substrate plane, i.e., the (111), (110), and (100) planes of the SrTiO3 substrate. No density change was observed for the (110) and (100) planes, whereas a defect layer was detected in the (111) plane. ZrO2-modified LiMn2O4 thin films prepared via PLD consisting of amorphous ZrO2 formed on the grain boundary and the outer layer of the LMO matrix [159]. The high capacity retention of 82% after 130 cycles of films of xZrO2-(1−x)LiMn2O4 (x = 0.025) monitored at the 4C rate was attributed to the decrease of the charge transfer resistance (Rct).
Epitaxial LiMn2O4/La0.5Sr0.5CoO3 (LMO/LSCO) bilayer thin films with sub-nano flat interfaces were deposited on (111)-oriented STO substrates at Ts = 650 °C in PO₂ = 4 Pa. After the first charge–discharge cycle, the decrease of the electrical conductivity of the LSCO buffer layer due to lattice oxygen loss induced capacity fading [139]. The PLD growth of a multilayer LMO thin film electrode demonstrated the compensation of lithium loss during deposition [160]. Such a sample prepared in the PLD conditions (Ts = 650 °C, Φ = 530 mJ·cm−2, and PO₂ = 1.3 Pa) showed the typical two pairs of voltammetry peaks at 0.82 and 1.02 V vs. Ag/AgCl in an aqueous cell. Kim et al. [161] prepared a Li0.17La0.61TiO3/LiMn2O4 (LLTO/LMO) hetero-epitaxial electrolyte/electrode by PLD with an energy fluence of Φ = 1.7 J·cm−2 in a PO₂ = 6.6 Pa atmosphere. The typical herostructure is composed of a 17.5-nm thick LMO, 7-nm thick interfacial layer, and 26.5-nm thick LLTO deposited on a (111)-oriented STO substrate. Voltammograms of the first and second cycles displayed redox peaks around 3.8 V attributed to an oxygen-deficient LMO and around 4.0 and 4.2 V, which are the typical redox voltages of the LMO spinel. Suzuli et al. [162] prepared multi-layer epitaxial LiMn2O4/SrRuO3 (LMO/SRO) thin film electrodes deposited for 30 min via PLD on (111)STO substrates heated at 650 °C using an Li1.2Mn2O4 target in PO₂ = 6.6 Pa. The LMO(33 nm)/SRO(38 nm) film exhibited a discharge capacity of 125 mAh·g−1 with the typical plateau regions of LMO in the charge–discharge reaction. Yim et al. substituted Sn for Mn in PLD LMO thin films [163]. The LiSnx/2Mn2−xO4 films were prepared on a Pt/Ti/SiO2/Si(100) substrate in the conditions of Ts = 450 °C, PO₂ = 26.7 Pa, Φ = 4.6 J·cm−2, 10 Hz pulse frequency, and 4 cm target-substrate distance. XPS and EXAFS measurements showed that Sn2+ cations replace Mn3+ ions, which resulted in an increase of the valence of Mn in the spinel lattice. A high specific capacity of ∼120 mAh·g−1 and cycleability with a capacity retention >81% at the 4C rate after 90 cycles was attributed to the Mn-deficient structure. A multi-layer PLD process was utilized to deposit LMO films (90 nm thick) on Si-based substrates coated with Pt as the current collector [164]. A reversible capacity of 2.6 µAh·cm−2 (corresponding to a specific capacity of ≈28 µAh·cm−2·µm−1 or 66 mAh·g−1 assuming a dense film with 4.3 g·cm−3) was reached at an extremely high current density of 1889 µA·cm−2 (equivalent to the 348C rate) with a capacity retention of 86% over 3500 cycles. A significant non-diffusion-controlled contribution (pseudocapacitive-like) was evidenced by cyclic voltammetry; however, the two typical voltage plateaus in the GCD of LMO (around 4 V) indicates that the faradaic redox reaction is the main process. For an easy comparison, Table 6 lists the electrochemical properties of PLD-prepared LMO thin film electrodes.

3.10. LiNi0.5Mn1.5O4 (LNM)

LiNixMn2−xO4 is a substituted oxide spinel that operates at high voltages >4.5 V upon Li extraction. Substituted spinel films of LixMn2−yMyO4 where M = Ni, Co and 0 ≤ y ≤ 0.25 as-prepared with a crystalline morphology (0.3 µm thick) showed the typical features of high-voltage electrodes without carbon additive or binder materials in the range of 2.0 to 5.8 V vs. Li+/Li [165]. Cyclic voltammetry showed that: (i) PLD LiMn1.9Ni0.1O4 films charged at 5.7 V do not show capacity fading; (ii) LiMn2O4 and LiMn1.75Co0.25O4 films present a good stability to 5.6 and 5.4 V vs. Li+/Li, respectively; and (iii) below 3 V the films exhibit the typical Jahn–Teller distortion. The compound, LiNi0.5Mn1.5O4 (LNM), is more interesting because of the oxidation state of cations: Ni2+ can be oxidized twice (i.e., 2e transfer) during charge while Mn4+ is electrochemically inactive. Xia et al. showed that laser-ablated LNM films 0.3 to 0.5 µm thick deposited on a stainless steel substrate heated at 600 °C under an oxygen partial pressure of 26 Pa exhibit excellent capacity retention (i.e., ~120 mAh·g−1 after 50 cycles) in the voltage range of 3 to 5 V vs. Li+/Li [166]. Well-crystallized oxygen deficient LiMn1.5Ni0.5O4−δ films deposited by PLD at a controlled fluence of Φ = 2 J·cm−2, Ts = 600 °C, and PO₂ = 26 Pa for 40 min exhibited a stepwise voltage profile near 4.7 V and a small plateau in the 4 V region. These disordered spinel structures had a stable specific capacity of 55 µA h cm−2·µm−1 in the voltage range of 3 to 5 V vs. Li+/Li. The good rate capability was due to the high kinetics for Li diffusion in the range of 10−12 to 10−10 cm2 s−1 measured by the potentiostatic intermittent titration technique (PITT). These values are comparable to that of layered LCO [167,168]. Epitaxial LNM films were grown on single-crystal oriented SrTiO3 (STO) substrates from an Li-enriched target with Li/(Ni + Mn) = 0.6. The film orientation, i.e., (100)-, (110)-, and (111)-oriented, were the replica of those of the STO substrates [169]. Depending on the film orientation and thickness, the discharge profiles exhibited two to three plateaus around 3.9, 4.5, and 4.7 V vs. Li+/Li, which were attributed to the Mn3+/Mn4+, Ni2+/Ni3+, and Ni3+/Ni4+ redox couples, respectively. Note that the emergence of the Mn3+/Mn4+ redox couple was due to the introduced oxygen vacancies [170].
Other 5-V class cathode thin films include LiCoMnO4. PLD LiCoMnO4 films prepared under standard conditions (i.e., Ts = 500 °C, PO₂ = 20–100 Pa, and Φ = 2 J·cm−2) had a composition of Li:Co:Mn = 0.99:0.98:1. These films were tested in the voltage range of 3.0 to 5.5 V vs. Li+/Li in SSMB consisting of Li/Li3PO4/LiCoMnO4 fabricated on Pt/Cr/SiO2 substrates [171]. Cyclic voltammetry showed that the higher capacity in the 5-V region was obtained for the film grown under PO₂ = 100 Pa (Figure 10). A specific discharge capacity of 90 mAh·g−1 remained after 20 cycles. Epitaxial Li0.92Co0.65Mn1.35O4 film with a cubic spinel structure was grown on a SrTiO3(111) single-crystal substrate using a layer-by-layer technique, which consisted of repeating a Li1.2Mn2O4/Li1.4CoO2 deposition process at Ts = 650 °C and PO₂ = 6.6 Pa using a KrF excimer laser (λ = 248 nm) [172]. For a film area of 0.7 mm2, thickness of 33.4 nm, and density of 4.38 g cm−3, the specific discharge capacity was 340 mA·g−1 at the second cycle. A capacity retention of 80% was observed after 20 cycles.

3.11. MnO2

Due to its environmental compatibility and low cost, manganese oxides are promising candidate materials for supercapacitor applications using neutral aqueous solution as the electrolyte (i.e., 0.5 mol·L−1 K2SO4) [173]. Xia et al. prepared a dense Mn3O4 spinel thin film grown by PLD, which transformed to nanoporous MnOx after electrochemical lithium insertion/extraction. After 2000 cycles, the MnOx film deposited at Ts = 600 °C under PO₂ = 26 Pa demonstrated a specific capacitance of 193 F·g−1 at a current density of 5 A·g−1 [174]. The fundamental aspects of the redox reaction were investigated on amorphous MnOx and crystalline Mn2O3 films prepared by PLD onto Si and (316)-stainless steel substrates [175]. Using standard conditions (Φ = 2–3 J·cm−2, PO₂ = 13 Pa, Ts = 200–500 °C) Co-doped MnOx films were grown from a hybrid Co3O4/Mn3O4 target. Non-doped and 3% Co-doped amorphous films (i.e., MnOx and Mn0.97Co0.03Ox film) exhibited a specific capacitance of 45 and 99 F·g−1 at a 5 mV·s−1 scan rate, respectively. The V2O5/Mn3O4 target was used to prepare PLD-grown V-doped MnOx thin films with a V content in the range of 3.3 to 10 atm. %. The specific capacitance of the crystalline PLD Mn2O3 film reached the value of 290 F·g−1 at a 1 mV·s−1 scan rate, while the 10 atm. % V-doped film (Mn0.9V0.1O2) had a higher specific current value [176].

3.12. LiMPO4 (M = Fe, Mn) Olivines

LiFePO4 (LFP) thin-film electrodes have been successfully fabricated by pulsed-laser deposition [156,177,178,179]. It was shown that, due to the film thickness and carbon content, the electrochemical performances are very sensitive, i.e., electronic conductivity and Li-ion diffusion. Iriyama et al. reported the PLD growth of olivine structured LFP thin films and their electrochemical properties characterized by cyclic voltammetry and charge–discharge tests [180,181]. The typical olivine features were evidenced by CV measurements in the range of 2.0 and 5.0 V vs. Li+/Li, i.e., a single couple of anodic and cathodic peaks at ~3.4 V. Song et al. synthesized PLD LFP films with a low carbon content (<1 wt. %) on stainless steel substrates utilizing an Ar atmosphere [182]. The 75-nm thick films showed reversible cycling of more than 80 mAh·g−1 after 60 cycles. Furthermore, 156-nm thick films grown using a target–substrate distance reduced to 5 cm had a layered surface texture and delivered more than 120 mAh·g−1 with a good capacity retention. LFP thin films with a needle-like morphology were prepared by an off-axis PLD technique [183]. The effect of the substrate on the structure and morphology was examined by Palomares et al. for PLD film deposited under argon gas kept at a pressure of 8 Pa [184]. Stainless steel was demonstrated to be the best substrate for the single-phase olivine (Pnma space group) with a temperature set at 500 °C.
LiFePO4 deposited by pulsed-laser deposition proved to be effective as a thin film electrode. Tang et al. stated that a well-crystallized pure olivine phase was grown using optimized deposition parameters (Ts = 500 °C, PAr = 20–30 Pa, pulse power of 120 mJ, pulse frequency of 10 Hz, λ = 248 nm) [185]. An electrochemical capacity of 38 µAh cm−2·µm−1 at the C/20 rate (36 µAh·cm−2·µm−1 at a rate of C/4) was measured at 25 °C. High substrate temperatures (500 ≤ Ts ≤ 700 °C) favored the presence of Fe3+ impurities, i.e., Li3Fe2(PO4)3 and Fe4(P2O7)3. In a second article, the same group analyzed the kinetics of Li+ ions in PLD LFP films using CV, GITT, and EIS measurements [179]. CV data provided average D* values of 10−14 cm2·s−1, while D* deduced from both GITT and EIS techniques was in the range of 10−14 to 10−18 cm2·s−1. A maximum D* value was observed at x = 0.5 for LixFePO4. Lu et al. prepared different composite thin films, i.e., LiFePO4–Ag and LiFePO4–C, with the aim of enhancing the electronic conductivity [177,178]. It was found that films grown with 2 mol% carbon and annealed at 600 °C for 6 h had an improved coulombic efficiency. Well-crystallized olivine-type structure LFP films were obtained by PLD coupled with high temperature annealing of 650 °C. The first discharge capacity was 27 mAh·g−1 with a retention of only 49% after 100 cycles. The low reversible capacity and poor cycling performance was attributed to the existence of an Fe2O3 impurity produced by the high temperature treatment and poor intrinsic conductivity [186]. Sauvage et al. published several reports on the electrochemical properties of PLD LFP thin films grown in different configurations [187,188,189,190]. First, it was shown that well-crystallized and homogeneous 300-nm thick LFP films deposited on Pt-capped Si substrates have intrinsic Li insertion properties evaluated both in aqueous and non-aqueous electrolytes, i.e., voltage plateau at 3.42 V vs. Li+/Li [187]. Second, the influence of the film thickness was studied in the range of 12 to 600 nm [188]. Third, the effect of the texture on the electrochemical performance was analyzed for PLD films deposited on a polycrystalline α-Al2O3 substrate coated with a 20-nm thick Pt layer from an LiFePO4 pellet as the target. The standard PLD conditions were used (i.e., (Φ = 2 J·cm−2, PAr = 8 Pa, Ts = 600 °C) [189]. Finally, the electrochemical stability of LFP films was analyzed as a function of the exposition to the most common lithium salt and for different current collectors (i.e., Si, Pt, Ti, Al, and (304)-stainless steel) [190]. A 270-nm thick film tested by CV at a 2 mV·s−1 scanning rate in 1 mol·L−1 LiClO4 in EC/DMC solution delivered a specific capacity of 1.52 µAh cm−2 after 150 cycles. Recently, Raveendran et al. reported the properties of FeSe and LiFeO2/FeSe bi-layers prepared by PLD as cathode materials [191]. Mangano-olivine LiMnPO4 (LMP) thin films were fabricated on Pt-coated SiO2 glass substrates using PLD parameters, e.g., Φ = 1.58 J·cm−2, Ts = 400–700 °C, and PAr = 2–100 Pa [45]. LMP films (50-nm thick, 0.09 cm2 area) were applied in Li/Li3PO4/LiMnPO4 microbatteries for 500 cycles. From the CV measurement, a capacity of 28 mAh·g−1 at 20 mV min−1 was reported.

3.13. V2O5

Another candidate material for the cathodes of microbatteries is V2O5, in which about 1 mol of Li+ ions can be inserted and extracted without the phase transformation of V2O5, leading to a theoretical specific capacity of 147 mAh·g−1. Due to its stable layered structure and its ability to accommodate large amounts of Li ions, V2O5 has been widely studied for the development of electrochromic displays, color memory devices, and lithium-battery cathodes [192]. Extensive works have evidenced the advantages of PLD for the preparation of V2O5 films with a good reproducible stoichiometry similar to the target material [193,194,195,196,197,198,199,200,201,202,203,204,205,206,207]. The first work related to the growth of V2O5 thin film by PLD as an electrode for a thin-film lithium battery was reported by the National Renewable Energy Labs (USA) [193] followed by Julien’s group [194,195]. A major advantage of laser ablation deposits is that it is possible to prepare thin layers of crystallized V2O5 under oxygen at a relatively low temperature of 200 °C [193]. The growth mechanism of PLD V2O5 thin films has been proposed by Ramana and coworkers [196]. It was reported that the grain size, surface roughness, and global morphology are highly sensitive to the nature and temperature of the substrate for films deposited in an oxygen partial pressure of PO₂ = 13 Pa. The functional influence of the growth temperature on the grain size for films deposited onto various substrates was also evidenced. Two main features should be pointed out: (i) The exponential variation of the grain size over the substrate temperature range of 25 to 500 °C; (ii) the variation is dependent on the substrate material, which is larger for the Si(00) wafer. McGraw et al. reported that pulsed-laser deposited V2O5 films can be grown on a number of low-cost substrates, including SnO2-coated glass, on which highly textured (001) films are obtained at Ts = 500 °C under PO2 in the range of 0.2 to 0.5 Pa [52,197,198]. PLD thin films of V2O5 were prepared for applications in lithium batteries using a ceramic V2O5 target and a KrF laser of a wavelength of 248 nm. Depending on the temperature of the substrates and the oxygen pressure during deposition, amorphous or crystallized layers are obtained. PLD-grown amorphous films exhibited a low capacity loss of ~2% over 100 discharge–charge cycles in the voltage range 4.1 to 1.8 V compared to 20% for crystalline film [45,199]. Thin layers of V2O5 were also prepared using a V2O3 target [200]. By making deposits at 200 °C with the same V2O3 target, amorphous layers were obtained in the absence of oxygen and layers crystallized in the presence of oxygen. Madhuri et al. [201] reported the successful crystallization of laser-ablated V2O5 thin films at Ts = 200 °C. These films were grown in the orthorhombic structure and exhibited a predominant (001) orientation. The growth of crystalline thin dense films without post-deposition annealing was claimed and the good electrochemical performance of PLD films was demonstrated. Iida et al. [202] addressed the electrochromic properties of V2O5 films deposited onto ITO glass as a function of the PLD parameters. The film recrystallization occurred in the range of 400 ≤ Ts ≤ 500 °C and the best morphology was obtained for PO₂ = 13.3 Pa. McGraw et al. deposited thin films of V2O5 for applications in lithium batteries using a ceramic V2O5 target and a KrF laser, with a wavelength of 248 nm. Depending on the temperature of the substrates and the oxygen pressure during deposition, amorphous or crystallized layers were obtained [193,199]. Thin layers of V2O5 were also prepared using a V2O3 target. By making deposits at 200 °C with the same V2O3 target, amorphous layers were obtained in the absence of oxygen and layers crystallized in the presence of oxygen. Stoichiometric amorphous V2O5 films can be grown onto substrates maintained at low temperatures (Ts < 100 °C) using a sintered V2O5 target. Ramana et al. revealed that stoichiometric V2O5 films can be grown with a layered structure onto amorphous glass substrates at temperatures as low as 200 °C and an oxygen partial pressure of 100 mTorr [203]. The onset of crystallization occurred at 200 °C with an activation energy of 0.43 to n0.55 eV [204]. Correlations between the growth conditions, microstructure, and optical properties were investigated for V2O5 thin films deposited over a wide substrate temperature range of 30 to 500 °C by Rutherford backscattering spectrometry (RBS), X-ray photoelectron spectroscopy (XPS), scanning electron microscopy (SEM), and UV−vis−NIR spectral measurements. As shown in Figure 11, the film grain size follows a power law of the substrate temperature and the optical energy bandgap decreases from 2.47 to 2.12 eV with the increase of Ts from 30 to 500 °C [205]. Bowman and Gregg investigated the effect of the applied strain on the resistance of V2O5 thin films grown from both metallic vanadium and a ceramic V2O5 target using a laser fluence of ~3.0 and ~1.5 J cm−2, respectively [208]. Deng et al. compared the growth of V2O5 films using femtosecond (f-PDL) and nanosecond (n-PDL) pulsed laser deposition using SEM, XRD, and Raman spectroscopy. Prior to annealing, f-PLD films showed a rougher texture and nano-crystalline character, while n-PLD films were much smoother and predominantly amorphous [209].
The PLD growth conditions were refined by an analysis of the surface properties for the production of high-quality V2O5 films. The investigations were carried out by AFM, SEM, FTIR, and XRD. AFM measurements showed a surface roughness of ~12 nm with a Gaussian-like height distribution of surface grains for films deposited at Ts = 200 °C under PO₂ = 10 Pa [210]. The local structure of 0.3-µm thick films grown on Si(100) substrates was characterized by Raman spectroscopy [211]. The influence of the deposition temperature on the microstructure was investigated by an examination of the rigid layer-like mode at 145 cm−1, which showed a frequency shift with increasing Ts. The ability of the V2O5 thin film lattice to accommodate Li+ ions was also investigated by Raman spectroscopy. The appearance of the δ- and γ-phases of LixV2O5 gave additional insight into the structural changes of lithiated films. From the photoluminescence spectra, Iida et al. evidenced a blue shift of the vanadyl V = O peak upon Li+ insertion for different electric charge in the range of 0 ≤ Q ≤ 20 mC of Li+ [212]. From X-ray diffraction and Raman spectroscopy data, Shibuya et al. derived a PO₂Ts phase diagram for V-O films grown on Si(100) substrates (Figure 12). The composition of V-O films was as follows: (i) A VO2 monoclinic phase was formed at Ts ≥ 450 °C and PO2 in the range of 5 to 20 mTorr; (ii) a V2O5 orthorhombic phase was obtained under oxidative conditions, i.e., at high PO₂; (iii) a V6O13 phase was grown under PO₂ between oxidative and reductive conditions; and (iv) metastable V4O9 and VO2(B) phases were formed for lower Ts (≤400 °C) and lower PO₂ (≤30 mTorr) [213].
V2O5 thin films have been widely used as electrochromic electrodes but few reports are devoted to PLD-grown films. Fang et al. obtained thin films deposited on In2O3:SnO2 (ITO)-coated glass and (111)Si wafer from a V2O5 target using an XeCl laser with a wavelength of 308 nm for applications in electrochromic devices [214,215]. Electrochromic tests over 60,000 cycles showed that a significant change in the optical density (bleached and colored states) was evaluated to be 0.13 at λ = 600 nm for as-prepared films at Ts = 200 °C. Crystallized c-axis oriented V2O5 films were obtained under oxygen and at a substrate temperature of 200 °C. The durability without long-term degradation of the electrochromic V2O5 films was tested over 8000 cycles in the voltage range of 1.2 to 1.4 V [216]. Ti-doped V2O5 thin films prepared by the pulsed laser ablation technique at Ts = 200 °C and Φ = 2 J·cm−2 were studied as the electrode for an electrochromic display that exhibits a neutral brownish blue color. The long-term durability was verified over 8000 cycles of a voltage cycled in the range from −1.0 to +1.0 V vs. SCE showing a charge of 35 mC·cm−2. The good cycleability was attributed to the layered structure of PLD crystalline films with a parallel orientation to the substrate, suitable for Li+-ions’ transport [215]. PLD thin films of the system, WO3-V2O5, were prepared with a laser fluence of 1 to 2 J·cm−2 on SnO2/F-coated glass substrates at Ts = 25 °C under PO₂ = 0.1 mbar. Such films with low V contents cycled in the protonic medium. The true color neutrality is the main advantage of V-based WO3 thin films; however, the cell capacity and coloration efficiency decrease with an increase of the V content [217]. The orthorhombic V2O5 phase is also applied as electrodes for sensors. Huotari reported that pure PLD films were obtained at Φ = 2.6 J cm−2, Ts = 400 °C, and PO₂ = 1.0 Pa with a post-annealing treatment at 400 °C for 1 h in normal ambient conditions [218]. The efficient response to NH3 at part-per-billion levels. indicates these films use as possible sensing materials for ammonia gas [219].
The electrochemical properties of V2O5 thin-film cathode material have been widely studied in cells with aprotic electrolytes (typically LiClO4 dissolved in propylene carbonate). The electrochemical charge–discharge profiles of PLD V2O5 films were also found to be dependent on Ts, exhibiting a marked difference for V2O5 films grown at Ts < 200 °C when compared to those grown at Ts ≥ 200 °C. The effect of the substrate temperature and hence the microstructure on the kinetics of the lithium intercalation process in V2O5 films is remarkable. The applicability of the grown PLD V2O5 films in lithium microbatteries indicates that PLD V2O5 films in the temperature range of 200 to 400 °C offer better electrochemical performance than films grown at other temperatures due to their excellent structural quality and stability [25,220]. As an experimental fact, pulsed laser deposited V2O5 thin films exhibit a higher initial voltage than the crystalline material, i.e., ~4.1 vs. ~3.5 V (Li+/Li). For instance, V2O5 thin-film cathodes, deposited from a V6O13 target at a fluence of ~12 J cm−2 on SnO2-coated glass at Ts = 200 °C, were efficient for Li+-ion incorporation. In (h00)-textured films, the specific capacity reached values between 50% and 80% of the theoretical value. On the other hand, amorphous films display a stable capacity corresponding to 1.2 F mol−1 in the voltage range of 4.1 to 1.5 V. Prior textured V2O5 films discharged beyond the threshold to 2.0 V vs. Li+/Li showed an immediate and continuous capacity fading and a quasi-total amorphization after 10 cycles [193,197]. The chemical diffusion coefficient of Li+ ions, D*, measured by PITT was found to be in the range of 1.7 × 10−12 to 5.8 × 10−15 cm2·s−1 in crystalline V2O5 films, which compares well to the value found in LixV2O5 phases, whereas D* displayed a smooth and continuous decrease as the Li content increased in amorphous films [198].
In an attempt to apply PLD V2O5 films in SSMB, a thin-film microbattery was constructed using a glassy Li1.4B2.5S0.1O4.9 electrolyte film with an ionic conductivity of 5 × 10−6 S·cm−1 at 25 °C and an Li anode film. This Li/Li1.4B2.5S0.1O4.9/V2O5 cell delivered a capacity of ~400 mC·cm−2·µm−1 at a current density of 15 µA·cm−2 [221]. Ag0.3V2O5 and LiPON thin films with a smooth surface were grown by PLD in an N2 and O2 atmosphere, respectively. The Li/LiPON/Ag0.3V2O5 SSMB displayed good cycleability at a current density of 7 µA·cm−2 in the voltage window of 1.0 to 3.5 V. The specific capacity was maintained at 40 µAh·cm−2·µm−1 after 100 cycles [222]. Recently, amorphous vanadium oxide a-VOx PLD films (650 nm thick) were grown on stainless steel substrates from a V2O5 PLD-target under PO₂ in the range of 0 to 30 Pa. Films prepared under PO₂ = 13 Pa had a smooth surface and bore an O/V atomic ratio of 2.13 with a higher atomic percentage of V5+ than that of V4+. Electrochemical tests carried out in Li cells with 1 mol·L−1 LiPF6 in ethylene carbonate (EC) and diethyl carbonate (DEC) (1:1 by volume) as the electrolyte showed a reversible specific capacity as high as 300 mAh·g−1 at the C/10 current rate and a capacity retention of 90% after 100 cycles [223]. Such studies were initiated by Zhang et al. in 1997 to obtain VOx films PLD grown at 200 °C and exhibiting a specific capacity of 340 mAh·g−1 at a current density of 0.1 mA·cm−2 and a capacity loss <2% at the end of 100 cycles [201]. A summary of the electrochemical properties of PLD-grown vanadium oxide thin film electrodes is given in Table 7.

3.14. V6O13

With the ability of vanadium cations (two V4+ every V5+) to be reduced, the mixed-valence vanadium oxide, V6O13, the structure of which is formed by alternated single and double layers of VO6 units, can insert reversibly about 6 mol of Li, giving a specific capacity of 311 mAh·g−1. It makes this compound a good candidate for the cathode material of rechargeable batteries [224]. V6O13 films were fabricated by the PLD technique using a pulsed KrF excimer laser (λ = 248 nm, 20 ns pulse duration, 10 Hz frequency, and 4 J·cm−2 laser fluence). A (100)-oriented Si substrate was maintained at a temperature of 500 °C. During the PLD process, the formation of crystalline V6O13 films (dark-bluish color) and the vanadium oxidation state (+2.166) was monitored by controlling the processing temperature and O2 partial pressure. The (002)-oriented V6O13 thin films (50 nm thick) were obtained after a post annealing at 400 °C under an oxygen partial pressure of 100 Pa [221]. The discharge profiles for Li//V6O13 thin-film cells were recorded in the voltage range of 3.3 to 2.5 V at a current density of 5 μA cm−2 (Figure 13). A film as-grown at Ts = 250 °C exhibited a steady discharge curve with an insertion uptake of 6Li per V6O13 formula unit, whereas the cell voltage decay was faster for a film deposited at Ts = 25 °C. However, the films deposited at Ts = 250 °C and annealed at 300 °C in an Ar atmosphere displayed a stepped discharge profile with the appearance of a voltage plateau at ca. 3.02 and 2.85 V vs. Li+/Li.

3.15. FeF2

Iron fluoride is a conversion-type cathode material with a high theoretical specific capacity of 571 mAh·g−1. Several groups reported electronic additive-free FeF2 films grown by the PLD technique at low temperatures [225,226,227,228]. The electrochemical properties of FeFx films were reported to be dependent on the substrate temperature. Using an FeF3 target, crystallized-like FeF2 film (P42/mnm space group) was obtained at Ts = 600 °C, while a mixed FeF3-FeF2 phase was grown at Ts = 25 °C and single FeF3 phase was prepared at Ts = −50 °C [226]. FeFx deposited on stainless steel substrates under vacuum (5 × 10−5 Pa) exhibited a capacity of ~600 mAh·g−1 at a current density of 0.56 µA·cm−2. Santos-Ortiz et al. reported the PLD growth of polycrystalline FeF2 thin films on oxide-etched Si(100) and glass substrates using standard conditions (Ts = 400 °C, Φ = 8 J·cm−2, growth rate of ~6 nm·min−1) [227]. A 50-nm thick PLD FeF2 film on stainless steel substrates held at 400 °C showed an initial specific discharge capacity of 167 mAh·g−1 when cycled 200 times in the potential range of 1 to 4 V vs. Li+/Li at the 1C current rate [228].

3.16. MoO3

MoO3 is an attractive cathode material for microbattery technology from several standpoints: (i) The orthorhombic a-phase is a layered structure favorable for Li insertion between slabs; (ii) Mo has the highest +6 oxidation state, making the high structural stability; (iii) the lattice can be reversibly inserted up to 1.5Li per mole of oxide, yielding a specific capacity of 280 mAh·g−1; and (iv) the capacity of the dense film can reach a value of ≈130 µAh·cm−2·µm−1, almost twice the value for LiCoO2 [229]. In addition to the use as cathode batteries, MoO3 is a material applied in electrochromics, gas sensors, and electro-optics. For certain applications, high-quality films grown by PLD are required.
Currently, PLD MoO3 thin films are grown using a KrF excimer laser (λ = 248 nm) with a fluence of 2 J cm−2 (energy of 300 mJ per pulse) and deposited on various substrates heated in the range of 25 ≤ Ts ≤ 500 °C under an atmosphere of O2 flow maintained at a pressure of 0.1 ≤ PO₂ ≤ 20 Pa. In the prior report, Julien et al. showed that the structure analyzed by optical spectroscopy strongly depends on Ts: For Ts < 150 °C, an amorphous phase is formed, the β-MoO3 phase grows at Ts ≈ 200 °C, and the layered α-MoO3 phase appears at Ts = 300 °C [230,231,232,233]. Al-Kuhaili et al. reported the growth of polycrystalline MoO3 films on unheated substrates using both XeF and KrF excimer lasers. By tuning the annealing temperature in the range of 300 to 500 °C, both the grain size and surface roughness increased. Films formed using the XeF laser (λ = 351 nm) and annealed at 400 °C have the best stoichiometry of MoO2.95 [233]. Analyzing the growth mechanism, Ramana and Julien concluded that the thermochemical reaction during ablation strongly influences the structural characteristics of PLD MoO3 films. Above Ts = 400 °C, the formation of compositional defects induces structural disorder, i.e., α-β-MoO3−x phase mixture [234,235].
The applicability of PLD films to an Li microbattery was demonstrated by the best electrochemical features: A discharge capacity of 90 µAh cm−2 µm−1 was obtained for Ts = 400 °C, while only 53 µAh·cm−2·µm−1 was delivered for Ts = 200 °C [236]. Puppala et al. investigated the microstructure and morphology of PLD MoO3−x thin films’ growth for catalytic applications using a femtosecond laser (f-PLD) and a nanosecond excimer-laser (n-PLD). Substantially textured films with a partially crystalline phase prior to annealing were obtained by the f-PDL laser, while the n-PLD-grown MoO3−x films were predominantly amorphous with a smooth surface [237]. Sunu et al. claimed that as-deposited PLD films (Ts = 400 °C, Φ = 4–5 J·cm−2, repetition rate of 15 to 20 Hz, and PO₂ = 500 Pa) are suboxide-like, i.e., mixture of η-Mo4O11 and χ-Mo4O11, which transformed to MoO3 after annealing at 500 °C in air for 5 h [238]. Several works reported the PLD growth of films (MoO3)1−x(V2O5)x with 0.0 ≤ x ≤ 0.3 prepared at room temperature under an oxygen pressure of 13.3 Pa. The effect of the V2O5 content on the coloring switching properties for thermochromic, gasochromic, photochromic, and electrochromic applications was investigated [239,240]. Contrary to pure MoO3, the electrochromism of MoO3-V2O5 films showed that the Mo oxidation state (+6) did not change considerably upon Li+ insertion, while V5+ was reduced considerably to V4+ [239]. A similar improvement of the gas-sensing properties, i.e., the shortest response time and highest transmittance change, was observed for V2O5-doped MoO3 films under an H2 atmosphere [240].

3.17. WO3

Tungsten oxide (WO3) belongs to the class of “chromogenic” materials, i.e., materials exhibiting coloration effects through electro-, photo-, gas-, laser-, and thermochromism processes, which requires the high homogeneity provided by the PLD technique. Preliminary studies of the growth of WO3 thin films by PLD were first attempted by Haro-Poniaowski et al. [233] in 1998. Later, Rougier et al. reported the PLD conditions for the growth of efficient WO3 films as electrochromics (EC) components [241]. The microstructure of films deposited on SnO2:F coated glass substrate is strongly sensitive to both the oxygen pressure and substrate temperature: (i) Crystallized films are formed for Ts = 400 °C and PO₂ = 10 Pa; (ii) amorphous films are obtained for PO₂ = 1 Pa at any Ts; (iii) for Ts = 25 °C and PO2 = 1 Pa, WO3 films are blue colored and conductive; and (iv) colorless insulator films are grown for Ts = 25 °C and PO₂ = 10 Pa, which display the best electrochromic properties. Qiu and Lu showed that oxygen deficient WO3−δ films with a deviated monoclinic structure were produced using PLD parameters as 2.5 J·cm−2, PO₂ = 26 Pa, and a target-Si(100) substrate distance of d = 5 cm [242]. Ramana et al. investigated the structural transformations of PLD WO3 as a function of the annealing treatment. Using standard conditions (Φ = 2 J·cm−2, Ts = 300 °C, PO₂ = 13.3 Pa), films deposited on glass substrates (200–500 nm thick) showed an atomic ratio of O/W ≈ 2.96 ± 0.05. The monoclinic phase of the as-prepared film transformed to an orthorhombic phase at 350 °C and to a hexagonal phase at 500 °C [243,244]. By varying the substrate temperature in the range of 150 to 800 °C and the oxygen pressure from 1 to 40 Pa, Mitsugi et al. obtained WO3 films with a different microstructure: Amorphous, crystallized tetragonal, and triclinic phases [245]. Hussain et al. obtained amorphous, polycrystalline, and nanocrystalline WO3 phases, and iso-epitaxial WO3(00l) thin films deposited on single-crystal SrTiO3 substrates at 600 °C and under PO₂ = 18 Pa [246]. Suda et al. deposited PLD WO3 thin films on flexible ITO substrates. They showed that films, prepared at Ts < 300 °C, are amorphous and polycrystalline phases were obtained at Ts > 400 °C, while the crystallinity of the film on glass substrates was not dependent on PO₂ [247]. Films deposited at 400 °C were porous with a nanocrystalline triclinic structure and showed the best cycleability [216,248,249].
The suitability of PLF WO3 films for EC applications was investigated as a function of the partial oxygen pressure during deposition. Studies of the texture and morphology of PLD 30-nm thick WO3 films deposited on Si(100) and SrTiO3(100) substrates under an O2 background of 2.5 Pa showed that: (i) The laser fluence (in the range of 5 to 15 J·cm−2) strongly influences the texture, (ii) the films grown on STO are biaxially textured with a smooth surface, and (iii) films deposited on Si are granular [250]. The fabrication of WO3 thin films with color neutrality for applications as EC materials was realized by the deposition of films containing 20% of vanadium onto SnO2:F coated glasses at Ts = 20 °C under PO₂ = 10 Pa. The blue color in the reduced state (−0.4 V) of the W-O-V films lost intensity and turned grey-blue (transmittance of 50%) as the V concentration increased [251]. Highly transparent WO3 films exhibiting strong coloration and fast and full bleaching were prepared under PLD conditions (Φ = 1 J·cm−2, Ts = 250 °C, PO₂ = 16 Pa, and d = 40 mm) [252]. WO3 films were also prepared using similar PLD parameters for applications in gas sensors [253,254,255].

4. Solid Electrolyte PLD Films

For the development of solid-state thin film batteries, thin films of solid electrolytes with excellent performances, i.e., high ionic conductivity (σi), good stability against the lithium anode, large electrochemical window (ΔV), and poor electronic conductivity (σe), are currently required. To fulfill these requirements, the thin films of oxide-, phosphate-, or sulphide-based solid electrolytes were grown by the PLD technique [256,257,258]. The facile manufacture of such thin films is due to the easy control of the PLD chamber’s atmosphere. Table 8 lists some typical solid electrolyte thin films prepared by PLD [41,259,260,261,262,263,264].

4.1. LiPON

In the early 1990s, Bates et al. prepared Li3PO4 thin films using a sputter-deposition technique in the presence of N2 gas that resulted in a nitrogen-doped lithium phosphate (called LiPON) of a typical chemical composition, Li3.3PO3.9N0.17 to Li2.9PO2.9N0.7. The structure consists of doubly and triply coordinated nitrogen atoms, which form cross-links between the phosphate chains [20]. LiPON displays a high chemical stability and an ionic conductivity of 2 × 10−6 S·cm−1 at 25 °C [265]. The growth of LiPON thin films by pulsed-laser deposition is also realized in nitrogen partial pressure with a moderate laser power influence [259,266].
Zhao et al. reported the growth LiPON thin films on three different substrates (i.e., Si wafer, Au-coated Si, and Al-coated glass plate) by reactive PLD in an N2 gas atmosphere in the range of 50 to 200 mTorr using a Li3PO4 target. The target was ablated by the beam of a Nd:YAG laser at the fluence of 5 to 20 mJ·cm−2. The influence of the ambient N2 pressure and the laser fluence on the ionic conductivity was systemically examined and the best result of 1.6 × 10−6 S·cm−1 with an activation energy of 0.58 eV at 25 °C was obtained for a film prepared under 200 mTorr at Φ = 15 J·cm−2. The mechanism of the nitridation of Li3PO4 was carried out by XPS measurements, showing that σi increases with the N/P ratio [259]. West et al. showed that a 17-nm thick layer of LiPON deposited at the solid electrolyte–electrode interface decreased the charge-transfer resistance from 4470 to 760 cm−2 in a Li/LiPON/LNM cell. The PLD amorphous films with σi = 1.5 × 10−8 S·cm−1 at 25 °C were deposited from a crystalline Li2PO2N target under the flow of N2 gas at PN₂ = 1 Pa [267].

4.2. LixLa2/3+yTiO3−d (LLTO)

Solid electrolytes, such as lithium lanthanum titanium oxides, LixLa2/3+yTiO3−δ (LLTO), based on a perovskite-like structure can accept vacancies at the Li (or La) and oxygen sites and show properties depending on the composition, with an electronic conductivity when Ti3+ cations (instead of Ti4+) are present and an ionic conductivity for Li-rich material. The typical growth of LLTO thin films fabricated by the laser ablation technique is obtained at deposition temperatures in the range of 600 to 800 °C under a controlled oxygen pressure from 0.1 to 100 Pa [268]. LLTO films, such as Li3xLa(2/3)−xTiO3, exhibit a high ionic conductivity of up to 10−5 S·cm−1 when deposited with pulsed laser deposition [269,270,271]. Li0.5La0.5TiO3 (LLTO) PLD thin films, prepared at 400 to 600 °C, are amorphous and show an ionic conductivity of ~2 × 10−5 S·cm−1 at room temperature. Contrary to crystalline films, the amorphous LLTO exhibits good stability in contact with lithium metal anodes. Half-cells based on LiCoO2 films covered with LLTO films deposited by pulsed laser deposition could be cycled for hundreds of cycles [269]. Furusawa et al. prepared amorphous LLTO films at Ts = 25 °C with a uniform thickness (0.46–0.63 µm) using a laser energy of 180 mJ per pulse at 10 Hz [270]. The authors stated a controlled pressure of ~10−6 Torr but did not mention the presence of O2 gas. The highest σi of 1.2 × 10−3 S·cm−1 (Ea = 0.35 eV) obtained for Li0.5La0.5TiO3 films deposited on an Ag substrate was due to the absence of grain boundaries. Maqueda optimized the PLD growth parameters to prepare La0.57Li0.29TiO3 dense films at Ts = 700 °C under PO₂ = 15 Pa with smooth surfaces [271]. The obtained nano-crystalline films exhibited domains, which are cubic and tetragonal modifications of the perovskite phase. Transport measurements showed an ionic conductivity of 8.2 × 10−4 S·cm−1 at 25 °C with Ea = 0.34 eV. Epitaxial Li0.33La0.56TiO3 solid electrolyte thin films were grown on NdGaO3(110) by PLD at Ts higher than 900 °C under PO₂ = 5 Pa [272]. These films showed a conductivity σi of 3.5 × 10−5 S·cm−1 at 25 °C with Ea = 0.35 eV (Figure 14).
The influence of different substrates and excess lithium in the target on the microstructure and ionic conductivity of PLD LLTO thin films was examined by Aguesse et al. [273] Despite a large lattice mismatch of up to +8.8% with the substrate, the epitaxial growth of LLTO is possible on different (001) oriented LaAlO3, SrTiO3, and MgO substrates using a sintered Li0.37La0.54TiO3 target and PLD parameters, such as Ts = 750–880 °C, PO₂ = 4–20 Pa, and laser fluence of 1.07 J cm−2. An ionic conductivity as high as 19.2 × 10−3 mS·cm−1 at 25 °C was obtained for 170-nm thick LLTO films grown on an STO substrate from an ablated 10 mol% lithium excess target. PLD LLTO films with a σi of 3 × 10−4 S·cm−1 and σe of 5 × 10−11 S·cm−1 were obtained by controlling the background PO₂ and Ts. Amorphous LLTO films were utilized in SSMB cycled up to 4.8 V vs. Li+/Li with high voltage LiNi0.5Mn1.5O4 spinel cathode thin films [274].
Another class of LLTO electrolytes consists of Ti-based solid electrolytes with a garnet-like structure, first reported by Weppner et al. [275]. Li6BaLa2Ta2O12 thin films were deposited on an MgO(100) substrate by the ablation of a target with a 5 mol% Li2O excess. In standard PLD conditions (Ts = 550 °C, PO₂ = 5 Pa, laser fluence of Φ = 2 J·cm−2, 40,000 laser pulses), an ionic conductivity of 2 × 10−6 S·cm−1 at 25 °C with an activation energy of 0.42 eV was obtained. This is comparable with the σi of LiPON. The electronic conductivity varied from 2.87 × 10−13 to 3.47 × 10−10 S·cm−1 in the range of the polarization voltage from 2.8 to 4.3 V [273]. Saccoccio et al. fabricated garnet Li6.4La3Zr1.4Ta0.6O12 films via PLD and studied the impact of PLD parameters (fluence of 1 to 4 J·cm−2, Ts in the range of 50 to 700 °C, and a post-annealing process) on the structural and transport properties. The ionic conductivity was measured by impedance spectroscopy. It was concluded that σi is not dependent on Ts but is strongly affected by the laser fluence [276]. Li7La3Zr2O12 (LLZO) garnet-like thin films were deposited on Si3N4/Si substrates at temperatures in the range of 50 ≤ Ts ≤ 750 °C under a fixed background of PO₂ = 1.3 Pa with a KrF excimer laser set at 0.6 J·cm−2 [277]. The best material, which exhibited an ionic conductivity of 6.3 × 10−3 S cm−1 at 400 °C (Ea = 0.6 eV), was obtained at Ts = 300 °C. The review of garnet-like solid electrolyte thin films grown via PLD is summarized in Table 9.

4.3. P- and Si-Based Electrolytes

Several phosphorus- or silicon-based oxides and sulfides are solid electrolytes for lithium batteries, such as Li3PO4, Li4SiO4-Li3PO4, Li2S-P2S5 glass ceramic, Li2+2xZn1−xGeO4 (LiSICON), Li3.25Ge0.25P0.75S4 (thio-LiSICON), etc., that can be prepared as thin films. Kuwata et al. prepared high quality Li3PO4 thin films by PLD for applications in Li/Li3PO4/LiCoO2 all-solid-state thin-film batteries. The Li3PO4 film exhibited an ionic conductivity of 4 × 10−7 S·cm−1 at 25 °C and an activation energy of 0.58 eV. This solid electrolyte showed an electrochemical stability in the potential range of 0.0 to 4.7 V vs. Li+/Li and was applied in Li/Li3PO4/LiCoO2 cells [14,282]. Amorphous PLD thin films of LiSICON display higher conductivities than that of Li4SiO4 and Li3PO4 films. The solid electrolyte 0.5Li4SiO4–0.5Li3PO4 dense films deposited on an Si wafer at Φ = 2–6 J·cm−2 under an argon gas of PAr = 0.01–5 Pa had an ionic conductivity of 1.6 × 10−6 S·cm−1 at 25 °C and an activation energy of 52 kJ·mol−1 [283]. Nakagawa et al. determined that PLD Li2SiO3 films stable to CO2 have an ionic conductivity of 2.5 × 10−8 S·cm−1 at 25 °C lower than that of Li2SiO3 films (i.e., 4.1 × 10−7 S·cm−1), which are unstable to CO2 [284]. PLD thin films of lithium meta-silicate (LSO) deposited at a growth rate of 0.17 Å per pulse on various substrates (i.e., SiO2, quartz, sapphire, Al2O3 ceramic, and MgO) from an Li2SiO3 sintered tablet were grown in the amorphous state. The ionic conductivity slightly depends on the substrate species with the best results (σi = 4.5 × 10−4 S·cm−1 at 300 °C, Ea = 0.88 eV) found for an 80-nm thick film deposited on SiO2 glass [262,285]. The PLD conditions for the growth of thio-LiSICON Li3.25Ge0.25P0.75S4 solid electrolyte thin films were carefully chosen (especially the Li content of 3.2 in the target, which maintains the number of Li vacancies) to obtain a high σi value of 1.7 × 10−4 S·cm−1 at 25 °C [265]. PLD 80Li2S-20P2O5 thin film prepared under PAr = 5 Pa exhibited an ionic conductivity and activation energy of 7.9 × 10−5 S·cm−1 and 43 kJ mol−1 at 25 °C, respectively. Heat treatment increased the σI to 2.8 × 10−4 S·cm−1 [286]. To avoid the formation of a Li-deficient phase, such as Li4P2S6, an Li2S-enriched Li3PS4 target was used to grow PLD solid-electrolyte thin films. Using an Li3.42PS4.21 target, PLD Li3PS4 films exhibited a higher ionic conductivity of 5.3 × 10−4 S·cm−1 at 20 °C [287].

4.4. PLD Electrolyte as Buffer Layers

Solid-state electrolyte (SSE) thin films have been used as a conductive buffer layer for the reduction of high resistance at the electrode/SSE interface of high-power all-solid-state lithium batteries. Coating the Li3PO4 thin films on electrode materials by the PLD method was found to be efficient for this purpose. Konishi et al. [288] reported the effect of surface Li3PO4 coating on LiNi0.5Mn1.5O4 epitaxial thin film electrodes. Amorphous Li3PO4 film (1–4 nm thick) was deposited at 25 °C with a laser energy of 150 mJ under PO₂ = 3.3 Pa. It was also pointed out that such a coating reduces the Mn dissolution in the non-aqueous electrolyte. Yubuchi et al. [289] fabricated the same coated electrode with Φ = 2 J cm−2 but under a lower oxygen gas pressure of 0.01 Pa. With a 100-nm thick Li3PO4 deposit, the total resistance of the Li cell decreased from 15 to 350 Ω. A PLD protective coating of 80Li2S–20P2S5 solid electrolytes on LiCoO2 particles was performed at room temperature under Ar gas at PAr = 5 Pa with a fluence of ca. 2 J·cm−2 (200 mJ per pulse). After SSE deposition for 120 min, the deposited film was ~150-nm thick, corresponding to 3 wt.% LiCoO2. Annealing the SSE deposit at 200 °C increased the capacity of the all-solid-state cell [42]. Another example of the buffer function of the Li2S−P2S5 solid electrolyte is given by the PLD coating of NiS-carbon fiber composite electrodes. The high ionic conductivity of 80Li2S−20P2S5 film deposited on an Si wafer was 7.9 × 10−5 S·cm−1 at 25 °C [290]. A capacity of 300 mAh·g−1 was delivered after 50 cycles at a current density of 3.8 mA·cm−2 (1C-rate). This SSE coating favors the lithium ion and electron conduction paths in the NiS framework. Ito et al. [291] successfully deposited Li2S–GeS2 thin films as the buffer electrolyte (σi = 1.8 × 10−4 S·cm−1) on LiCoO2 particles by the PLD technique. The amorphous 78Li2S–22GeS2 solid electrolyte thin films prepared using standard PLD conditions exhibited an ionic conductivity of 1.8 × 10−4 S·cm−1 at 25 °C. These SSE films were applied to form an electrode–electrolyte buffer interface with LiCoO2 [291]. The coating of a LiNbO3 SSE buffer coated onto the LiMn2O4 cathode resulted in an enhancement of the high rate capability and cycling stability of the electrode [292]. A similar process ensured a high thermal stability for the LiNi0.8Co0.15Al0.05O2 electrode operating over 500 charge–discharge cycles at 150 °C [293].

4.5. Li2O-V2O5-Si2O (LVSO)

PLD films of Li2.2V0.54Si0.46O3.4 are amorphous solid-state electrolytes of the system, Li2O-V2O5-Si2O (LVSO), which exhibits a conductivity of ~2.5 × 10−7 S·cm−1 at 25 °C [44]. Li4SiO4 thin films were successfully deposited by PLD using both an Nd:YAG laser (λ = 266 nm) and ArF excimer laser (λ = 193 nm) at the fluence of 2.5 J·cm−2 in the flow of O2 gas at PO₂ = 0.2 Pa. Having a conductivity of 4.1 × 10−7 S·cm−1 at 25 °C, thermally activated with Ea = 0.52 eV, these films were applied in SSMBs [44]. Zhao et al. prepared PLD Li–V–Si–O thin films’ electrolytes on an Si wafer and Al-coated glass as substrates placed 4 cm from the target. The film deposition was carried out at a fluence of 1.2 J cm−2 under an ambient of PO₂ = 6 Pa [294]. For Ts = 300 °C, the Li–V–Si–O film exhibited σi = 3.98 × 10−7 S·cm−1 at 25 °C and Ea = 0.55 eV. Workers at Kawamura’s lab reported the PLD growth of several LVSO solid electrolytes. The Li2.2V0.54Si0.46O3.4 film deposited with a continuous flow of O2 gas maintained at PO₂ = 0.2 Pa displayed an ionic conductivity of 2.5 × 10−7 S·cm−1 at 25 °C with an activation energy of 0.54 eV [41]. PLD amorphous 0.6(Li4SiO4)–0.4(Li3VO4) films deposited on Si(111) or fused silica plate exhibited an ionic conductivity of 10−7 S cm−1·at 25 °C, which is one order higher than the value for PLD Li2TiO3 film [295]. All-solid-state thin film batteries were fabricated using both LCO and LMO PLD film cathodes and amorphous LVSO solid electrolytes as shown in Figure 15 [121].

4.6. LiNbO3

Because of its high room-temperature ionic conductivity and low electronic conductivity (10−5 and 10−11 S·cm−1, respectively), LiNbO3 (R3c crystal structure) is considered as a good SSE for electrode coating [296]. LiNbO3 was applied as a buffer layer between an LCO cathode and thio-LISICON electrolyte (Li3.25Ge0.25P0.75S4). The resultant electrochemical cell showed low interfacial resistance and a high-rate capability [297]. A high quality was obtained for PLD LiNbO3 thin films deposited at 730 °C on sapphire substrates by using a relatively high oxygen partial pressure of PO₂ = 133 Pa and a laser fluence of 3 to 5 J cm−2 [298]. Contrastingly, Perea et al. prepared PLD LiNbO3 films using a lower laser fluence (0.8 to 1.6 J·cm−2) in a residual pressure of ≈4 × 10−4 Pa [299].

5. Negative Electrode PLD Films

5.1. TiO2

Due to the theoretical capacity of ~335 mAh·g−1 of titanium dioxide (comparable to ~372 mAh·g−1 for graphite and the small volume expansion (~4% for anatase)) significant interest has been devoted to the applied anode material in Li-ion batteries. The tetragonal anatase polymorph of TiO2 is a good anode candidate due to its insertion potential of around 1.5 V vs. Li+/Li [300]. Several works of the literature report the growth of TiO2 thin films with either a rutile or anatase structure fabricated by the PLD technique [301,302]. The growth conditions were studied on TiO2 films deposited by PLD using an Nd:YAG laser (532 nm wavelength beam) and a rutile-type TiO2 target. The effects of the substrate temperature (Ts) and oxygen partial pressure (PO₂) were investigated by Raman spectroscopy [13]. The parameters of Ts = 300 °C and PO₂ = 50 mTorr were optimized to obtain crystalline TiO2 films with a preferential (110) orientation. Kim et al. discussed the effects of the target morphology and target density on the size and distribution density of crystalline in PLD rutile-type TiO2 films deposited on (100)-oriented Si wafers maintained at 700 °C in a chamber with an oxygen partial pressure of 1.33 Pa [303]. A nearly particulate-free film was obtained from a dense target and the laser shots were adjusted for clear ripple patterns from the target surface. The optical bandgap energies of TiO2 PLD films grown on an α-Al2O3 (0001) substrate with an anatase and rutile structure were evaluated to be 3.22 and 3.03 eV, respectively [304]. Inoue et al. reported that films deposited at Ts = 150 °C have an anatase structure, while Ts = 300 °C provides rutile-type TiO2 films [305]. Choi et al. [306] prepared anatase TiO2 thin films with nanograins of 11 to 28 nm using a TiC target with Ts = 500 °C under 4 Pa O2 gas.

5.2. Li4Ti5O12 (LTO)

Li4Ti5O12 (LTO) cubic structure (Li[Li1/3Ti5/3]O4 in spinel notation), considered as a “zero-strain” anode material, exhibits the advantage of very minor volumetric changes (<0.2%) upon cycling. This electrode displays a large voltage plateau at ~1.5 V vs. Li+/Li and a theoretical specific capacity of 175 mAh·g−1 [307]. The first PLD growth of LTO thin films deposited onto Pt/Ti/SiO2/Si substrates using a KrF excimer laser beam (248 nm, 250 mJ) were reported by Deng et al. [308]. Films annealed at 800 °C (410 nm thick) exhibited a cubic structure with a lattice constant 8.375 Å larger than that of the LTO crystal (8.359 Å). The SEM cross-section image (Figure 16a) revealed the porous morphology induced by the high temperature treatment. The discharge specific capacity was the largest for films annealed at 700 °C due to the optimized adhesion strength between the film and substrate (Figure 16b). The anode films discharged at a current density of 10 µA·cm−2 (0.58C rate) showed excellent cycleability; the discharge capacity remained as 149 mAh·g−1 after 50 cycles. Li4Ti5O12 films (545 nm thick) deposited on conducting fluorine-doped tin oxide (LTO/FTO) with a crystallite size of 50 to 80 nm were investigated as electrochromic active material with the highest contrast at a wavelength of 705 nm (transmittance change of ~48%) [309]. Epitaxial LTO thin-film grown on SrTiO3 single crystal from an Li-rich target, Li5.2Ti5O12, have a structural orientation identical to the substrate and are impurity-free when deposited at Ts = 700 °C. The electrochemical features of LTO film anodes (20 nm thick) exhibited discharge capacities of ~200 and ~250 mAh·g−1 for the (100)- and (111)-orientation, respectively [310]. Kim et al. prepared nano-sized epitaxial LTO(110) deposited on Nb:SrTiO3(110) substrate. These films (~28 nm thick) were tested by cyclic voltammetry at a scan rate of 1 mV·s−1 and exhibited redox peaks at 1.53 and 1.60 V, corresponding to the insertion and extraction of Li+ ions. As-deposited films at a substrate temperature of 700 °C in a 6.6 Pa oxygen partial pressure exhibited a high initial capacity (~200 mAh·g−1) but poor stability [311]. Kumatani et al. investigated the PLD growth process of epitaxial LTO films deposited on an MgAl2O4 (111) substrate. With Ts = 800 °C and PO₂ = 1 × 10−3 Torr, LTO films had excellent crystallinity and a low resistivity of 3.3 × 10−4 Ω cm. at 25 °C. At lower PO₂, the PLD LiTi2O4 film was formed, while at higher PO₂, Ti was segregated as TiO2 rutile and Li0.74Ti3O6 [312].
Studies of the electrochemical performance and kinetic behavior of PLD LTO films deposited on Pt/Ti/SiO2/Si substrates were reported by Deng et al. [313]. Using an Li-rich target (i.e., excess 5 wt.% Li2O), the films annealed at 700 °C for 2 h in air were well-crystallized items with densely packed grains. The galvanic charge–discharge plateau was observed around 1.56 V and an initial specific capacity of 159 mAh g−1 was delivered with a retention of 93.7% after 20 cycles. The diffusion coefficient of Li+ ions in such an LTO framework was in the range of 10−15 to 10−12 cm2·s−1. The energy barrier of the diffusion of lithium ions was estimated to be Ea = 0.11 eV in LTO (111)-oriented PLD thin films (190 nm thick) grown on a spinel MgAl2O4 (111) substrate [314].
Zhao et al. reported the optical properties of epitaxially grown LTO films on (001)-oriented MgAl2O4 substrate. The optical bandgap of 3.14 eV was measured for 86 nm thick films (surface roughness of 4.61 nm) [315]. Schichtel et al. fabricated all-solid-state microbatteries with LTO as the positive electrode. PLD films were obtained on various substrates at Ts = 650 °C under a 0.3 Pa pure oxygen atmosphere using a commercially available LTO powder. As-prepared films (650 nm thick) revealed columnar growth that allowed a coulombic efficiency >97% after the second cycle and a discharge capacity of 33 µAh·cm−2 at a 3.5 µA cm−2 current density [43]. Pfenninger et al. demonstrated that LTO thin films deposited by PLD on an MgO substrate kept at 500 °C using a dense Li7.1Ti5O12 target sintered at 1000 °C for 12 h are compatible with the Li6.25Al0.25La3Zr2O12 electrolyte pellet. Such films display a stable structure and cycleability almost close to 175 mAh·g−1. The typical voltage plateau at 1.57 V (oxidation) and 1.53 V (reduction) was observed at a rate of 2.5 mA·g−1 [316]. Among the Li1+xTi1−xO4 ternary system, LiTi2O4 thin films were grown by the PLD route in the temperature range of 400 to 800 °C using a target with a higher Li/Ti ratio of 0.8 [317]. Chopdekar et al. grew epitaxial PLD LiTi2O4 thin films on various crystalline-oriented substrates, such as single crystalline substrates of MgAl2O4, MgO, and SrTiO3 [318]. The authors state the PLD conditions with Ts held at 450 to 600 °C in a vacuum of better than 5 × 10−6 Torr without any mention of the oxygen partial pressure, while Kumatani determined that stoichiometric LiTi2O4 thin films were obtained at a PO₂ of 5 × 10−6 Torr with Ts = 800 °C [312]. Recently, PLD LTO films grown on Nd-doped oriented STO substrates at Ts = 700 °C under PO₂ = 20 Pa showed high discharge capacities of 280 to 310 mAh·g−1. The best rate performance of 30C was obtained for the (100)-oriented Li4Ti5O12 films [319].

5.3. LiNiVO4

Amorphous LiNiVO4 thin-film anodes for microbatteries were grown by pulsed laser deposition using a sintered Li1.2NiVO4 target. The film grown at Ts = 25 °C and PO₂ = 8 mTorr showed the best electrochemical performance with a retainable capacity as high as 410 µAh·cm−2·µm−1 after 50 cycles [320].

5.4. TiNb2O7

An alternative to LTO, titanium-niobium oxide, TiNb2O7 (TNO), is considered a promising anode material for long life Li-ion batteries, due to its high Li+ ion transport, average voltage of 1.66 V, and theoretical capacity of ∼387 mAh·g−1 [321]. Fabrication of PLD TiNb2O7 thin films as anode electrodes for Li-ion micro-batteries was demonstrated by the ablation of a Nb2O5 + TiO2 mixture as a target at a laser fluence of 4.6 J·cm−2. Pure monoclinic TNO films were deposited on Pt/TiO2/SiO2/Si(100) substrates at 750 °C under an O2 gas of PO₂ = 6–13 Pa. The 380-nm thick films grown at PO₂ = 13 Pa delivered an initial specific capacity of 142 µAh cm−2 µm−1 at a current density of 50 µA·cm−2 with a 58% capacity retention after 25 cycles [322]. Recently, the same co-workers reported a high specific discharge capacity of 226 µAh·cm−2·µm−1 (~460 mAh·g−1) at a current density of 17 µA·cm−2 for amorphous TNO films grown by PLD. Li+ diffusion coefficient of ≈10−13 cm2·s−1 and an electronic conductivity of ≈10−9 S·cm−1 were also reported [323].

5.5. Silicon

Pulsed-laser deposited silicon thin films have been widely studied for applications in opto-electronics. With a large theoretical capacity (4200 mAh·g−1), silicon is also considered as a promising anode material for the replacement of graphite anode (LiC6, 372 mAh·g−1) for Li-ion batteries [324]. Despite the huge volume expansion of >300% during lithiation up to Li22Si5, it is possible to obtain anodes with Si thin films grown by physical vapor deposition (PVD), reaching a cycling life of up to 3000 cycles due to the limited volume change in the 2D film [325]. For example, a film deposited on Ni foil maintained a capacity of 3000 mAh·g−1 at a 12C rate over 1000 cycles [326]. PLD-grown Mg2Si thin film (30–380 nm thick) exhibited electrochemical activity with a stable cycling behavior in the voltage range of 0.1 to 1.0 V vs. Li+/Li; however, the initial irreversible capacity loss increased with the film thickness. The superior capacity of the 30-nm thick film was attributed to the formation of Li-Si alloys at the Si-rich surface [327]. Park et al. prepared PLD amorphous Si (a-Si) thin films on a stainless-steel substrate at temperature of 500 °C under an Ar gas pressure PAr = 6.5 mPa. Furthermore, 1.5-µm thick a-Si films were obtained at the growth rate of 25 nm·min−1. Electrochemical tests carried out in the voltage range of 0.005 to 1.5 V showed a first discharge capacity of 9 to 0.7 µAh·cm−2 with a 54.4% coulombic efficiency. Although, after 70 cycles, the 1-µm thick Si film exhibited a good cyclic performance [328]. Xia et al. reported the growth of a-Si using the standard conditions (Ts = 25 °C, P = 1.3 mPa, fluence of 150 to 160 mJ per pulse, deposition time of 30 min). Electrochemical tests showed that 120-nm thick a-Si films exhibited an initial charge capacity of ~64 µAh·cm−2 at a current density of 100 µA·cm−2, a discharge capacity of ~50 µAh cm−2 was maintained after 40 cycles, and the diffusion coefficient of Li ions determined from the cyclic voltammograms was ~10−13 cm2·s−1 [329]. Some Si-based composite thin films prepared by PLD combine the advantages of both components. The most popular are the carbon-based composites [330,331,332,333]. Chou et al. obtained a flexible anode material by the deposition of Si film onto single-wall carbon nanotubes (SWCNTs) using standard PLD conditions (λ = 248 nm, Ts ≈ 30 °C, Φ = 3 J·cm−2, PAr = 13 Pa, target–substrate distance of 50 mm). After 50 cycles, this Si/SWCNT nanocomposite delivered a specific capacity of 163 mAh·g−1 at a 25 mA·g−1 current density, which is 60% higher than for CNT [330]. The ultrathin film of Si grown by PLD was deposited on multilayer graphene (MLG) by CVD on an Ni foam substrate. The specific capacity of this binder-free Si/MLG anode appeared to be stable at ca. 1400 mAh·g−1 [331]. Silicon nitride SiN0.92 thin films were prepared by PLD and investigated as a negative electrode in lithium batteries. A 200-nm thick film grown on buffed stainless-steel substrates kept at 25 °C from an Si3N4 pellet as the target delivered a high specific capacity of 1300 mAh·g−1 after 100 cycles [334].

5.6. Graphene

Most of the commercial lithium batteries have a carbon anode. Graphene is the most conductive form of carbon, and as such, it is considered as a promising electrode, especially when it is doped with nitrogen [229]. A recent review was devoted to PLD-graphene synthesized from a solid carbon source [335]. Since nitrogen modifies the intrinsic properties of graphene, it is important to control its concentration. PLD, which allows for a one-step synthesis of N-doped carbon films, is particularly suited to this purpose. The first N-doped amorphous carbon film (a-C:N) synthesized by PLD dates from 2013 [336]. It contained 2 at. % of nitrogen. More recently, using the same approach, the upper nitrogen concentration in PLD a-C:N film was raised to 3.2 at.% [337]. These films, however, were not used as electrodes. On the other hand, a N-doped graphene (NG) electrode prepared by PLD coupled with in-situ thermal annealing (PLD-TA) was achieved by Fortgang et al. [338]. More recently, Bourquard et al. used the PLD-TA process to form an N-doped graphene film by high temperature condensation of the laser-induced carbon plasma plume onto the Si electrode previously covered by an Ni catalytic film [339], using a protocol published by the same group [340] Carbon was ablated at 780 °C from the graphite target using a femtosecond laser (λ = 800 nm, pulse width of 60 ns, repetition rate of 1 kHz, and Φ = 5 J·cm−2) at a distance of 36 mm from the graphite target, with PN = 10 Pa in the vacuum chamber. The electrochemical properties were measured with the thus-obtained 40 nm-thick film with a nitrogen concentration of 1.75% as the working electrode and an active area of 0.07 cm2, saturated calomel electrode as the reference electrode, and platinum as the counter electrode, in a 0.5 mol·L−1 1,1′ ferrocene-dimethanol solution of 0.1 mol·L−1 NaClO4. Aiming to detect H2O2 in 0.1 mol L−1 phosphate buffer saline (PBS) solution (pH 7.4), linear sweep voltammetry was used in the anodic range from 0 to 1000 mV with a scan rate of 100 mV·s−1. The electrode showed excellent reversibility, 60 mV, close to the theoretical value, and a detection limit of 1 mM of hydrogen peroxide, which constitutes a major improvement of the electroanalytical oxidation of H2O2 in comparison with undoped graphite electrodes. Such results are recent, and to our knowledge, no such electrode has been tested yet as an anode for lithium batteries.

5.7. Other PLD Anodes

Significant efforts have been devoted to the design and development of new PLD-grown anode materials for SSMBs, yielding a high energy density (from 500 to 1500 mAh·g−1) and electrochemical stability [341,342,343,344,345,346,347,348,349,350,351,352,353,354]. Table 10 summarizes the PLD-growth conditions and electrochemical properties of some new anodes proposed for lithium microbatteries. All transition-metal oxide MxOy materials are subjected to electrochemical lithiation via a conversion reaction, which implies a two-step process, i.e., first, fine metallic (M) nanoparticles embedded in an insulating matrix, such as Li2O, are in situ formed during the first (irreversible) discharge, and secondly, an alloying reversible reaction (Li-M) occurs on subsequent cycles [341].
Recently, Wu et al. proposed a novel anode consisting of a Li3P-VP nanocomposite fabricated by PLD [353] using a 5-Hz Nd:YAG laser (λ = 355 nm, Φ = 2 J·cm−2) concentrated on the target surface with an incident angle of 45°, with PAr =10 Pa in the vacuum chamber. The stainless-steel substrate was placed 3 cm from the surface and kept at 400 °C. The weight of the film thus obtained (without the substrate) was 0.10 mg·cm−2. The excessive lithium in this composite was used to stabilize the VP2 structure after the first charge. Electrochemical tests were made with this film as the working electrode, while lithium metal sheets were used as counter and reference electrodes with 1 mol·L−1 LiPF6 in EC:DMC (1:1 in volume) electrolyte. When cycled in the range of 0.01 and 4 V vs. Li+/Li at a current density of 5 μA·cm−2, a capacity of 1040 mAh·g−1 was delivered at the second discharge, 987 mAh·g−1 at the 50th cycle.

6. Discussion

There is general agreement on the beneficial results of the pulsed-laser deposition of thin films used as active materials in lithium microbatteries. This is a technique prone to fabricate thin film rapidly, from a small amount of target material, while maintaining the ideal stoichiometry by the control of different growth parameters. We observed (Table 2) that there is a strong trend to develop microcell technologies using LiCoO2 film (typical thickness of 4-µm) as the cathode, LiPON thin film (typical thickness of 1 µm) as the solid electrolyte, and Li thin film anode, which may have advantages in terms of the following key requirements: High energy density, high voltage, sustainability, and easy fabrication. For such microbatteries, let us recall the energy units used by authors. For a comparison of the volumetric specific energy, one generally refers to that of the cathode material (i.e., the limiting electrode); thus, considering a dense LCO film (d = 4.3 g·cm−3), a gravimetric specific energy of 100 mAh·g−1 is equivalent to 43 µAh·cm−2·µm−1 or 154.8 mC·cm−2·µm−1.
Let us compare and discuss the growth conditions that allow the best electrochemical performance of each component, i.e., cathode, anode, and electrolyte. Regarding the growth of lithiated oxides used as cathode materials, an excess of lithium (at least ~15 wt. %) is mandatory for any material to compensate the loss of volatile lithium species during the ablation process. Given the different materials’ available candidates for the cathode, LiCoO2 appears to be the most electrochemically efficient. From Figure 17, comparing the Ragone plots of LiCoO2 and LiMn2O4 thin films, each curve displays a series of discharge profiles for a lithium microbattery with the cathode thickness (in µm). As the capacity delivered by a cathode is proportional to the mass of the material, thicker films provide high energies, but often at the expense of the power performance [26]. The specific energy for a planar Li//LiCoO2 cell with a thick cathode can reach 500 µWh·cm−2·µm−1. However, the rate capability depends strongly on the plane orientation of the film, which can be controlled by the nature of the substrate. Thus, the preferred orientation is the (003)-plane parallel [29]. However, a surface–electrolyte interface (SEI) is formed on epitaxial LiCoO2 films with different orientations of (104), (110), and (003) that result in anisotropic reactions of intercalation activity [102]. It has also been demonstrated that the minimized strain energy in thick LCO films allows preferential (101) and (104) textures [32]. Nevertheless, the best well-crystallized LCO thin films were fabricated in the following PLD conditions: Ts = 500 °C, PO₂ = 13 Pa, Φ = 2 J·cm−2, substrate–target distance of 30 to 40 mm, and laser beam-target incident angle of 45°. Interesting results reported in Ref. [53] showed impurity-free LCO films, highly (003)-oriented with a very small lattice expansion during charge (at 4.2 V), when grown on stainless steel substrates at relatively low temperatures (Ts ≈ 300 °C). In this case, the films had a texture between the amorphous and well-crystalline state with very small grains, which is suitable for short pathways for electrons and ions during the (de)intercalation reaction.
To avoid the poor performance of LIBs derived from hindered lithium-ion diffusion at the interface between the LCO positive electrode and electrolyte, modifications of the cathode surface have been realized by the deposition of a thin layer of a fast-ionic Li+ conductor, such as amorphous Li2WO4 or Li3PO4. This layer reduces the interfacial Li+-ion transfer resistance that results in a rapid charge–discharge rate. The a-Li2WO3/LCO/Pt/Cr/SiO2 electrode cycled at a high rate of 20C with a high capacity retention [84]. Another electrode exhibiting a fast charge–discharge rate as high as 348C has been fabricated by the multilayer PLD technique, but in this case, the LMO thin film exhibited a significant pseudocapacitive behavior (non-diffusion-controlled) instead of a faradaic mechanism. An additional promising improvement is the fabrication of an LCO thin film sandwiched between a PLD-prepared SrRuO3 film as the electronic conductor and the film of Li3PO4 (3.2 nm thick) as the ionic conductor with the result being limited surface structural change in the high voltage range (4.4 V) [71].
The influence of the PLD conditions on the texture of LiMn2O4 thin films has shown that Ts = 500 °C and PO₂ = 20 Pa are the optimum values that maintain the Li/Mn ratio close to 1, when an Li-enriched target is used, and obtains the best mass transfer [122]. It was also noticed that any substrate does not strongly influence the stoichiometry, but affects the out-of-plane preferred texture. The applicability of the PLD-grown V2O5 films in lithium microbatteries has been evidenced that, in the range of 200 < Ts < 400 °C under PO₂ = 6 Pa, the films offer better electrochemical performance than those grown at other temperatures in terms of their structural quality and stability. Only two works were devoted to the fabrication of solid-state thin-film batteries with vanadium oxide as the cathode materials: The Li/Li1.4B2.5S 0.1O4.9/V2O5 cell delivered a capacity of ~400 mC·cm−2·µm−1 [221], while the Li/LiPON/Ag0.3V2O5 maintained a specific capacity of 40 µAh·cm−2·µm−1 after 100 cycles [222], but the low current density was due to the poor electronic conductivity of the positive electrode. Recently, a solid-state thin-film battery, Li/Li3PO4/LiMnPO4, was successfully fabricated by PLD [48]. Such a cell delivered a modest specific capacity of 10 µAh·cm−2·µm−1, which was limited by the slow chemical diffusion coefficient of the Li+ ion in the olivine framework (3 × 10−17 cm2·s−1).
Lithium phosphates, i.e., Li3PO4 and LiPON, are the most widely used solid electrolytes in microbatteries; they are easily fabricated by PLD using an ArF excimer laser and show a good ionic conductivity. The LiPON electrolyte is known to exhibit a better chemical stability than Li3PO4 [282]. However, the electrochemical stability of PLD-prepared Li3PO4 thin films is greater than 4.7 V [265].
Few works have attempted to replace the lithium metal thin-film anode by other lithiated materials (i.e., intercalation compound or alloy) for the fabrication of microbatteries. The most stable insertion compound should be Li4Ti5O12 spinel with minor volumetric changes but the high voltage plateau of 1.5 V is a great penalty for high energy density. The In/80Li2S–20P2S5/LiCoO2 microbattery developed by Sakuda et al. seems to be promising as the Li-In alloy allows a high specific discharge capacity at moderate current density of 0.13 mA·g−1 [42].

7. Concluding Remarks

The results of the intensive research on the growth of thin films by pulsed laser deposition in recent years have been reviewed. Due to careful investigations of the mechanism of the sample preparation, optimized materials with adequate properties for energy storage and conversion have been obtained. The PLD technique is considered to be suitable for improving the density and adhesion properties of films. A huge effort has been mainly concentrated on the deposition of lithiated oxides, which require specific conditions due to the volatile character of lithium vapor species during the PLD process. Due to the outstanding performance of the conventional cathode materials, LiCoO2 and LiMn2O4, PLD films exhibiting a specific capacity close to the theoretical one are the most popular. The progress concerns mainly the epitaxial films grown with an orientation favorable to a high rate of transport of Li ions at the electrode/electrolyte interface. For instance, the pyramidal-type LiMn2O4 films cycled at the 3.3C rate demonstrate a specific capacity of 90 mAh·g−1 after 1000 cycles.
The PLD technique has proved to also be efficient for the preparation of thin films of anode materials. The best example is the production of LTO, which is a “zero-strain” compound. Other anode thin film materials, such as silicon and conversion-type oxides, are attractive due to their high specific capacity and easy PLD fabrication.
So far, solid-electrolyte thin-films have been fabricated essentially by thermal vacuum evaporation and rf-sputtering. The manufacture of solid-electrolyte thin films by PLD has brought improvements in their intrinsic properties. For example, the electronic conductivity of PLD films is small in comparison with rf-sputtered films. LiPON and Li–V–S–O are the most popular solid-electrolyte films.
In recent years, due to a strong demand for smaller power sources, the interest in rechargeable micro-batteries has gradually increased. The progress on lithium microbatteries is remarkable, mainly due to the PLD growth of high-quality, pinhole-free, solid-state electrolyte thin films, such as Li6.1V0.61Si0.39O5.36. The rechargeable thin-film lithium-ion battery designed by the Japanese group at Tohoku University was fabricated using the sequential PLD technique. This microcell delivered a specific capacity of 9.5 Ah cm−2 discharged at a current density of 44 µA·cm−2 using an Li-Sn alloy film as the anode and showed good reversibility over 100 cycles.

Funding

This research received no external funding.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Chrisey, D.G.; Hubler, G.K. (Eds.) Pulsed Laser Deposition of Thin Films; Wiley: New York, NY, USA, 1994. [Google Scholar]
  2. Bauerle, D. Laser Processing and Chemistry; Springer: Berlin, Germany, 1996. [Google Scholar]
  3. Julien, C.; Camacho-Lopez, M.A.; Escobar-Alarcon, L.; Haro-Poniatowski, E. Fabrication of LiCoO2 thin-film cathodes for rechargeable lithium microbatteries. Mater. Chem. Phys. 2001, 68, 210–216. [Google Scholar] [CrossRef]
  4. Antaya, M.; Dahn, J.R.; Preston, J.S.; Rossen, E.; Reimers, J.N. Preparation and characterization of LiCoO2 thin films by laser ablation deposition. J. Electrochem. Soc. 1993, 140, 575–578. [Google Scholar] [CrossRef]
  5. Tsuruhama, T.; Hitosugi, T.; Oki, H.; Hirose, Y.; Hasegawa, T. Preparation of layered-rhombohedral LiCoO2 epitaxial thin films using pulsed laser deposition. Appl. Phys. Express 2009, 2, 085502. [Google Scholar] [CrossRef]
  6. Perkins, J.D.; Bahn, C.S.; McGraw, J.M.; Parilla, P.A.; Ginley, D.S. Pulsed laser deposition and characterization of crystalline lithium cobalt dioxide  ( LiCoO2 )  thin films. J. Electrochem. Soc. 2001, 148, A1301–A1312. [Google Scholar] [CrossRef]
  7. Julien, C.; Haro-Poniatowski, E.; Camacho-Lopez, M.A.; Escobar-Alarcon, L.; Jimenez-Jarquin, J. Growth of LiMn2O4 thin films by pulsed-laser deposition and their electrochemical properties in lithium microbatteries. Mater. Sci. Eng. B 2000, 72, 36–46. [Google Scholar] [CrossRef]
  8. Julien, C.; Haro-Poniatowski, E.; Hussain, O.M.; Ramana, C.V. Structure and electrochemistry of thin-film oxides grown by laser-pulsed deposition. Ionics 2001, 7, 165–171. [Google Scholar] [CrossRef]
  9. Montenegro, M.J.; Lippert, T. Films for electrochemical applications. In Pulsed Laser Deposition of Thin Films: Applications-Led Growth of Functional Materials; Eason, R., Ed.; John Wiley & Son Inc.: Hoboken, NJ, USA, 2007; Chapter 22; pp. 563–584. [Google Scholar]
  10. Iriyama, Y.; Inaba, M.; Abe, T.; Ogumi, Z. Preparation of c-axis oriented thin films of LiCoO2 by pulsed laser deposition and their electrochemical properties. J. Power Sources 2001, 94, 175–182. [Google Scholar] [CrossRef]
  11. Huo, W.; Li, J.; Chen, G.; Wang, Y.; Zou, W.; Rao, Q.; Zhou, A.; Chang, A.; Wang, Q. LiCoO2 thin film cathode fabricated by pulsed laser deposition. Rare Met. 2011, 30 (Suppl. 1), 106–110. [Google Scholar] [CrossRef]
  12. Xia, H.; Tang, S.B.; Lu, L. Thin film microbatteries prepared by pulsed laser deposition. J. Korean Phys. Soc. 2007, 51, 1055–1062. [Google Scholar] [CrossRef]
  13. Escobar-Alarcon, L.; Haro-Poniatowski, E.; Camacho-Lopez, M.A.; Fernandez-Guasti, M.; Jımenez-Jarquın, J.; Sanchez-Pineda, A. Structural characterization of TiO2 thin films obtained by pulsed laser deposition. Appl. Surf. Sci. 1999, 137, 38–44. [Google Scholar] [CrossRef]
  14. Kuwata, N.; Iwagami, N.; Matsuda, Y.; Tanji, Y.; Kawamura, J. Thin film batteries with Li3PO4 solid electrolytes fabricated by pulsed laser deposition. ECS Trans. 2009, 16, 53–60. [Google Scholar]
  15. Julien, C. Solid State Batteries. In Handbook of Solid State Electrochemistry; Gellings, P.J., Bouwmeester, H.J.M., Eds.; CRC Press: Boca Raton, FL, USA, 1997; pp. 371–406. [Google Scholar]
  16. Julien, C. Thin film technology and microbatteries. In Lithium Batteries-New Materials-Development and Perspectives; Pistoia, G., Ed.; Elsevier: Amsterdam, The Netherlands, 1994; pp. 167–237. [Google Scholar]
  17. Sator, A. Pile réversible dont l’électrolyte est un cristal déposé en lame mince par évaporation. C. R. Acad. Sci. Paris 1952, 234, 2283–2285. [Google Scholar]
  18. Kennedy, J.H. Thin film electrolyte systems. Thin Solid Film 1977, 43, 41–92. [Google Scholar] [CrossRef]
  19. Bates, J.B.; Dudney, N.J.; Gruzalski, G.R.; Luck, C.F. Thin Film Battery and Method for Making Same. U.S. Patent 5,338,625, 16 April 1994. [Google Scholar]
  20. Bates, J.B.; Dudney, N.J.; Gruzalski, G.R.; Zuhr, R.A.; Choudhury, A.; Luck, C.F.; Robertson, J.D. Electrical properties of amorphous lithium electrolyte thin films. Solid State Ion. 1992, 53, 647–654. [Google Scholar] [CrossRef]
  21. Bates, J.B.; Dudney, N.J.; Gruzalski, G.R.; Zuhr, R.A.; Choudhury, A.; Luck, C.F.; Robertson, J.D. Fabrication and characterization of amorphous lithium electrolyte thin films and rechargeable thin-film batteries. J. Power Sources 1993, 43, 103–110. [Google Scholar] [CrossRef]
  22. Bates, J.B.; Dudney, N.J.; Lubben, D.C.; Gruzalski, G.R.; Kwak, B.S.; Yu, X.H.; Zuhr, R.A. Thin-film rechargeable lithium batteries. J. Power Sources 1995, 54, 58–62. [Google Scholar] [CrossRef] [Green Version]
  23. Bates, J.B.; Dudney, N.J.; Luck, C.F.; Sales, B.C.; Zuhr, R.A.; Robertson, J.D. Deposition and characterization of Li2O-SiO2-P2O5 thin films. J. Am. Ceram. Soc. 1993, 76, 929–943. [Google Scholar] [CrossRef]
  24. Bates, J.B.; Gruzalski, G.R.; Dudney, N.J.; Luck, C.F.; Yu, X.H. Rechargeable thin-film lithium batteries. Solid State Ion. 1994, 70, 619–628. [Google Scholar] [CrossRef]
  25. Julien, C. Lithium microbatteries. In Materials for Lithium-Ion Batteries; Julien, C., Stoynov, Z., Eds.; Springer: Dordrecht, The Netherlands, 2000; pp. 381–400. [Google Scholar]
  26. Dudney, N.J. Thin film micro-batteries. Electrochem. Interface 2001, 17, 44–48. [Google Scholar]
  27. Oudenhoven, J.F.M.; Baggetto, L.; Notten, P.H.L. All-solid-state lithium-ion microbatteries: A review of various three-dimensional concepts. Adv. Energy Mater. 2011, 1, 10–33. [Google Scholar] [CrossRef]
  28. Wang, Y.; Liu, B.; Li, Q.; Cartmell, S.; Ferrara, S.; Deng, Z.D.; Xiao, J. Lithium and lithium-ion microbatteries for applications in microelectronic devices: A review. J. Power Sources 2015, 286, 330–345. [Google Scholar] [CrossRef]
  29. Ferrari, S.; Loveridge, M.; Beattie, S.D.; Jahn, M.; Dashwood, R.J.; Bhagat, R. Latest advances in the manufacturing of 3D rechargeable lithium microbatteries. J. Power Sources 2015, 286, 25–46. [Google Scholar] [CrossRef] [Green Version]
  30. Meng, Y.S.; McGilway, T.; Yang, M.-C.; Gostovic, D.; Wang, F.; Zeng, D.; Zhu, Y.; Graetz, J. In situ analytical electron microscopy for probing nanoscale electrochemistry. Electrochem. Interface 2011, 20, 49–53. [Google Scholar] [CrossRef]
  31. Jones, S.D.; Akridge, J.R. Athin film solid state microbattery. Solid State Ion. 1992, 53, 628–634. [Google Scholar] [CrossRef]
  32. Laïk, B.; Ressejac, I.; Venet, C.; Pereira-Ramos, J.P. Comparative study of electrochemical performance of commercial solid-state thin film Li microbatteries. Thin Solid Films 2018, 649, 69–74. [Google Scholar] [CrossRef]
  33. Enfucell Powers the Internet of Things Revolution. Available online: http://www.enfucell.com/softbattery (accessed on 28 September 2018).
  34. EnerchipTM Smart Solid State Batteries. Available online: http://www.cymbet.com/products/enerchip-solid-state-batteries.php (accessed on 1 January 2019).
  35. Thin Film Battery Technology. Available online: http://www.excellatron.com./thin-film-battery-technology (accessed on 1 January 2018).
  36. Thinergy Micro-Energy Cell Product Family. Available online: https://www.batterypoweronline.com/markets/batteries/thinergy-micro-energy-cell-product-family/ (accessed on 1 January 2019).
  37. NanoEnergyTM. Available online: http:/www.frontedgetechnology.com./gen.htm (accessed on 1 January 2014).
  38. Ulvac Enables Flexible Li-Ion Battery. Available online: https://fr.scribd.com/document/118211050/196729-Ulvac-Enables (accessed on 22 December 2008).
  39. Batteries Ultraminces à Durée de vie Etendue de STMicroelectronics Pour L’Alimentation Des Technologies Miniatures de Demain. Available online: https://www.digikey.fr/fr/product-highlight/s/stmicroelectronics/efl700a39-enfilm-rechargeable-battery/ (accessed on 13 January 2016).
  40. Thin Film Rechargeable Batteries from Korea. Available online: https://www.idtechex.com/research/articles/thin-film-rechargeable-batteries-from-korea-00003130.ja.asp?donotredirect=true&setlang=ja (accessed on 23 February 2011).
  41. Kuwata, N.; Kawamura, J.; Toribami, K.; Hattori, T.; Sata, N. Thin-film lithium-ion battery with amorphous solid electrolyte fabricated by pulsed laser deposition. Electrochem. Commun. 2004, 6, 417–421. [Google Scholar] [CrossRef]
  42. Sakuda, A.; Hayashi, A.; Ohtomo, T.; Hama, S.; Tatsumisago, M. All-solid-state lithium secondary batteries using LiCoO2 particles with pulsed laser deposition coatings of Li2S–P2S5 solid electrolytes. J. Power Sources 2011, 196, 6735–6741. [Google Scholar] [CrossRef]
  43. Schichtel, P.; Geiß, M.; Leichtweiß, T.; Sann, J.; Weber, D.A.; Janek, J. On the impedance and phase transition of thin film all-solid-state batteries based on the Li4Ti5O12 system. J. Power Sources 2017, 360, 593–604. [Google Scholar] [CrossRef]
  44. Nakagawa, A.; Kuwata, N.; Matsuda, Y.; Kawamura, J. Thin film lithium battery using stable solid electrolyte Li4SiO4 fabricated by PLD. ECS Trans. 2010, 25, 155–161. [Google Scholar]
  45. Fujimoto, D.; Kuwata, N.; Matsuda, Y.; Kawamura, J.; Kang, F. Fabrication of solid-state thin-film batteries using LiMnPO4 thin films deposited by pulsed laser deposition. Thin Solid Films 2015, 579, 81–88. [Google Scholar] [CrossRef]
  46. Ohzuku, T.; Ueda, A. Solid-state redox reactions of LiCoO2 (R3m) for 4 volt secondary lithium cells. J. Electrochem. Soc. 1994, 141, 2972–2977. [Google Scholar] [CrossRef]
  47. Striebel, K.A.; Deng, C.Z.; Wen, S.J.; Cairns, E.J. Electrochemical behavior of LiMn2O4 and LiCoO2 thin films produced with pulsed laser deposition. J. Electrochem. Soc. 1996, 143, 1821–1827. [Google Scholar] [CrossRef]
  48. Bouwman, P.J.; Boukamp, B.A.; Bouwmeester, J.M.; Notten, P.H.L. Influence of diffusion plane orientation on electrochemical properties of thin film LiCoO2 electrodes. J. Electrochem. Soc. 2002, 149, A699–A709. [Google Scholar] [CrossRef]
  49. McGraw, J.; Bahn, C.; Parilla, P.; Perkins, J.; Readey, D.; Ginley, D. Li ion diffusion measurements in V2O5 and Li(Co1−xAlx)O2 thin-film battery cathodes. Electrochim. Acta 1999, 45, 187–196. [Google Scholar] [CrossRef]
  50. Escobar-Alarcon, L.; Haro-Poniatowski, E.; Massot, M.; Julien, C. Physical properties of lithium-cobalt oxides grown by laser ablation. Mater. Res. Soc. Symp. Proc. 1999, 548, 223–228. [Google Scholar] [CrossRef]
  51. Okada, K.; Ohnishi, T.; Mitsuishi, K.; Takada, K. Epitaxial growth of LiCoO2 thin films with (001) orientation. AIP Adv. 2017, 7, 115011. [Google Scholar] [CrossRef]
  52. Ohnishi, T.; Hang, B.T.; Xu, X.; Osada, M.; Takada, K. Quality control of epitaxial LiCoO2 thin films grown by pulsed laser deposition. J. Mater. Res. 2010, 25, 1886–1889. [Google Scholar] [CrossRef]
  53. Ohnishi, T.; Nishio, K.; Takada, K. Composition controlled LiCoO2 epitaxial thin film growth by pulsed laser deposition. Proc. SPIE 2015, 9364, 93640L. [Google Scholar]
  54. Zhang, Y.; Zhong, Z.Y.; Zhu, M. Effect of deposition condition on structure and morphology of pulsed-laser deposited LiCoO2 films. J. Funct. Mater. 2007, 38, 1971–1974. [Google Scholar]
  55. Bouwman, P.J.; Boukamp, B.A.; Bouwmeester, H.J.M.; Wondergem, H.J.; Notten, P.H.L. Structural analysis of submicrometer LiCoO2 films. J. Electrochem. Soc. 2001, 148, A311–A317. [Google Scholar] [CrossRef]
  56. Tang, S.B.; Lu, L.; Lai, M.O. Characterization of a LiCoO2 thin film cathode grown by pulsed laser deposition. Philos. Mag. 2005, 85, 2831–2842. [Google Scholar] [CrossRef]
  57. Nishio, K.; Ohnishi, T.; Akatsuka, K.; Takada, K. Crystal orientation of epitaxial LiCoO2 films grown on SrTiO3 substrates. J. Power Sources 2014, 247, 687–691. [Google Scholar] [CrossRef]
  58. Rao, M.C.; Muntaz-Begum, S.K. Substrate effect on microstructure of LiCoO2 thin film cathodes. Optoelectron. Adv. Mater. 2012, 6, 508–510. [Google Scholar]
  59. Xia, H.; Lu, L.; Ceder, G. Substrate effect on the microstructure and electrochemical properties of LiCoO2 thin films grown by PLD. J. Alloy Compd. 2006, 417, 304–310. [Google Scholar] [CrossRef]
  60. Nishio, K.; Ohnishi, T.; Mitsuishi, K.; Ohta, N.; Watanabe, K.; Takada, K. Orientation alignment of epitaxial LiCoO2 thin films on vicinal SrTiO3 (100) substrates. J. Power Sources 2016, 325, 306–310. [Google Scholar] [CrossRef]
  61. Liu, D.; Chao, Y.; Wu, S.; Zhong, S.; Lei, M. Morphology control of LiCoO2 thin film prepared by pulsed laser deposition. In Proceedings of the 2013 International Conference on Materials for Renewable Energy and Environment, Chengdu, China, 19–21 August 2014; Volume 2, pp. 529–531. [Google Scholar]
  62. Ohnishi, T.; Takada, K. High-rate growth of high-crystallinity LiCoO2 epitaxial thin films by pulsed laser deposition. Appl. Phys. Express 2012, 5, 055502. [Google Scholar] [CrossRef]
  63. Nishio, K.; Ohnishi, T.; Osada, M.; Ohta, N.; Watanabe, K.; Takada, K. Influences of high deposition rate on LiCoO2 epitaxial films prepared by pulsed laser deposition. Solid State Ion. 2016, 285, 91–95. [Google Scholar] [CrossRef]
  64. Funayama, K.; Nakamura, T.; Kuwata, N.; Kawamura, J.; Kawada, T.; Amezawa, K. Electromotive force measurements of LiCoO2 electrode on a lithium ion-conducting glass ceramics under mechanical stress. Solid State Ion. 2016, 285, 75–78. [Google Scholar] [CrossRef]
  65. Rao, M.C. Structure and surface morphology of LiCoO2 thin film cathodes prepared by pulsed laser deposition. J. Intense Pulsed Lasers Appl. Adv. Phys. 2012, 2, 35–38. [Google Scholar]
  66. Rao, M.C. Optical absorption studies of LiCoO2 thin films grown by pulsed laser deposition. Int. J. Pure Appl. Phys. 2010, 6, 365–370. [Google Scholar]
  67. Rao, M.C.; Hussain, O.M. Optical and electrical properties of laser ablated amorphous LiCoO2 thin film cathodes. In IOP Conference Series: Materials Science and Engineering; IOP Publishing: Bristol, UK, 2009; Volume 2, p. 012037. [Google Scholar]
  68. Bouwman, P.J.; Boukamp, B.A.; Bouwmeester, H.J.M.; Notten, P.H.L. Structure-related intercalation behaviour of LiCoO2 films. Solid State Ion. 2002, 152, 181–188. [Google Scholar] [CrossRef]
  69. Xia, H.; Lunney, J.; Ceder, G. Li diffusion in LiCoO2 thin films prepared by pulsed laser deposition. J. Power Sources 2006, 159, 1422–1427. [Google Scholar] [CrossRef]
  70. Xia, H.; Lu, L. Texture effect on the electrochemical properties of LiCoO2 thin films prepared by PLD. Electrochim. Acta 2007, 52, 7014–7020. [Google Scholar] [CrossRef]
  71. Taminato, S.; Hirayama, M.; Suzuki, K.; Tamura, K.; Minato, T.; Arai, H.; Uchimoto, Y.; Ogumi, Z.; Kanno, R. Lithium intercalation and structural changes at the LiCoO2 surface under high voltage battery operation. J. Power Sources 2016, 307, 599–603. [Google Scholar] [CrossRef]
  72. Vasanthi, R.; Mangani, I.R.; Selladurai, S.; Manoravi, P. Comparison of LiCo0.97Mg0.03O2 PLD film with bulk for the application of lithium micro-batteries. In Proceedings of the 8th Asian Conference on Solid State Ionics: Trends in the New Millennium, Langkawi, Malaysia, 15–19 December 2002; pp. 163–170. [Google Scholar]
  73. Rao, M.C.; Hussain, O.M. Growth and characterization of tetravalent doped LiCoO2 thin film cathodes. Indian J. Eng. Mater. Sci. 2009, 16, 335–340. [Google Scholar]
  74. Rao, M.C.; Hussain, O.M. Synthesis and electrochemical properties of Ti doped LiCoO2 thin film cathodes. J. Alloy. Compd. 2010, 491, 503–506. [Google Scholar] [CrossRef]
  75. Hirayama, M.; Sonoyama, N.; Abe, T.; Minoura, M. Characterization of electrode/electrolyte interface for lithium batteries using in situ synchrotron X-ray reflectometry—A new experimental technique for LiCoO2 model electrode. J. Power Sources 2007, 168, 493–500. [Google Scholar] [CrossRef]
  76. Yamada, I.; Iriyama, Y.; Abe, T.; Ogumi, Z. Lithium-ion transfer on a LixCoO2 thin film electrode prepared by pulsed laser deposition—Effect of orientation. J. Power Sources 2007, 172, 933–937. [Google Scholar] [CrossRef]
  77. Tang, S.B.; Lai, M.O.; Lu, L. Li-ion diffusion in highly (003) oriented LiCoO2 thin film cathode prepared by pulsed laser deposition. J. Alloy. Compd. 2008, 449, 300–303. [Google Scholar] [CrossRef]
  78. Takeuchi, S.; Tan, H.; Bharathi, K.K.; Stafford, G.R.; Shin, J.; Yasui, S.; Takeuchi, I.; Bendersky, L.A. Epitaxial LiCoO2 films as a model system for fundamental electrochemical studies of positive electrodes. ACS Appl. Mater. Interfaces 2015, 7, 7901–7911. [Google Scholar] [CrossRef]
  79. Tan, H.; Takeuchi, S.; Bharathi, K.K.; Takeuchi, I.; Bendersky, L.A. Microscopy study of structural evolution in epitaxial LiCoO2 positive electrode films during electrochemical cycling. ACS Appl. Mater. Interfaces 2016, 8, 6727–6735. [Google Scholar] [CrossRef] [PubMed]
  80. Li, Z.; Yasui, S.; Takeuchi, S.; Creuziger, A.; Maruyama, S.; Herzing, A.A.; Takeuchi, I.; Bendersky, L.A. Structural study of epitaxial LiCoO2 films grown by pulsed laser deposition on single crystal SrTiO3 substrates. Thin Solid Films 2016, 612, 472–482. [Google Scholar] [CrossRef]
  81. Hussain, O.M.; Rao, M.C.; Julien, C.M. Growth of nano-structured TiO2-doped LiCoO2 PLD films and their electrochemical properties. In New Trends in Intercalation Compounds for Energy Storage and Conversion; Zaghib, K., Julien, C.M., Prakash, J., Eds.; The Electrochemical Society: Pennington, NJ, USA, 2003; pp. 627–631. [Google Scholar]
  82. Kato, M.; Hayashi, T.; Hasegawa, G.; Lu, X.; Miyazaki, T.; Matsuda, Y.; Kuwata, N.; Koji Kurihara, K.; Kawamura, J. Electrochemical properties of LiCoO2 thin film surface modified by lithium tantalate and lithium niobate coatings. Solid State Ion. 2017, 308, 54–60. [Google Scholar] [CrossRef]
  83. Hayashi, T.; Okada, J.; Toda, E.; Kuzuo, R.; Matsuda, Y.; Kuwata, N.; Kawamura, J. Electrochemical effect of lithium tungsten oxide modification on LiCoO2 thin film electrode. J. Power Sources 2015, 285, 559–567. [Google Scholar] [CrossRef]
  84. Hayashi, T.; Matsuda, Y.; Kuwata, N.; Kawamura, J. High-power durability of LiCoO2 thin film electrode modified with amorphous lithium tungsten oxide. J. Power Sources 2017, 354, 41–47. [Google Scholar] [CrossRef]
  85. Hayashi, T.; Toda, E.; Kuzuo, R.; Matsuda, Y.; Kuwata, N.; Junichi Kawamura, J. Contribution of randomly oriented Li2WO4 with tetragonal symmetry to Li+ ion transfer resistance reduction in lithium ion batteries. Int. J. Electrochem. Sci. 2015, 10, 8150–8157. [Google Scholar]
  86. Hayashi, T.; Miyazaki, T.; Matsuda, Y.; Kuwata, N.; Saruwatari, M.; Furuichi, Y.; Kurihara, K.; Kuzuo, R.; Kawamura, J. Effect of lithium-ion diffusibility on interfacial resistance of LiCoO2 thin film electrode modified with lithium tungsten oxides. J. Power Sources 2016, 305, 46–53. [Google Scholar] [CrossRef]
  87. Kuwata, N.; Ise, K.; Matsuda, Y.; Kawamura, J.; Tsurui, T.; Kamishima, O. Detection of degradation in LiCoO2 thin films by in situ micro Raman microscopy. In Ionics for Sustainable World, Proceedings of the 13th Asian Conference on Solid State Ionics, Sendai, Japan, 17–20 July 2012; Chowdari, B.V.R., Kawamura, J., Mizusaki, J., Amezawa, K., Eds.; World Scientific Publishing Co.: Singapore, 2013; pp. 138–143. [Google Scholar]
  88. Matsuda, Y.; Kuwata, N.; Kawamura, J. Thin-film lithium batteries with 0.3–30 μm thick LiCoO2 films fabricated by high-rate pulsed laser deposition. Solid State Ion. 2014, 320, 38–44. [Google Scholar] [CrossRef]
  89. Shiraki, S.; Oki, H.; Takagi, Y.; Suzuki, T.; Kumatani, A.; Shimizu, R.; Haruta, M.; Ohsawa, T.; Sato, Y.; Ikuhara, Y.; et al. Fabrication of all-solid-state battery using epitaxial LiCoO2 thin films. J. Power Sources 2014, 267, 881–887. [Google Scholar] [CrossRef]
  90. Wen, S.-J.; von Rottkay, K.; Rubin, M. Electrochromic lithium nickel oxide thin film by pulsed laser deposition. ECS Symp. Proc. 1997, 96, 54–63. [Google Scholar]
  91. Bouessay, I.; Rougier, A.; Tarascon, J.M. Electrochromic mechanism in nickel oxide thin films grown by pulsed laser deposition. Electrochem. Soc. Proc. 2003, 2003, 91–102. [Google Scholar]
  92. Jahromi, S.P.; Huang, N.M.; Kamalianfar, A.; Lim, H.N.; Muhamad, M.R.; Yousef, R. Facile synthesis of porous-structured nickel oxide thin film by pulsed laser deposition. J. Nanomater. 2012, 2012, 173825. [Google Scholar]
  93. Wang, H.; Wang, Y.; Wang, X. Pulsed laser deposition of the porous nickel oxide thin film at room temperature for high-rate pseudocapacitive energy storage. Electrochem. Commun. 2012, 18, 92–95. [Google Scholar] [CrossRef]
  94. López-Iturbe, J.; Camacho-López, M.A.; Escobar-Alarcón, L.; Camps, E. Synthesis and characterization of LiNiO2 targets for thin film deposition by pulsed laser ablation. Superficies Y Vacío 2005, 18, 27–30. [Google Scholar]
  95. Rao, M.C.; Ravindranadh, K.; Srikanth, K.S. Microstructural and electrochemical studies on laser ablated LiNiO2 thin films. J. Chem. Pharm. Res. 2016, 8, 677–684. [Google Scholar]
  96. Yuki, Y.; Uchiyama, T.; Yamamoto, K.; Uchimoto, Y. Preparation of Li1−xNi1+xO2 thin films by pulsed laser deposition and the electrochemical performance for oxygen evolution reaction in alkaline media. In ESC Meeting Abstracts; The Electrochemical Society: Pennington, NJ, USA, 2018; p. 2585. [Google Scholar]
  97. Razeg, K.H.; Al-Hilli, M.F.; Khalefa, A.A.; Aadim, K.A. Structural and optical properties of (LixNi2−xO2) thin films deposited by pulsed laser deposited (PLD) technique at different doping ratio. Int. J. Phys. 2017, 5, 46–52. [Google Scholar]
  98. Julien, C.; El-Farh, L.; Rangan, S.; Massot, M. studies of LiNi0.8Co0.2O2 cathode material prepared by the citric acid-assisted sol-gel method for lithium batteries. J. Sol-Gel Sci. Technol. 1999, 15, 63–72. [Google Scholar] [CrossRef]
  99. Ramana, C.V.; Zaghib, K.; Julien, C.M. Synthesis, structural and electrochemical properties of pulsed laser deposited Li(Ni,Co)O2 films. J. Power Sources 2006, 159, 1310–1315. [Google Scholar] [CrossRef]
  100. Ramana, C.V.; Zaghib, K.; Julien, C.M. Highly oriented growth of pulsed-laser deposited LiNi0.8Co0.2O2 films for application in microbatteries. Chem. Mater. 2006, 18, 1397–1400. [Google Scholar] [CrossRef]
  101. Ramana, C.V.; Zaghib, K.; Julien, C.M. Growth and electrochemical properties of pulsed laser deposited Al-doped Li-Ni-Co oxide films. In Proceedings of the Extended Abstracts of the 231st ACS National Meeting, Atlanta, GA, USA, 26–30 March 2006; p. 944631. [Google Scholar]
  102. Hirayama, M.; Sakamoto, K.; Hiraide, T.; Mori, D.; Yamada, A.; Kanno, R.; Sonoyama, N.; Tamura, K.; Mizuki, J. Characterization of electrode/electrolyte interface using in situ X-ray reflectometry and LiNi0.8Co0.2O2 epitaxial film electrode synthesized by pulsed laser deposition method. Electrochim. Acta 2007, 53, 871–881. [Google Scholar] [CrossRef]
  103. Wang, G.X.; Lindsay, M.J.; Ionescu, M.; Bradhurst, D.H.; Dou, S.X.; Liu, H.K. Physical and electrochemical characterization of LiNi0.8Co0.2O2 thin-film electrodes deposited by laser ablation. J. Power Sources 2001, 97, 298–302. [Google Scholar] [CrossRef]
  104. Imanishi, N.; Shizuka, K.; Matsumura, T.; Hirano, A.; Takeda, Y.; Kanno, R. Impedance analysis of PLD LiNi0.8Co0.2O2 film electrode. J. Power Sources 2007, 174, 751–755. [Google Scholar] [CrossRef]
  105. Baskaran, R.; Kuwata, N.; Kamishima, O.; Kawamura, J.; Selvasekarapandian, S. Structural and electrochemical studies on thin film LiNi0.8Co0.2O2 by PLD for micro battery. Solid State Ion. 2009, 180, 636–643. [Google Scholar] [CrossRef]
  106. Xia, H.; Lu, L.; Meng, Y.S. Growth of layered LiNi0.5Mn0.5O2 thin films by pulsed laser deposition for application in microbatteries. Appl. Phys. Lett. 2008, 92, 011912. [Google Scholar] [CrossRef]
  107. Xia, H.; Lu, L.; Lai, M.O. Li diffusion in LiNi0.5Mn0.5O2 thin film electrodes prepared by pulsed laser deposition. Electrochim. Acta 2009, 54, 5986–5991. [Google Scholar] [CrossRef]
  108. Sakamoto, K.; Konishi, H.; Sonoyama, N.; Yamada, A.; Tamura, K.; Mizuki, J.; Kanno, R. Mechanistic study on lithium intercalation using a restricted reaction field in LiNi0.5Mn0.5O2. J. Power Sources 2007, 174, 678–682. [Google Scholar] [CrossRef]
  109. Sakamoto, K.; Hirayama, M.; Konishi, H.; Sonoyama, N.; Dupre, N.; Guyomard, D.; Tamura, K.; Mizuki, J.; Kanno, R. Structural changes in surface and bulk LiNi0.5Mn0.5O2 during electrochemical reaction on epitaxial thin-film electrodes characterized by in situ X-ray scattering. Phys. Chem. Chem. Phys. 2010, 12, 3815–3823. [Google Scholar] [CrossRef]
  110. Ramana, C.V.; Zaghib, K.; Julien, C.M. Pulsed-laser deposited LiNi0.8Co0.15Al0.05O2 thin films for application in microbatteries. Appl. Phys. Lett. 2007, 90, 21916. [Google Scholar] [CrossRef]
  111. Deng, J.; Xi, L.; Wang, L.; Wang, Z.; Chung, C.Y.; Han, X.; Zhou, H. Electrochemical performance of LiNi1/3Co1/3Mn1/3O2 thin film electrodes prepared by pulsed laser deposition. J. Power Sources 2012, 217, 491–497. [Google Scholar] [CrossRef]
  112. Jacob, C.; Lynch, T.; Chen, A.; Jian, J.; Wang, H. Highly textured Li(Ni0.5Mn0.3Co0.2)O2 thin films on stainless steel as cathode for lithium-ion battery. J. Power Sources 2013, 241, 410–414. [Google Scholar] [CrossRef]
  113. Smyrek, P.; Kim, H.; Zheng, Y.; Seifert, H.J.; Piqué, A.; Pfleging, W. Laser-printing and femtosecond laser-structuring of electrode materials for the manufacturing of 3D lithium-ion micro-batteries. Proc. SPIE 2016, 9738, 973806. [Google Scholar]
  114. Abe, M.; Iba, H.; Suzuki, K.; Minamishima, H.; Hirayama, M.; Tamura, K.; Mizuki, J.; Saito, T.; Ikuhara, Y.; Kanno, R. Study on the deterioration mechanism of layered rock-salt electrodes using epitaxial thin films: Li(Ni, Co, Mn)O2 and their Zr-O surface modified electrodes. J. Power Sources 2017, 345, 108–119. [Google Scholar] [CrossRef]
  115. Bendersky, L.A.; Tan, H.; Karuppanan, K.B.; Li, Z.P.; Johnston-Peck, A.C. Crystallography and growth of epitaxial oxide films for fundamental studies of cathode materials used in advanced Li-ion batteries. Crystals 2017, 7, 127. [Google Scholar] [CrossRef]
  116. Johnston-Peck, A.C.; Takeuchi, S.; Kamala-Bharathi, K.; Herzing, A.A.; Bendersky, L.A. Domain formation in lithium-rich manganese-nickel-cobalt-oxide epitaxial thin films and implications for interpretation of electrochemical behavior. Thin Solid Films 2018, 647, 40–49. [Google Scholar] [CrossRef]
  117. Yan, B.; Liu, J.; Song, B.; Xiao, P.; Lu, L. Li-rich thin film cathode prepared by pulsed laser deposition. Sci. Rep. 2013, 3, 3332. [Google Scholar] [CrossRef] [PubMed]
  118. Taminato, S.; Hirayama, M.; Suzuki, K.; Yamada, N.L.; Yonemura, M.; Son, J.Y.; Kanno, R. Highly reversible capacity at the surface of a lithium-rich manganese oxide: A model study using an epitaxial film system. Chem. Commun. 2015, 51, 1673–1676. [Google Scholar] [CrossRef] [PubMed]
  119. Zheng, Y.; Taminato, S.; Suzuki, K.; Hirayama, M.; Kanno, R. Fabrication and lithium intercalation properties of epitaxial Li2RuO3 thin films. Thin Solid Films 2012, 520, 4889–4893. [Google Scholar] [CrossRef]
  120. Taminato, S.; Hirayama, M.; Suzuki, K.; Kim, K.; Zheng, Y.; Tamura, K.; Mizuki, J.I.; Kanno, R. Mechanistic studies on lithium intercalation in a lithium-rich layered material using Li2RuO3 epitaxial film electrodes and in situ surface X-ray analysis. J. Mater. Chem. A 2014, 2, 17875–17882. [Google Scholar] [CrossRef]
  121. Kuwata, N.; Kumar, R.; Toribami, K.; Suzuki, T.; Hattori, T.; Kawamura, J. Thin film lithium ion batteries prepared only pulsed laser deposition. Solid State Ion. 2006, 177, 2827–2832. [Google Scholar] [CrossRef]
  122. Morcrette, M.; Barboux, P.; Perrière, J.; Brousse, T. LiMn2O4 thin films for lithium ion sensors. Solid State Ion. 1998, 112, 249–254. [Google Scholar] [CrossRef]
  123. Camacho-Lopez, M.A.; Escobar-Alarcon, L.; Haro-Poniatowski, E.; Julien, C. LiMn2O4 films grown by pulsed-laser deposition. Ionics 1999, 5, 244–250. [Google Scholar] [CrossRef]
  124. Singh, D.; Houriet, R.; Vacassy, R.; Hofinann, H.; Craciun, V.; Singh, R.K. Pulsed laser deposition and characterization of LiMn2O4 thin films for applications in Li-ion rechargeable battery systems. Mater. Res. Soc. Symp. Proc. 1999, 575, 83–89. [Google Scholar] [CrossRef]
  125. Camacho-Lopez, M.A.; Escobar-Alarcon, L.; Haro-Poniatowski, E.; Julien, C. Physico-chemical properties of LiMn2O4 films grown by laser ablation. In Materials for Lithium-Ion Batteries; Julien, C., Stoynov, Z., Eds.; NATO Science Series; Kluwer Academic Publishers: Dordrecht, The Netherlands, 2000; Volume 3–85, pp. 535–541. [Google Scholar]
  126. Camacho-Lopez, M.A.; Escobar-Alarcon, L.; Haro-Poniatowski, E.; Julien, C. Raman studies of LiMn2O4 films grown by laser ablation. Mater. Res. Soc. Symp. Proc. 2000, 617. [Google Scholar] [CrossRef]
  127. Simmen, F.; Lippert, T.; Novák, P.; Neuenschwander, B.; Döbeli, M.; Mallepell, M.; Wokaun, A. The influence of lithium excess in the target on the properties and compositions of Li1+xMn2O4−δ thin films prepared by PLD. Appl. Phys. A 2008, 93, 711–716. [Google Scholar] [CrossRef]
  128. Gao, X.; Ikuhara, Y.H.; Fisher, C.A.; Moriwake, H.; Kuwabara, A.; Oki, H.; Kohama, K.; Yoshida, R.; Huang, R.; Ikuhara, Y. Structural distortion and compositional gradients adjacent to epitaxial LiMn2O4 thin film interfaces. Adv. Mater. Interfaces 2014, 1, 1400143. [Google Scholar] [CrossRef]
  129. Hendriks, R.; Monteiro Cunha, D.; Singh, D.P.; Huijben, M. Enhanced lithium transport by control of crystal orientation in spinel LiMn2O4 thin film cathodes. ACS Appl. Energy Mater. 2018, 1, 7046–7051. [Google Scholar] [CrossRef] [PubMed]
  130. Sonoyama, N.; Iwase, K.; Takatsuka, H.; Matsumura, T.; Imanishi, N.; Takeda, Y.; Kanno, R. Electrochemistry of LiMn2O4 epitaxial films deposited on various single crystal substrates. J. Power Sources 2009, 189, 561–565. [Google Scholar] [CrossRef]
  131. Canulescu, S.; Papadopoulou, E.L.; Anglos, D.; Lippert, T.; Schneider, C.W.; Wokaun, A. Mechanisms of the laser plume expansion during the ablation of LiMn2O4. J. Appl. Phys. 2009, 105, 063107. [Google Scholar] [CrossRef]
  132. Hussain, O.M.; Hari Krishna, K.; Kalai-Vani, V.; Julien, C.M. Pulsed laser deposition of highly oriented LiCoO2 and LiMn2O4 thin films for microbattery applications. In Proceedings of the 10th Asian Conference on Solid State Ionics, Kandy, Sri Lanka, 12–16 June 2006; pp. 136–145. [Google Scholar]
  133. Hussain, O.M.; Hari-Krishna, K.; Kalai-Vani, V.; Julien, C.M. Structural and electrical properties of lithium manganese oxide thin films grown by pulsed laser deposition. Ionics 2007, 13, 455–459. [Google Scholar] [CrossRef]
  134. Tang, S.B.; Lai, M.O.; Lu, L.; Tripathy, S. Comparative study of LiMn2O4 thin film cathode grown at high medium and low temperatures by pulsed laser deposition. J. Solid State Chem. 2006, 179, 3831–3838. [Google Scholar] [CrossRef]
  135. Tang, S.B.; Xia, H.; Lai, M.O.; Lu, L. Characterization of LiMn2O4 thin films grown on Si substrates by pulsed laser deposition. J. Alloy. Compd. 2008, 449, 322–325. [Google Scholar] [CrossRef]
  136. Morcrette, M.; Barboux, P.; Perrière, J.; Brousse, T.; Traverse, A.; Boilot, J.P. Non-stoichiometry in LiMn2O4 thin films by laser ablation. Solid State Ion. 2001, 138, 213–219. [Google Scholar] [CrossRef]
  137. Yamada, I.; Abe, T.; Iriyama, Y.; Ogumi, Z. Lithium-ion transfer at LiMn2O4 thin film electrode prepared by pulsed laser deposition. Electrochem. Commun. 2003, 5, 502–505. [Google Scholar] [CrossRef]
  138. Albrecht, D.; Wulfmeier, H.; Fritze, H. Preparation and characterization of c-LiMn2O4 thin films prepared by pulsed laser deposition for lithium-ion batteries. Energy Technol. 2016, 4, 1558–1564. [Google Scholar] [CrossRef]
  139. Yu, X.; Chen, X.; Buchholz, D.B.; Li, Q.; Wu, J.; Fenter, P.A.; Bedzyk, M.J.; Dravid, V.P.; Barnett, S.A. Pulsed laser deposition and characterization of heteroepitaxial LiMn2O4/La0.5Sr0.5CoO3 bilayer thin films as model lithium Ion battery cathodes. ACS Appl. Nano Mater. 2018, 1, 642–653. [Google Scholar] [CrossRef]
  140. Singh, D.; Houriet, R.; Giovannini, R.; Hofmann, H.; Cruciun, V.; Singh, R.K. Challenges in making of thin films for LixMnyO4 rechargeable lithium batteries for MEMS. J. Power Sources 2001, 97, 826–831. [Google Scholar] [CrossRef]
  141. Rao, M.C. Thermal properties of laser ablated LiMn2O4 thin films. J. Chem. Biol. Phys. Sci. C 2014, 4, 2555–2559. [Google Scholar]
  142. Rougier, A.; Striebel, K.A.; Wen, S.J.; Cairns, E.J. Cyclic voltammetry of pulsed laser deposited LixMn2O4 thin films. J. Electrochem. Soc. 1998, 145, 2975–2980. [Google Scholar] [CrossRef]
  143. Singh, D.; Singh, R.K. Lithium-Based Rechargeable Batteries. U.S. Patent 20030186128A1, 2 October 2003. [Google Scholar]
  144. Rao, M.C.; Ravindranadh, K.; Ramesh, K.K.D.; Srikanth, K.S. Morphological and electrochemical properties of LiMn2O4 thin films. Der Pharma Chemica 2016, 8, 74–79. [Google Scholar]
  145. Suzuki, K.; Kim, K.S.; Taminato, S.; Komo, M.; Hagiwara, A.; Son, J.Y.; Inami, T.; Konishi, H.; Tamura, K.; Mizuki, J.; et al. Modification effects on structural changes of LiMn2O4 electrode during the electrochemical process. In Proceedings of the Extended Abstracts of the 222nd Electrochem. Soc. Meeting, Honolulu, HI, USA, 7–12 October 2012; p. 216. [Google Scholar]
  146. Striebel, K.; Sakai, E.; Cairns, E. Impedance behavior of the LiMn2O4/LiPF6-DMC-EC interface during cycling. In Workshop on Interfaces, Phenomena, and Nanostructures in Lithium Batteries; Kingler, R., Landgrebe, A., Maupin, P., Eds.; The Electrochemical Society: Pennington, NJ, USA, 2000; Volume 2000–36, pp. 61–67. [Google Scholar]
  147. Hirayama, M. Interfacial reactions of intercalation electrodes in lithium ion batteries. Electrochemistry 2015, 83, 701–706. [Google Scholar] [CrossRef]
  148. Hirayama, M.; Ido, H.; Kim, K.S.; Cho, W.; Tamura, K.; Mizuki, J.; Kanno, R. Dynamic structural changes at LiMn2O4/electrolyte interface during lithium battery reaction. J. Am. Chem. Soc. 2010, 132, 15268–15276. [Google Scholar] [CrossRef] [PubMed]
  149. Inaba, M.; Doi, T.; Iriyama, Y.; Abe, T.; Ogumi, Z. Electrochemical STM observation of LiMn2O4 thin films prepared by pulsed laser deposition. J. Power Sources 1999, 81, 554–557. [Google Scholar] [CrossRef]
  150. Tang, S.B.; Lai, M.O.; Lu, L. Electrochemical studies of low-temperature processed nano-crystalline LiMn2O4 thin film cathode at 55 °C. J. Power Sources 2007, 164, 372–378. [Google Scholar] [CrossRef]
  151. Xie, J.; Kohno, K.; Matsumura, T.; Imanishi, N.; Hirano, A.; Takeda, Y.; Yamamoto, O. Li-ion diffusion kinetics in LiMn2O4 thin films prepared by pulsed laser deposition. Electrochim. Acta 2008, 54, 376–381. [Google Scholar] [CrossRef]
  152. Tang, S.B.; Lai, M.O.; Lu, L. Study on Li+-ion diffusion in nano-crystalline LiMn2O4 thin film cathode grown by pulsed laser deposition using CV, EIS, PITT techniques. Mater. Chem. Phys. 2008, 111, 149–153. [Google Scholar] [CrossRef]
  153. Ishiyama, H.; Jeong, S.C.; Watanabe, Y.X.; Hirayama, Y.; Imai, N.; Jung, H.S.; Miyatake, H.; Oyaizu, M.; Osa, A.; Otokawa, Y.; et al. Direct measurement of nanoscale lithium diffusion in solid battery materials using radioactive tracer of 8Li. Nucl. Instrum. Methods Phys. Res. B 2016, 376, 379–381. [Google Scholar] [CrossRef]
  154. Kuwata, N.; Nakane, M.; Miyazaki, T.; Mitsuishi, K.; Kawamura, J. Lithium diffusion coefficient in LiMn2O4 thin films measured by secondary ion mass spectrometry with ion-exchange method. Solid State Ion. 2018, 320, 266–271. [Google Scholar] [CrossRef]
  155. Marchini, F.; Rubi, D.; del Pozo, M.; Federcio, W.J.; Calvo, E.J. Surface chemistry and lithium-ion exchange in LiMn2O4 for the electrochemical selective extraction of LiCl from natural salt lake brines. J. Phys. Chem. C 2016, 120, 15875–15883. [Google Scholar] [CrossRef]
  156. Matsumura, T.; Imanishi, N.; Hirano, A.; Sonoyama, N.; Takeda, Y. Electrochemical performances for preferred oriented PLD thin-film electrodes of LiNi0.8Co0.2O2, LiFePO4 and LiMn2O4. Solid State Ion. 2008, 179, 2011–2015. [Google Scholar] [CrossRef]
  157. Tang, S.B.; Lai, M.O.; Lu, L. Properties of nano-crystalline LiMn2O4 thin films deposited by pulsed laser deposition. Electrochim. Acta 2006, 52, 1161–1168. [Google Scholar] [CrossRef]
  158. Hirayama, M.; Sonoyama, N.; Ito, M.; Minoura, M.; Mori, D.; Yamada, A.; Tamura, K.; Mizuki, J.; Kanno, R. Characterization of electrode electrolyte interface with X-ray reflectometry and epitaxial-film LiMn2O4 electrode. J. Electrochem. Soc. 2007, 154, A1065–A1072. [Google Scholar] [CrossRef]
  159. Shin, D.W.; Choi, J.W.; Ahn, J.P.; Choi, W.K.; Cho, Y.S.; Yoon, S.J. ZrO2-modified LiMn2O4 thin-film cathodes prepared by pulsed laser deposition. J. Electrochem. Soc. 2010, 157, A567–A570. [Google Scholar] [CrossRef]
  160. Fehse, M.; Trocoli, R.; Hernandez, E.; Ventosa, E.; Sepulveda, A.; Morata, A.; Tarancon, A. An innovative multi-layer pulsed laser deposition approach for LiMn2O4 thin film cathodes. Thin Solid Films 2018, 648, 108–112. [Google Scholar] [CrossRef]
  161. Kim, S.; Hirayama, M.; Suzuki, K.; Kanno, R. Hetero-epitaxial growth of Li0.17La0.61TiO3 solid electrolyte on LiMn2O4 electrode for all solid-state batteries. Solid State Ion. 2014, 262, 578–581. [Google Scholar] [CrossRef]
  162. Suzuki, K.; Kim, K.; Taminato, S.; Hirayama, M.; Kanno, R. Fabrication and electrochemical properties of LiMn2O4/SrRuO3 multi-layer epitaxial thin film electrodes. J. Power Sources 2013, 226, 340–345. [Google Scholar] [CrossRef]
  163. Yim, H.; Shin, D.W.; Choi, J.W. LiMn2O4-based cathode thin films for Li thin-film batteries. J. Korean Phys. Soc. 2016, 68, 41–53. [Google Scholar] [CrossRef]
  164. Fehse, M.; Trocoli, R.; Hernandez, E.; Ventosa, E.; Hernandez, E.; Sepulveda, A.; Morata, A.; Tarancon, A. Ultrafast dischargeable LiMn2O4 thin-film electrodes with pseudocapacitive properties for microbatteries. ACS Appl. Mater. Interfaces 2017, 9, 5295–5301. [Google Scholar] [CrossRef] [PubMed]
  165. Striebel, K.A.; Rougier, A.; Horne, C.R.; Reade, R.P.; Cairns, E.J. Electrochemical studies of substituted spinel thin films. J. Electrochem. Soc. 1999, 146, 4339–4347. [Google Scholar] [CrossRef]
  166. Xia, H.; Tang, S.B.; Lu, L.; Meng, Y.S.; Ceder, G. The influence of preparation conditions on electrochemical properties of LiNi0.5Mn1.5O4 thin film electrodes by PLD. Electrochim. Acta 2007, 52, 1120–1126. [Google Scholar] [CrossRef]
  167. Xia, H.; Meng, Y.S.; Lu, L.; Ceder, G. Electrochemical properties of nonstoichiometric LiNi0.5Mn1.5O4−δ thin-film electrodes prepared by pulsed laser deposition. J. Electrochem. Soc. 2007, 154, A737–A743. [Google Scholar] [CrossRef]
  168. Xia, H.; Lu, L. Li diffusion in spinel LiNi0.5Mn1.5O4 thin films prepared by pulsed laser deposition. Phys. Scr. 2007, 2007, 43–48. [Google Scholar] [CrossRef]
  169. Konishi, H.; Suzuki, K.; Taminato, S.; Kim, K.; Kim, S.; Lim, J.; Hirayama, M.; Kanno, R. Structure and electrochemical properties of LiNi0.5Mn1.5O4 epitaxial thin film electrodes. J. Power Sources 2014, 246, 365–370. [Google Scholar] [CrossRef]
  170. Liu, D.; Zhu, W.; Trottier, J.; Gagnon, C.; Barray, F.; Gariépy, V.; Guerfi, A.; Mauger, A.; Groult, H.; Julien, C.M.; et al. Spinel materials for high-voltage cathodes in Li-ion batteries. RSC Adv. 2014, 4, 154–167. [Google Scholar] [CrossRef]
  171. Kuwata, N.; Kudo, S.; Matsuda, Y.; Kawamura, J. Fabrication of thin-film lithium batteries with 5-V-class LiCoMnO4 cathodes. Solid State Ion. 2014, 262, 165–169. [Google Scholar] [CrossRef]
  172. Lim, J.; Lee, S.; Suzuki, K.; Kim, K.; Kim, S.; Taminato, S.; Hirayama, M.; Oshima, Y.; Takayanagi, K.; Kanno, R. Synthesis, structure and electrochemical properties of novel Li–Co–Mn–O epitaxial thin-film electrode using layer-by-layer deposition process. J. Power Sources 2015, 279, 502–509. [Google Scholar] [CrossRef]
  173. Yang, D. Pulsed laser deposition of manganese oxide thin films for supercapacitor applications. J. Power Sources 2011, 196, 8843–8849. [Google Scholar] [CrossRef]
  174. Xia, H.; Lai, M.O.; Lu, L. Nanoporous MnOx thin-film electrodes synthesized by electrochemical lithiation/delithiation for supercapacitors. J. Power Sources 2011, 196, 2398–2402. [Google Scholar] [CrossRef]
  175. Yang, D. Pulsed laser deposition of cobalt-doped manganese oxide thin films for supercapacitor applications. J. Power Sources 2012, 198, 416–422. [Google Scholar] [CrossRef]
  176. Yang, D. Pulsed laser deposition of vanadium-doped manganese oxide thin films for supercapacitor applications. J. Power Sources 2013, 228, 89–96. [Google Scholar] [CrossRef]
  177. Lu, Z.G.; Cheng, H.; Lo, M.F.; Chung, C.Y. Pulsed laser deposition and electrochemical characterization of LiFePO4–Ag composite thin films. Adv. Funct. Mater. 2007, 17, 3885–3896. [Google Scholar] [CrossRef]
  178. Lu, Z.G.; Lo, M.F.; Chung, C.Y. Pulse laser deposition and electrochemical characterization of LiFePO4–C composite thin films. J. Phys. Chem. C 2008, 112, 7069–7078. [Google Scholar] [CrossRef]
  179. Tang, K.; Yu, X.; Sun, J.; Li, H.; Huang, X. Kinetic analysis on LiFePO4 thin films by CV, GITT, and EIS. Electrochim. Acta 2011, 56, 4869–4875. [Google Scholar] [CrossRef]
  180. Iriyama, Y.; Yokoyama, M.; Yada, C.; Jeong, S.K.; Yamada, I.; Abe, T.; Inaba, M.; Ogumi, Z. Preparation of LiFePO4 thin films by pulsed laser deposition and their electrochemical properties. Electrochem. Solid State Lett. 2004, 7, A340–A342. [Google Scholar] [CrossRef]
  181. Yada, C.; Iriyama, Y.; Jeong, S.K.; Abe, T.; Inaba, M.; Ogumi, Z. Electrochemical properties of LiFePO4 thin films prepared by pulsed laser deposition. J. Power Sources 2005, 146, 559–564. [Google Scholar] [CrossRef]
  182. Song, S.W.; Reade, R.P.; Kostecki, R.; Striebel, K.A. Electrochemical studies of the LiFePO4 thin films prepared with pulsed laser deposition. J. Electrochem. Soc. 2006, 153, A12–A19. [Google Scholar] [CrossRef]
  183. Sun, J.; Tang, K.; Yu, X.; Li, H.; Huang, X. Needle-like LiFePO4 thin films prepared by an off-axis pulsed laser deposition technique. Thin Solid Films 2009, 517, 2618–2622. [Google Scholar] [CrossRef]
  184. Palomares, V.; Ruiz de Larramendi, I.; Alonso, J.; Bengoechea, M.; Goñi, A.; Miguel, O.; Rojo, T. LiFePO4 thin films grown by pulsed laser deposition: Effect of the substrate on the film structure and morphology. Appl. Surf. Sci. 2010, 256, 2563–2568. [Google Scholar] [CrossRef]
  185. Tang, K.; Sun, J.; Yu, X.; Li, H.; Huang, X. Electrochemical performance of LiFePO4 thin films with different morphology and crystallinity. Electrochim. Acta 2009, 54, 6565–6569. [Google Scholar] [CrossRef]
  186. Xue, M.Z.; Fu, Z.W. Electrochemical properties of LiFePO4 cathode thin film fabricated by pulsed laser deposition. Acta Phys. Chem. Sin. 2005, 21, 707–710. [Google Scholar]
  187. Sauvage, F.; Baudrin, E.; Morcrette, M.; Tarascon, J.M. Pulsed laser deposition and electrochemical properties of LiFePO4 thin films. Electrochem. Solid-State Lett. 2004, 7, A15–A18. [Google Scholar] [CrossRef]
  188. Sauvage, F.; Baudrin, E.; Gengembre, L.; Tarascon, J.M. Effect of texture on the electrochemical properties of LiFePO4 thin films. Solid State Ion. 2005, 176, 1869–1876. [Google Scholar] [CrossRef]
  189. Sauvage, F.; Baudrin, E.; Laffont, L.; Tarascon, J.M. Origin of electrochemical reactivity enhancement of post-annealed LiFePO4 thin films: Preparation of heterosite-type FePO4. Solid State Ion. 2007, 178, 145–152. [Google Scholar] [CrossRef]
  190. Sauvage, F.; Laffont, L.; Tarascon, J.M.; Baudrin, E. Factors affecting the electrochemical reactivity vs. lithium of carbon-free LiFePO4 thin films. J. Power Sources 2008, 175, 495–501. [Google Scholar] [CrossRef]
  191. Raveendran, N.R.; Kumary, T.G.; Rajkumari, S.; Pandian, R.; Janaki, J.; Mani, A. Synthesis and study of FeSe thin films and LiFeO2/FeSe bi-layers. AIP Conf. Proc. 2017, 1832, 130034. [Google Scholar]
  192. Mauger, A.; Julien, C.M. V2O5 thin films for energy storage and conversion. AIMS Mater. Sci. 2018, 5, 349–401. [Google Scholar] [CrossRef]
  193. McGraw, J.; Perkins, J.; Zhang, J.G.; Liu, P.; Parilla, P.; Turner, J.A.; Schulz, D.L.; Curtis, C.J.; Ginley, D.S. Next generation V2O5 cathode materials for Li rechargeable batteries. Solid State Ion. 1998, 113, 407–413. [Google Scholar] [CrossRef]
  194. Julien, C.; Haro-Poniatowski, E.; Camacho-Lopez, M.A.; Escobar-Alarcon, L.; Jimenez-Jarquin, J. Applications of V2O5 thin films obtained by pulsed laser deposition in solid state batteries. Bull. Am. Phys. Soc. 1999, 44, 1274. [Google Scholar]
  195. Julien, C.; Haro-Poniatowski, E.; Camacho-Lopez, M.A.; Escobar-Alarcon, L.; Jimenez-Jarquin, J. Growth of V2O5 thin films by pulsed laser deposition and their applications in lithium microbatteries. Mater. Sci. Eng. B 1999, 65, 170–176. [Google Scholar] [CrossRef]
  196. Ramana, C.V.; Smith, R.J.; Hussain, O.M.; Julien, C. On the growth mechanism of pulsed-laser deposited vanadium oxide thin films. Mater. Sci. Eng. B 2004, 111, 218–225. [Google Scholar] [CrossRef]
  197. McGraw, J.; Perkins, J.; Zhang, J.G.; Parilla, P.; Ciszek, T.F.; Fu, M.L.; Trickett, D.M.; Turner, J.A.; Ginley, D. Structural phase transformations in V2O5 thin-film cathode material for rechargeable batteries. Mater. Res. Soc. Symp. Proc. 1998, 496, 373–378. [Google Scholar] [CrossRef]
  198. McGraw, J.; Bahn, C.; Parilla, P.; Perkins, J.; Readey, D.; Ginley, D. Li ion diffusion measurements in crystalline and amorphous V2O5 thin-film battery cathodes. Mater. Res. Soc. Symp. Proc. 1999, 575, 103–108. [Google Scholar] [CrossRef]
  199. McGraw, J.M.; Perkins, J.D.; Hasoon, F.; Parilla, P.A.; Warmsingh, C.; Ginley, D.S. Pulsed laser deposition of oriented V2O5 thin films. J. Mater. Res. 2000, 15, 2249–2265. [Google Scholar] [CrossRef]
  200. Zhang, J.G.; McGraw, J.M.; Turner, J.; Ginley, D. Charging capacity and cycling stability of  VOx films prepared by pulsed laser deposition. J. Electrochem. Soc. 1997, 144, 1630–1634. [Google Scholar] [CrossRef]
  201. Madhuri, K.V.; Rao, K.S.; Naidu, B.S.; Hussain, O.M.; Pinto, R. Characterization of laser-ablated V2O5 thin films. J. Mater. Sci. Mater. Electron. 2002, 13, 425–432. [Google Scholar] [CrossRef]
  202. Iida, Y.; Kaneko, Y.; Kanno, Y. Fabrication of pulsed-laser deposited V2O5 thin films for electrochromic devices. J. Mater. Process. Technol. 2008, 197, 261–267. [Google Scholar] [CrossRef]
  203. Ramana, C.V.; Hussain, O.M.; Pinto, R.; Julien, C.M. Microstructure features of pulsed-laser deposited V2O5 thin films. Appl. Surf. Sci. 2003, 207, 135–138. [Google Scholar] [CrossRef]
  204. Ramana, C.V.; Smith, R.J.; Hussain, O.M.; Julien, C.M. Growth and surface characterization of V2O5 thin films made by pulsed laser deposition. J. Vac. Sci. Technol. A 2004, 22, 2453–2458. [Google Scholar] [CrossRef]
  205. Ramana, C.V.; Smith, R.J.; Hussain, O.M.; Chusuei, C.C.; Julien, C.M. Correlation between growth conditions, microstructure and optical properties in pulsed-laser-deposited V2O5 thin films. Chem. Mater. 2005, 17, 1213–1219. [Google Scholar] [CrossRef]
  206. Iida, Y.; Kanno, Y. Doping effect of M (M = Nb, Ce, Nd, Dy, Sm, Ag, and/or Na) on the growth of pulsed-laser deposited V2O5 thin films. J. Mater. Process. Technol. 2009, 209, 2421–2427. [Google Scholar] [CrossRef]
  207. Beke, S.; Giorgio, S.; Korosi, L.; Nanai, L.; Marine, W. Structural and optical properties of pulsed laser deposited V2O5 thin films. Thin Solid Films 2008, 516, 4659–4664. [Google Scholar] [CrossRef]
  208. Bowman, R.M.; Gregg, J.M. VO2 thin films: Growth and the effect of applied strain on their resistance. J. Mater. Sci.-Mater. Electron. 1998, 9, 187–191. [Google Scholar] [CrossRef]
  209. Deng, Y.; Pelton, A.; Mayanovic, R. Comparison of vanadium oxide thin films prepared using femtosecond and nanosecond pulsed laser deposition. MRS Adv. 2016, 1, 2737–2742. [Google Scholar] [CrossRef]
  210. Ramana, C.V.; Hussain, O.M.; Pinto, R.; Julien, C.M. Surface and material properties of nanostructured vanadium oxide films made by pulsed-laser ablation. In Proceedings of the 203rd ECS Meeting, Paris, France, 27 April–2 May 2003. [Google Scholar]
  211. Ramana, C.V.; Smith, R.J.; Hussain, O.M.; Massot, M.; Julien, C.M. Surface analysis of pulsed-laser-deposited V2O5 films and their lithium intercalated products studied by Raman spectroscopy. Surf. Interface Anal. 2005, 37, 406–411. [Google Scholar] [CrossRef]
  212. Iida, Y.; Takashima, H.; Kanno, Y. Optical and electrical properties of V2O5 thin films fabricated by pulsed laser deposition. Jpn. J. Appl. Phys. 2008, 47, 633–636. [Google Scholar] [CrossRef]
  213. Shibuya, K.; Sawa, A. Optimization of conditions for growth of vanadium dioxide thin films on silicon by pulsed-laser deposition. AIP Adv. 2015, 5, 107118. [Google Scholar] [CrossRef] [Green Version]
  214. Fang, G.J.; Liu, Z.L.; Wang, Y.Q.; Liu, H.H.; Yao, K.L. Orientated growth of V2O5 electrochromic thin films on transparent conductive glass by pulsed excimer laser ablation technique. J. Phys. D Appl. Phys. 2000, 33, 3018–3021. [Google Scholar] [CrossRef]
  215. Fang, G.J.; Yao, K.L.; Liu, Z.L. Fabrication and electrochromic properties of double layer WO3(V)/V2O5(Ti) thin films prepared by pulsed laser ablation technique. Thin Solid Films 2001, 394, 64–71. [Google Scholar] [CrossRef]
  216. Fang, G.J.; Liu, Z.L.; Wang, Y.; Liu, Y.H.; Yao, K.L. Synthesis and structural, electrochromic characterization of pulsed laser deposited vanadium oxide thin films. J. Vac. Sci. Technol. A 2001, 19, 887–892. [Google Scholar] [CrossRef]
  217. Rougier, A.; Blyr, A. Electrochromic properties of vanadium tungsten oxide thin films grown by pulsed laser deposition. Electrochim. Acta 2001, 46, 1945–1950. [Google Scholar] [CrossRef]
  218. Huotari, J.; Lappalainen, J.; Eriksson, J.; Bjorklund, R.; Heinonen, E.; Miinalainen, I.; Puustinen, J.; Spetz, A.L. Synthesis of nanostructured solid-state phases of V7O16 and V2O5 compounds for ppb-level detection of ammonia. J. Alloy. Compd. 2016, 675, 433–440. [Google Scholar] [CrossRef]
  219. Huotari, J.; Bjorklund, R.; Lappalainen, J.; Lloyd Spetz, A. Nanostructured mixed phase vanadium oxide thin films as highly sensitive ammonia sensor material. Procedia Eng. 2014, 87, 1035–1038. [Google Scholar] [CrossRef]
  220. Ramana, C.V.; Julien, C.M. Electrochemical performance of pulsed-laser deposited V2O5 films. In ESC Meeting Abstracts; The Electrochemical Society: Pennington, NJ, USA, 2006; Volume 210, p. 221. [Google Scholar]
  221. Julien, C. Technology of microbatteries and its applications. In Trends in Materials Science; Radhakrishna, S., Ed.; Narosa: New Delhi, India, 1997; pp. 24–43. [Google Scholar]
  222. Zhao, S.L.; Wen, J.B.; Zhu, Y.M.; Qin, Q.Z. Fabrication and electrochemical properties of all solid state 0.3Ag0V2O5/Lipon/Li thin film battery. J. Funct. Mater. 2008, 39, 91–94. [Google Scholar]
  223. Petnikota, S.; Chua, R.; Zhou, Y.; Edison, E.; Srinivasan, M. Amorphous vanadium oxide thin films as stable performing cathodes of lithium and sodium-ion batteries. Nanoscale Res. Lett. 2018, 13, 363. [Google Scholar] [CrossRef] [PubMed]
  224. West, K.; Zachau-Christiansen, B.; Jacobsen, T.; Atlung, S. V6O13 as cathode material for lithium cells. J. Power Sources 1985, 14, 235–245. [Google Scholar] [CrossRef]
  225. Makimura, Y.; Rougier, A.; Tarascon, J.M. Pulsed laser deposited iron fluoride thin films for lithium-ion batteries. Appl. Surf. Sci. 2006, 252, 4587–4592. [Google Scholar] [CrossRef]
  226. Makimura, Y.; Rougier, A.; Laffont, L.; Womes, M.; Jumas, J.C.; Leriche, J.B.; Tarascon, J.M. Electrochemical behaviour of low temperature grown iron fluoride thin films. Electrochem. Commun. 2006, 8, 1769–1774. [Google Scholar] [CrossRef]
  227. Santos-Ortiz, R.; Volkov, V.; Schmid, S.; Kuo, F.L.; Kisslinger, K.; Nag, S.; Banerjee, R.; Zhu, Y.; Shepherd, N.D. Microstructure and electronic band structure of pulsed laser deposited iron fluoride thin film for battery electrodes. ACS Appl. Mater. Interfaces 2013, 5, 2387–2391. [Google Scholar] [CrossRef] [PubMed]
  228. Al-Khateeb, S.; Lind, A.G.; Santos-Ortiz, R.; Shepherd, N.D.; Jones, K.S. Cycling performance and morphological evolution of pulsed laser-deposited FeF2 thin film cathodes for Li-ion batteries. J. Mater. Sci. 2015, 50, 5174–5182. [Google Scholar] [CrossRef]
  229. Julien, C.M.; Mauger, A.; Vijh, A.; Zaghib, K. Lithium Batteries: Science and Technology; Springer: Cham, Switzerland, 2016. [Google Scholar]
  230. Haro-Poniatowski, E.; Jouanne, M.; Morhange, J.F.; Julien, C.; Diamant, R.; Fernandez-Guasti, M.; Fuentes, G.A.; Alonso, J.C. Micro-Raman characterization of WO3 and MoO3 thin films obtained by pulsed laser irradiation. Appl. Surf. Sci. 1998, 127, 674–678. [Google Scholar] [CrossRef]
  231. Ramana, C.V.; Hussain, O.M.; Julien, C.M. Effect of growth temperature on the optical properties of pulsed-laser deposited MoO3 thin films. Mater. Sci. Indian J. 2005, 1, 10–15. [Google Scholar]
  232. Ramana, C.V.; Hussain, O.M.; Julien, C.M. Electronic properties of MoO3 thin films grown by PLD and ARE techniques and their performance upon lithium intercalation. In Proceedings of the Extended Abstract of the 208th ECS Meeting, Los Angeles, CA, USA, 16–21 October 2005; p. 885. [Google Scholar]
  233. Al-Kuhaili, M.; Durrani, S.; Bakhtiari, I. Pulsed laser deposition of molybdenum oxide thin films. Appl. Phys. A Mater. Sci. Process. 2010, 98, 609–615. [Google Scholar] [CrossRef]
  234. Ramana, C.V.; Julien, C.M. Chemical and electrochemical properties of molybdenum oxide thin films prepared by reactive pulsed-laser assisted deposition. Chem. Phys. Lett. 2006, 428, 114–118. [Google Scholar] [CrossRef]
  235. Ramana, C.V.; Atuchin, V.V.; Pokrovsky, L.D.; Beker, U.; Julien, C.M. Structure and chemical properties of molybdenum oxide thin films. J. Vac. Sci. Technol. A 2007, 25, 1166–1171. [Google Scholar] [CrossRef]
  236. Ramana, C.V.; Hussain, O.M.; Julien, C.M. Electronic properties and performance upon lithium intercalation of MoO3 thin films grown by PLD. ECS Trans. 2006, 1, 1–7. [Google Scholar]
  237. Puppala, H.K.; Pelton, A.T.; Mayanovic, R.A. A comparative characterization study of molybdenum oxide thin films grown using femtosecond and nanosecond pulsed laser deposition. Mater. Res. Soc. Adv. 2016, 1, 2585–2590. [Google Scholar] [CrossRef]
  238. Sunu, S.S.; Prabhu, E.; Jayaraman, V.; Gnanasekar, K.I.; Gnanasekaran, T. Gas sensing properties of PLD made MoO3 films. Sens. Actuators B Chem. 2003, 94, 189–196. [Google Scholar] [CrossRef]
  239. Ashrafi, M.A.; Ranjbar, M.; Kalhori, H.; Salamati, H. Pulsed laser deposition of Mo-V-O thin films for chromogenic applications. Thin Solid Films 2017, 621, 220–228. [Google Scholar] [CrossRef]
  240. Chang, C.C.; Luo, J.Y.; Chen, T.K.; Yeh, K.W.; Huang, T.W.; Hsu, C.H.; Chao, W.H.; Ke, C.T.; Hsu, P.C.; Wang, M.J.; et al. Pulsed laser deposition of (MoO3)1−x(V2O5)x thin films: Preparation, characterization and gasochromic studies. Thin Solid Films 2010, 519, 1552–1557. [Google Scholar] [CrossRef]
  241. Rougier, A.; Portemer, F.; Quédé, A.; El Marssi, M. Characterization of pulsed laser deposited WO3 thin films for electrochromic devices. Appl. Surf. Sci. 1999, 153, 1–9. [Google Scholar] [CrossRef]
  242. Qiu, H.; Lu, Y.F. Studies of pulsed laser deposition mechanism of WO3 thin films. Jpn. J. Appl. Phys. 2001, 40, 183–187. [Google Scholar] [CrossRef]
  243. Ramana, C.V.; Utsunomiya, S.; Ewing, R.C.; Julien, C.M.; Becker, U. Electron microscopy investigation of structural transformations in tungsten oxide (WO3) thin films. Phys. Status Solidi A 2005, 202, R108–R110. [Google Scholar] [CrossRef]
  244. Ramana, C.V.; Utsunomiya, S.; Ewing, R.C.; Julien, C.M.; Becker, U. Structural stability and phase transitions in WO3 thin films. J. Phys. Chem. B 2006, 110, 10430–10435. [Google Scholar] [CrossRef] [PubMed]
  245. Mitsugi, F.; Hiraiwa, E.; Ikegami, T.; Ebihara, K.; Thareja, R.K. WO3 thin films prepared by pulsed laser deposition. Jpn. J. Appl. Phys. 2002, 41, 5372–5375. [Google Scholar] [CrossRef]
  246. Hussain, O.M.; Swapnasmitha, A.S.; John, J.; Pinto, R. Structure and morphology of laser-ablated WO3 thin films. Appl. Phys. A 2005, 81, 1291–1297. [Google Scholar] [CrossRef]
  247. Suda, Y.; Kawasaki, H.; Ohshima, T.; Yagyu, Y. Characteristics of tungsten oxide thin films prepared on the flexible substrates using pulsed laser deposition. Thin Solid Films 2008, 516, 4397–4401. [Google Scholar] [CrossRef]
  248. Fang, G.J.; Liu, Z.L.; Sun, G.C.; Yao, K.L. Preparation and electrochromic properties of nanocrystalline WO3 thin films prepared by pulsed excimer laser ablation technique. Phys. Status Solidi A 2001, 184, 129–137. [Google Scholar] [CrossRef]
  249. György, E.; Socol, G.; Mihailescu, I.N.; Ducu, C.; Ciuca, S. Structural and optical characterization of WO3 thin films for gas sensor applications. J. Appl. Phys. 2005, 97, 093527. [Google Scholar] [CrossRef]
  250. Caruana, A.J.; Cropper, M.D. Orientationally textured thin films of WOx deposited by pulsed laser deposition. Mater. Res. Express 2014, 1, 046408. [Google Scholar] [CrossRef]
  251. Impey, S.A.; Garcia-Miguel, J.L.; Allen, S.; Blyr, A.; Boessay, I.; Rougier, A. Colour neutrality for thin oxide films from pulsed laser deposition and sol-gel. Electrochem. Soc. Proc. 2003, 2003–17, 103–118. [Google Scholar]
  252. Boyadjiev, S.I.; Stefan, N.; Szilágyi, I.M.; Mihailescu, N.; Visan, A.; Mihailescu, I.N.; Stan, G.E.; Besleaga, C.; Iliev, M.T.; Gesheva, K.A. Characterization of MAPLE deposited WO3 thin films for electrochromic applications. J. Phys. Conf. Ser. 2017, 780, 012013. [Google Scholar] [CrossRef]
  253. Kawasaki, H.; Ueda, T.; Suda, T.; Ohshima, T. Properties of metal doped tungsten oxide thin films for NOx gas sensors grown by PLD method combined with sputtering process. Sens. Actuators B 2004, 100, 266–269. [Google Scholar] [CrossRef]
  254. Leidinger, M.; Huotari, J.; Sauerwald, T.; Lappalainen, J.; Schütze, A. Selective detection of naphthalene with nanostructured WO3 gas sensors prepared by pulsed laser deposition. J. Sens. Sens. Syst. 2016, 5, 147–156. [Google Scholar] [CrossRef]
  255. Boyadjiev, S.I.; Georgieva, V.; Stefan, N.; Stan, G.E.; Mihailescu, N.; Visan, A.; Mihailescu, I.N.; Besleag, C.; Szilágyi, I.M. Characterization of PLD grown WO3 thin films for gas sensing. Appl. Surf. Sci. 2017, 417, 218–223. [Google Scholar] [CrossRef]
  256. Xia, H.; Wang, H.L.; Xiao, W.; Lai, M.O.; Li, L. Thin film Li electrolytes for all-solid-state micro-batteries. Int. J. Surf. Sci. Eng. 2009, 3, 23–43. [Google Scholar] [CrossRef]
  257. Ren, Y.; Kai Chen, K.; Chen, R.; Liu, T.; Zhang, Y.; Nan, C.W. Oxide electrolytes for lithium batteries. J. Am. Ceram. Soc. 2015, 98, 3603–3623. [Google Scholar] [CrossRef]
  258. Cao, C.; Li, Z.B.; Wang, X.L.; Zhao, X.B.; Han, W.Q. Recent advances in inorganic solid electrolytes for lithium batteries. Front. Energy Res. 2014, 2, 25. [Google Scholar] [CrossRef]
  259. Zhao, S.L.; Fu, Z.W.; Qin, Q.Z. A solid-state electrolyte lithium phosphorous oxynitride film prepared by pulsed laser deposition. Thin Solid Films 2002, 415, 108–113. [Google Scholar] [CrossRef]
  260. Ohta, N.; Takada, K.; Osada, M.; Zhang, L.Q.; Sasaki, T.; Watanabe, M. Solid electrolyte, thio-LiSICON, thin film prepared by pulsed laser deposition. J. Power Sources 2005, 146, 707–710. [Google Scholar] [CrossRef]
  261. Ahn, J.K.; Yoon, S.G. Characteristics of amorphous lithium lanthanum titanate electrolyte thin films grown by PLD for use in rechargeable lithium microbatteries. Electrochem. Solid-State Lett. 2005, 8, A75–A78. [Google Scholar] [CrossRef]
  262. Furusawa, S.I.; Kamiyama, A.; Tsurui, T. Fabrication and ionic conductivity of amorphous lithium meta-silicate thin film. Solid State Ion. 2008, 179, 536–542. [Google Scholar] [CrossRef]
  263. Furusawa, S.I.; Shimizu, S.; Sekine, K.; Tabuchi, H. Preparation and ionic conductivity of β-LiAlSiO4 thin film. Solid State Ion. 2004, 167, 325–329. [Google Scholar]
  264. Reinacher, J.; Berendts, S.; Janek, J. Preparation and electrical properties of garnet-type Li6BaLa2Ta2O12 lithium solid electrolyte thin films prepared by pulsed laser deposition. Solid State Ion. 2014, 258, 1–7. [Google Scholar] [CrossRef]
  265. Yu, X.-H.; Bates, J.B.; Jellison, G.E.; Hart, F.X. A stable thin-film lithium electrolyte: Lithium phosphorus oxynitride. J. Electrochem. Soc. 1997, 144, 524–532. [Google Scholar] [CrossRef]
  266. Zhao, S.L.; Wen, J.B.; Fan, L.M.; Qin, Q.Z. preparation of thin film battery by pulse laser deposition. J. Funct. Mater. 2006, 37, 172–177. [Google Scholar]
  267. West, W.C.; Hood, Z.D.; Adhikari, S.P.; Liang, C.; Lachgar, A.; Motoyama, M.; Iriyama, Y. Reduction of charge-transfer resistance at the solid electrolyte-electrode interface by pulsed laser deposition of films from a crystalline Li2PO2N source. J. Power Sources 2016, 312, 116–122. [Google Scholar] [CrossRef]
  268. Morcrette, M.; Llorente, A.G.; Laurent, A.; Perrière, J.; Barboux, P.; Boilot, J.P.; Raymond, O.; Brousse, T. Growth by laser ablation of Ti-based oxide films with different valency states. Appl. Phys. A 1998, 67, 425–428. [Google Scholar] [CrossRef]
  269. Ahn, J.K.; Yoon, S.G. Characteristics of perovskite (Li0.5La0.5)TiO3 solid electrolyte thin films grown by pulsed laser deposition for rechargeable lithium microbattery. Electrochim. Acta 2004, 50, 371–374. [Google Scholar] [CrossRef]
  270. Furusawa, S.; Tabuchi, H.; Sugiyama, T.; Tao, S.; Irvine, J.T.S. Ionic conductivity of amorphous lithium lanthanum titanate thin film. Solid State Ion. 2005, 176, 553–558. [Google Scholar] [CrossRef]
  271. Maqueda, O.; Sauvage, F.; Laffont, L.; Martinez-Sarrion, M.L.; Mestres, L.; Baudrin, E. Structural, microstructural and transport properties study of lanthanum lithium titanium perovskite thin films grown by pulsed laser deposition. Thin Solid Films 2008, 516, 1651–1655. [Google Scholar] [CrossRef]
  272. Ohnishi, T.; Takada, K. Synthesis and orientation control of Li-ion conducting epitaxial Li0.33La0.56TiO3 solid electrolyte thin films by pulsed laser deposition. Solid State Ion. 2012, 228, 80–82. [Google Scholar] [CrossRef]
  273. Aguesse, F.; Roddatis, V.; Roqueta, J.; García, P.; Pergolesi, D.; Santiso, J.; Kilner, J.A. Microstructure and ionic conductivity of LLTO thin films: Influence of different substrates and excess lithium in the target. Solid State Ion. 2015, 272, 1–8. [Google Scholar] [CrossRef] [Green Version]
  274. Lee, J.L.; Wang, Z.; Xin, H.L.; Wynn, T.A.; Meng, Y.S. Amorphous lithium lanthanum titanate for solid-state microbatteries. J. Electrochem. Soc. 2017, 164, A6268–A6273. [Google Scholar] [CrossRef]
  275. Thangadurai, V.; Kaack, H.; Weppner, W. Novel fast lithium ion conduction in garnet-type Li5La3M2O12 (M = Nb, Ta). J. Am. Ceram. Soc. 2003, 86, 437–440. [Google Scholar] [CrossRef]
  276. Saccoccio, M.; Yu, J.; Lu, Z.; Kwok, S.C.T.; Wang, J.; Yeung, K.K.; Yuen, M.M.F.; Ciucci, F. Low temperature pulsed laser deposition of garnet Li6.4La3Zr1.4Ta0.6O12 films as all solid-state lithium battery electrolytes. J. Power Sources 2017, 365, 43–52. [Google Scholar] [CrossRef]
  277. Garbayo, I.; Struzik, M.; Bowman, W.J.; Pfenninger, R.; Stilp, E.; Rupp, J.L.M. Glass-type polyamorphism in Li-garnet thin film solid state battery conductors. Adv. Energy Mater. 2018, 8, 1702265. [Google Scholar] [CrossRef]
  278. Tan, J.; Tiwari, A. Fabrication and characterization of Li7La3Zr2O12 thin films for lithium ion battery. ECS Solid State Lett. 2012, 1, Q57–Q60. [Google Scholar] [CrossRef]
  279. Park, J.S.; Cheng, L.; Zorba, V.; Mehta, A.; Cabana, J.; Chen, G.; Doeff, M.M.; Richardson, T.J.; Park, J.H.; Son, J.-W.; et al. Effects of crystallinity and impurities on the electrical conductivity of Li–La–Zr–O thin films. Thin Solid Films 2015, 576, 55–60. [Google Scholar] [CrossRef]
  280. Kim, S.; Hirayama, M.; Taminato, S.; Kanno, R. Epitaxial growth and lithium ion conductivity of lithium-oxide garnet for an all solid-state battery electrolyte. Dalton Trans. 2013, 42, 13112. [Google Scholar] [CrossRef]
  281. Rawlence, M.; Garbayo, I.; Buecheler, S.; Rupp, J.L.M. On the chemical stability of post-lithiated garnet Al-stabilized Li7La3Zr2O12 solid state electrolyte thin films. Nanoscale 2016, 8, 14746–14753. [Google Scholar] [CrossRef]
  282. Kuwata, N.; Iwagami, N.; Tanji, Y.; Matsuda, Y.; Kawamura, J. Characterization of thin-film lithium batteries with stable thin-film Li3PO4 solid electrolytes fabricated by ArF excimer laser deposition. J. Electrochem. Soc. 2011, 157, A521–A527. [Google Scholar] [CrossRef]
  283. Sakurai, Y.; Sakuda, A.; Hayashi, A.; Tatsumisago, M. Preparation of amorphous Li4SiO4-Li3PO4 thin films by pulsed laser deposition for all-solid-state lithium secondary batteries. Solid State Ion. 2011, 182, 59–63. [Google Scholar] [CrossRef]
  284. Nakagawa, A.; Kuwata, N.; Matsuda, Y.; Kawamura, J. Characterization of stable solid electrolyte lithium silicate for thin film lithium battery. J. Phys. Soc. Jpn. A 2010, 79, 98–101. [Google Scholar] [CrossRef]
  285. Furusawa, S.I.; Kasahara, T.; Kamiyama, A. Fabrication and ionic conductivity of Li2SiO3 thin film. Solid State Ion. 2009, 180, 649–653. [Google Scholar] [CrossRef]
  286. Sakuda, A.; Hayashi, A.; Hama, S.; Tatsumisago, M. Preparation of highly lithium-ion conductive 80Li2S-20P2O5 thin-film electrolytes using pulsed laser deposition. J. Am. Ceram. Soc. 2010, 93, 765–768. [Google Scholar] [CrossRef]
  287. Quan, Z.; Hirayama, M.; Sato, D.; Zheng, Y.; Yano, T.; Hara, K.; Suzuki, K.; Hara, M.; Kanno, R. Effect of excess Li2S on electrochemical properties of amorphous Li3PS4 films synthesized by pulsed laser deposition. J. Am. Ceram. Soc. 2017, 100, 746–753. [Google Scholar] [CrossRef]
  288. Konishi, H.; Suzuki, K.; Taminato, S.; Kim, K.; Zheng, Y.; Kim, S.; Lim, J.; Hirayama, M.; Son, J.Y.; Cui, Y.; et al. Effect of surface Li3PO4 coating on LiNi0.5Mn1.5O4 epitaxial thin film electrodes synthesized by pulsed laser deposition. J. Power Sources 2014, 269, 293–298. [Google Scholar] [CrossRef]
  289. Yubuchi, S.; Ito, Y.; Matsuyama, T.; Hayashi, A.; Tatsumisago, M. 5 V class LiNi0.5Mn1.5O4 positive electrode coated with Li3PO4 thin film for all-solid-state batteries using sulfide solid electrolyte. Solid State Ion. 2016, 285, 79–82. [Google Scholar] [CrossRef]
  290. Aso, K.; Sakuda, A.; Hayashi, A.; Tatsumisago, M. All-solid-state lithium secondary batteries using NiS-carbon fiber composite electrodes coated with Li2S−P2S5 solid electrolytes by pulsed laser deposition. ACS Appl. Mater. Interfaces 2013, 5, 686–690. [Google Scholar] [CrossRef]
  291. Ito, Y.; Sakuda, A.; Ohtomo, T.; Hayashi, A.; Tatsumisago, M. Preparation of Li2S–GeS2 solid electrolyte thin films using pulsed laser deposition. Solid State Ion. 2013, 236, 1–4. [Google Scholar] [CrossRef]
  292. Zhang, Z.; Chou, S.; Gu, Q.; Liu, H.; Li, H.; Ozawa, K.; Wang, J. Enhancing the high rate capability and cycling stability of LiMn2O4 by coating of solid-state electrolyte LiNbO3. ACS Appl. Mater. Interfaces 2014, 6, 22155–22165. [Google Scholar] [CrossRef]
  293. Uemura, T.; Goto, K.; Ogawa, M.; Harada, K. High-power all-solid secondary battery with high heat resistance. SEI Technol. Rev. 2013, 77, 132–136. [Google Scholar]
  294. Zhao, S.; Qin, Q. Li–V–Si–O thin film electrolyte for all-solid-state Li-ion battery. J. Power Sources 2003, 122, 174–180. [Google Scholar] [CrossRef]
  295. Kawamura, J.; Kuwata, N.; Toribami, K.; Sata, N.; Kamishima, O.; Hattori, T. Preparation of amorphous thin films by pulsed laser deposition. Solid State Ion. 2004, 175, 273–276. [Google Scholar] [CrossRef]
  296. Glass, A.; Nassau, K.; Negran, T. Ionic Conductivity of quenched alkali niobate and tantalate glasses. J. Appl. Phys. 1978, 49, 4808–4811. [Google Scholar] [CrossRef]
  297. Ohta, N.; Takada, K.; Sakaguchi, I.; Zhang, L.; Ma, R.; Fukuda, K.; Osada, M.; Sasaki, T. LiNbO3-coated LiCoO2 as cathode material for all solid-state lithium secondary batteries. Electrochem. Commun. 2007, 9, 1486–1490. [Google Scholar] [CrossRef]
  298. Marsh, A.M.; Harkness, S.D.; Qian, F.; Singh, R.K. Pulsed laser deposition of high quality LiNbO3 films on sapphire substrates. Appl. Phys. Lett. 1993, 62, 952–954. [Google Scholar] [CrossRef]
  299. Perea, A.; Gonzalo, J.; Afonso, C.N.; Vivien, C.; Leborgne, C. Fast imaging of the plume expansion produced by laser ablation of LiNbO3. Appl. Phys. A 1999, 69, S141–S144. [Google Scholar] [CrossRef]
  300. Madian, M.; Eychmüller, A.; Giebeler, L. Current advances in TiO2-based nanostructure electrodes for high performance lithium ion batteries. Batteries 2018, 4, 7. [Google Scholar] [CrossRef]
  301. Murugesan, S.; Kuppusami, P.; Parvathavarthini, N.; Mohandas, E. Pulsed laser deposition of anatase and rutile TiO2 thin films. Surf. Coat. Technol. 2007, 201, 7713–7719. [Google Scholar] [CrossRef]
  302. Suda, Y.; Kawasaki, H.; Ueda, T.; Ohshima, T. Preparation of high-quality nitrogen doped TiO2 thin film as a photocatalyst using a pulsed laser deposition method. Thin Solid Films 2004, 453, 162–166. [Google Scholar] [CrossRef]
  303. Kim, J.H.; Seonghoon Lee, S.; Im, H.S. The effect of target density and its morphology on TiO2 thin films grown on Si(100) by PLD. Appl. Surf. Sci. 1999, 151, 6–16. [Google Scholar] [CrossRef]
  304. Yamamoto, S.; Sumita, T.; Sugiharuto, M.; Miyashita, A.; Naramoto, H. Preparation of epitaxial TiO2 films by pulsed laser deposition technique. Thin Solid Films 2001, 401, 88–93. [Google Scholar] [CrossRef]
  305. Inoue, N.; Yuasa, H.; Okoshi, M. TiO2 thin films prepared by PLD for photocatalytic applications. Appl. Surf. Sci. 2002, 197, 393–397. [Google Scholar] [CrossRef]
  306. Choi, Y.; Yamamoto, S.; Umebayashi, T.; Yoshikawa, M. Fabrication and characterization of anatase TiO2 thin film on glass substrate grown by pulsed laser deposition. Solid State Ion. 2004, 172, 105–108. [Google Scholar] [CrossRef]
  307. Ohzuku, T.; Ueda, A.; Yamamoto, N. Zero-strain insertion material of Li[Li1/3Ti5/3]O4 for rechargeable lithium cells. J. Electrochem. Soc. 1995, 142, 1431–1435. [Google Scholar] [CrossRef]
  308. Deng, J.Q.; Lu, Z.G.; Belharouak, I.; Amine, K.; Chung, C.Y. Preparation and electrochemical properties of Li4Ti5O12 thin film electrodes by pulsed laser deposition. J. Power Sources 2009, 193, 816–821. [Google Scholar] [CrossRef]
  309. Yu, X.; Wang, R.; He, Y.; Hu, Y.; Li, H.; Huang, X. Electrochromic behavior of transparent Li4Ti5O12/FTO electrode. Electrochem. Solid-State Lett. 2010, 13, J99–J101. [Google Scholar] [CrossRef]
  310. Hirayama, M.; Kim, K.; Toujigamori, T.; Cho, W.; Kanno, R. Epitaxial growth and electrochemical properties of Li4Ti5O12 thin-film lithium battery anodes. Dalton Trans. 2011, 40, 2882–2887. [Google Scholar] [CrossRef] [PubMed]
  311. Kim, K.; Toujigamori, T.; Suzuki, K.; Taminato, S.; Tamura, K.; Mizuki, J.; Hirayama, M.; Kanno, R. Characterization of nano-sized epitaxial Li4Ti5O12(110) film electrode for lithium batteries. Electrochemistry 2012, 80, 800–803. [Google Scholar] [CrossRef]
  312. Deng, J.; Lu, Z.; Chung, C.Y.; Hanc, X.; Wang, Z.; Zhou, H. Electrochemical performance and kinetic behavior of lithium ion in Li4Ti5O12 thin film electrodes. Appl. Surf. Sci. 2014, 314, 936–941. [Google Scholar] [CrossRef]
  313. Sugiyama, J.; Umegaki, I.; Uyama, T.; McFadden, R.M.L.; Shiraki, S.; Hitosugi, T.; Salman, Z.; Saadaoui, H.; Morris, G.D.; MacFarlane, W.A.; et al. Lithium diffusion in spinel Li4Ti5O12 and LiTi2O4 films detected with 8Li β-NMR. Phys. Rev. B 2017, 96, 094402. [Google Scholar] [CrossRef]
  314. Zhao, M.; Lian, J.; Jia, Y.; Jin, K.; Xu, L.; Hu, Z.; Yang, X.; Kang, S. Investigation of the optical properties of LiTi2O4 and Li4Ti5O12 spinel films by spectroscopic ellipsometry. Opt. Mater. Express 2016, 6, 3366–3374. [Google Scholar] [CrossRef]
  315. Pfenninger, R.; Afyon, S.; Garbayo, I.; Struzik, M.; Rupp, J.L.M. Lithium titanate anode thin films for Li-ion solid state battery based on garnets. Adv. Funct. Mater. 2018, 28, 1800879. [Google Scholar] [CrossRef]
  316. Mesoraca, S.; Kleibeuker, J.E.; Prasad, B.; MacManus-Driscoll, J.L.; Blamire, M.G. Lithium outdiffusion in LiTi2O4 thin films grown by pulsed laser deposition. J. Cryst. Growth 2016, 454, 134–138. [Google Scholar] [CrossRef]
  317. Chopdekar, R.V.; Wong, F.J.; Takamura, Y.; Arenholz, E.; Suzuki, Y. Growth and characterization of superconducting spinel oxide LiTi2O4 thin films. Phys. C Supercond. 2009, 469, 1885–1891. [Google Scholar] [CrossRef]
  318. Kumatani, A.; Ohsawa, T.; Shimizu, R.; Takagi, Y.; Shiraki, S.; Hitosugi, T. Growth processes of lithium titanate thin films deposited by using pulsed laser deposition. Appl. Phys. Lett. 2012, 101, 123103. [Google Scholar] [CrossRef]
  319. Cunha, D.M.; Hendriks, T.A.; Vasileiadis, A.; Vos, C.M.; Verhallen, T.; Singh, D.P.; Wagemaker, M.; Huijben, M. Doubling reversible capacities in epitaxial Li4Ti5O12 thin film anodes for microbatteries. ACS Appl. Energy Mater. 2019, 2, 3410–3418. [Google Scholar] [CrossRef]
  320. Tang, S.B.; Xia, H.; Lai, M.O.; Lu, L. Amorphous LiNiVO4 thin-film anode for microbatteries grown by pulsed laser deposition. J. Power Sources 2006, 159, 685–689. [Google Scholar] [CrossRef]
  321. Han, T.; Goodenough, J.B. 3-V full cell performance of anode framework TiNb2O7/spinel LiNi0.5Mn1.5O4. Chem. Mater. 2011, 23, 3404–3407. [Google Scholar] [CrossRef]
  322. Daramalla, V.; Rao Penki, T.; Munichandraiah, N.; Krupanidhi, S.B. Fabrication of TiNb2O7 thin film electrodes for Li-ion micro-batteries by pulsed laser deposition. Mater. Sci. Eng. B 2016, 213, 90–97. [Google Scholar] [CrossRef]
  323. Daramalla, V.; Venkatesh, G.; Kishore, B.; Munichandraiah, N.; Krupanidhi, S.B. Superior electrochemical performance of amorphous titanium niobium oxide thin films for Li-ion thin film batteries. J. Electrochem. Soc. 2018, 165, A764–A772. [Google Scholar] [CrossRef]
  324. Mauger, A.; Xie, H.; Julien, C.M. Composite anodes for lithium-ion batteries: Status and trends. AIMS Mater. Sci. 2016, 3, 1054–1106. [Google Scholar] [CrossRef]
  325. Chen, L.B.; Xie, J.Y.; Yu, H.C.; Wang, T.H. An amorphous thin film Si anode with high capacity and long cycling life for lithium ion batteries. J. Appl. Electrochem. 2009, 39, 1157–1162. [Google Scholar] [CrossRef]
  326. Ohara, S.; Suzuki, J.; Sekine, K.; Takamura, T. A thin film silicon anode for Li-ion batteries having a very large specific capacity and long cycle life. J. Power Sources 2004, 136, 303–306. [Google Scholar] [CrossRef]
  327. Song, S.W.; Sriebel, K.A.; Reade, R.P.; Roberts, G.A.; Cairns, E.J. Electrochemical studies of nanoncrystalline Mg2Si thin film electrodes prepared by pulsed laser deposition. J. Electrochem. Soc. 2003, 150, A121–A127. [Google Scholar] [CrossRef]
  328. Park, M.S.; Wang, G.X.; Liu, H.K.; Dou, S.X. Electrochemical properties of Si thin film prepared by pulsed laser deposition for lithium ion micro-batteries. Electrochim. Acta 2006, 51, 5246–5249. [Google Scholar] [CrossRef]
  329. Xia, H.; Tang, S.; Lu, L. Properties of amorphous Si thin film anodes prepared by pulsed laser deposition. Mater. Res. Bull. 2007, 42, 1301–1309. [Google Scholar] [CrossRef]
  330. Chou, S.L.; Zhao, Y.; Wang, J.Z.; Chen, Z.X.; Liu, H.K.; Dou, S.X. Silicon/single-walled carbon nanotube composite paper as a flexible anode material for lithium ion batteries. J. Phys. Chem. C 2010, 114, 15862–15867. [Google Scholar] [CrossRef]
  331. Radhakrishnan, G.; Adams, P.M.; Foran, B.; Quinzio, M.V.; Brodie, M.J. Pulsed laser deposited Si on multilayer graphene as anode material for lithium ion batteries. APL Mater. 2013, 1, 062103. [Google Scholar] [CrossRef] [Green Version]
  332. Biserni, E.; Xie, M.; Brescia, R.; Scarpellini, A.; Hashempour, M.; Movahed, P.; George, S.M.; Bestetti, M.; Li Bassi, A.; Bruno, P. Silicon algae with carbon topping as thin-film anodes for lithium-ion microbatteries by a two-step facile method. J. Power Sources 2015, 274, 252–259. [Google Scholar] [CrossRef]
  333. Garino, N.; Biserni, E.; Bassi, A.L.; Bruno, P.; Gerbaldi, C. Mesoporous Si and multi-layered Si/C films by pulsed laser deposition as Li-ion microbattery anodes. J. Electrochem. Soc. 2015, 162, A1816–A1822. [Google Scholar] [CrossRef]
  334. Suzuki, N.; Butch-Cervera, R.; Ohnishi, T.; Takada, K. Silicon nitride thin film electrode for lithium-ion batteries. J. Power Sources 2013, 231, 186–189. [Google Scholar] [CrossRef]
  335. Bleu, Y.; Bourquard, F.; Tite, T.; Loir, A.-S.; Maddi, C.; Donnet, C.; Garrelie, F. Review of graphene growth from a solid carbon source by pulsed laser deposition (PLD). Front. Chem. 2018, 6, 574. [Google Scholar] [CrossRef] [PubMed]
  336. Kumar, S.R.; Nayak, P.K.; Hedhili, M.N.; Khan, M.A.; Alshareef, H.N. In situ growth of p and n-type graphene thin films and diodes by pulsed laser deposition. Appl. Phys. Lett. 2013, 103, 192109. [Google Scholar] [CrossRef] [Green Version]
  337. Ren, P.; Pu, E.; Liu, D.; Wang, Y.; Xiang, B.; Ren, X. Fabrication of nitrogen-doped graphenes by pulsed laser deposition and improved chemical enhancement for Raman spectroscopy. Mater. Lett. 2017, 204, 65–68. [Google Scholar] [CrossRef]
  338. Fortgang, P.; Tite, T.; Barnier, V.; Zehani, N.; Maddi, C.; Lagarde, F.; Loir, A.-S.; Jaffrezic-Renault, N.; Donnet, C.; Garrelie, F.; et al. Robust electrografting on self-organized 3D graphene electrodes. ACS Appl. Mater. Interfaces 2015, 8, 1424–1433. [Google Scholar] [CrossRef] [PubMed]
  339. Bourquard, F.; Bleu, Y.; Loi, A.-S.; Caja-Munoz, B.; Avila, J.; Asensio, M.-C.; Riamondi, G.; Shokouhi, M.; Rassas, I.; Farre, C.; et al. Electroanalytical performance of nitrogen-doped graphene films processed in one step by pulsed laser deposition directly coupled with thermal annealing. Materials 2019, 12, 666. [Google Scholar] [CrossRef]
  340. Maddi, C.; Bourquard, F.; Barnier, V.; Avila, J.; Asensio, M.C.; Tite, T.; Donnet, C.; Garrelie, F. Nano-architecture of nitrogen-doped graphene films synthesized from a solid CN source. Sci. Rep. 2018, 8, 3247. [Google Scholar] [CrossRef]
  341. Yu, S.H.; Feng, X.; Zhang, N.; Seok, J.; Abruna, H.D. Understanding conversion-type electrodes for lithium rechargeable batteries. Acc. Chem. Res. 2018, 51, 273–281. [Google Scholar] [CrossRef]
  342. Wang, K.X.; Li, Y.; Wu, X.Y.; Chen, J.S. Carbon nanocolumn arrays prepared by pulsed laser deposition for lithium ion batteries. J. Power Sources 2012, 203, 140–144. [Google Scholar] [CrossRef]
  343. Wei, K.; Zhao, Y.; Cui, Y.; Wang, J.; Cui, Y.; Zhu, R.; Zhuang, Q.; Xue, M. Lithium phosphorous oxynitride (LiPON) coated NiFe2O4 anode material with enhanced electrochemical performance for lithium ion batteries. J. Alloy Compd. 2018, 769, 110–119. [Google Scholar] [CrossRef]
  344. Pralong, V.; Leriche, J.B.; Beaudoin, B.; Naudin, E.; Morcrette, M.; Tarascon, J.M. Electrochemical study of nanometer Co3O4, Co, CoSb3 and Sb thin films toward lithium. Solid State Ion. 2004, 166, 295–305. [Google Scholar] [CrossRef]
  345. Song, S.W.; Reade, R.P.; Cairns, E.; Vaughey, J.T.; Thackeray, M.M.; Striebel, K.A. Cu2Sb thin-film electrodes prepared by pulsed laser deposition for lithium batteries. J. Electrochem. Soc. 2004, 151, A1012–A1019. [Google Scholar] [CrossRef]
  346. Teng, X.; Qin, Y.; Wang, X.; Li, H.; Shang, X.; Fan, S.; Li, Q.; Xu, J.; Cao, D.; Li, S. A nanocrystalline Fe2O3 film anode prepared by pulsed laser deposition for lithium-ion batteries. Nanoscale Res. Lett. 2018, 13, 60. [Google Scholar] [CrossRef] [PubMed]
  347. Fu, Z.W.; Wang, Y.; Zhang, Y.; Qin, Q.Z. Electrochemical reaction of nanocrystalline Co3O4 thin film with lithium. Solid State Ion. 2004, 170, 105–109. [Google Scholar] [CrossRef]
  348. Wang, Y.; Wang, H.; Wang, X. The cobalt oxide/hydroxide nanowall array film prepared by pulsed laser deposition for supercapacitors with superb-rate capability. Electrochim. Acta 2013, 92, 298–303. [Google Scholar] [CrossRef]
  349. Zhou, Y.; Zhang, H.; Xue, M.; Wu, C.; Wu, X.; Fu, Z. The electrochemistry of nanostructured In2O3 with lithium. J. Power Sources 2006, 162, 1373–1378. [Google Scholar] [CrossRef]
  350. Xue, M.Z.; Fu, Z.W. Pulsed laser deposited Sb2Se3 anode for lithium-ion batteries. J. Alloy. Compd. 2008, 458, 351–356. [Google Scholar] [CrossRef]
  351. Yu, L.; Wang, H.X.; Liu, Z.Y.; Fu, Z.W. Pulsed laser deposited FeOF as negative electrodes for rechargeable Li batteries. Electrochim. Acta 2010, 56, 767–775. [Google Scholar] [CrossRef]
  352. Rambabu, A.; Senthilkumar, B.; Dayamani, A.; Krupanidhi, S.B.; Barpanda, P. Preferentially oriented SrLi2Ti6O14 thin film anode for Li-ion microbatteries fabricated by pulsed laser deposition. Electrochim. Acta 2018, 269, 212–216. [Google Scholar] [CrossRef]
  353. Yu, X.Q.; He, Y.; Sun, J.P.; Tang, K.; Li, H.; Chen, L.Q.; Huang, X.J. Nanocrystalline MnO thin film anode for lithium ion batteries with low overpotential. Electrochem. Commun. 2009, 11, 791–794. [Google Scholar] [CrossRef]
  354. Wu, H.; Wei, K.; Tang, B.; Cui, Y.; Zhao, Y.; Xue, M.; Li, C.; Cui, Y. A novel Li3P-VP nanocomposite fabricated by pulsed laser deposition as anode material for high-capacity lithium ion batteries. J. Electroanal. Chem. 2019, 841, 21–25. [Google Scholar] [CrossRef]
Figure 1. Design principles for lithium microbatteries composed of a lithium film (Li), fast-ion conductor (FIC), mixed ionic-electronic conductor (MIC), current collector(s) (CC), silicon substrate (Si), glass substrate (Sub), and buffer layer (Buf). (a) four-layer design on a conducting substrate and (b) six-layer stack incorporating two metallic current collectors (Reproduced with permission from [25]. Copyright 2000 Springer).
Figure 1. Design principles for lithium microbatteries composed of a lithium film (Li), fast-ion conductor (FIC), mixed ionic-electronic conductor (MIC), current collector(s) (CC), silicon substrate (Si), glass substrate (Sub), and buffer layer (Buf). (a) four-layer design on a conducting substrate and (b) six-layer stack incorporating two metallic current collectors (Reproduced with permission from [25]. Copyright 2000 Springer).
Coatings 09 00386 g001
Figure 2. The relationship between impurity inclusions and growth conditions of PLD-grown LCO films established from spectroscopic Raman data (Reproduced with permission from [51]. Copyright 2017 AIP Publishing).
Figure 2. The relationship between impurity inclusions and growth conditions of PLD-grown LCO films established from spectroscopic Raman data (Reproduced with permission from [51]. Copyright 2017 AIP Publishing).
Coatings 09 00386 g002
Figure 3. Schematic representation of the fast-laser ablation growth according to the Li/Co ratio variation of the plume for a stoichiometric (a) and Li-enriched target (b). (Reproduced with permission from [62]. Copyright 2012 IOP Publishing).
Figure 3. Schematic representation of the fast-laser ablation growth according to the Li/Co ratio variation of the plume for a stoichiometric (a) and Li-enriched target (b). (Reproduced with permission from [62]. Copyright 2012 IOP Publishing).
Coatings 09 00386 g003
Figure 4. Electrochemical features of PLD-grown LCO thin films: specific discharge capacity and discharge mid-voltage vs. substrate temperature.
Figure 4. Electrochemical features of PLD-grown LCO thin films: specific discharge capacity and discharge mid-voltage vs. substrate temperature.
Coatings 09 00386 g004
Figure 5. (a) SEM image cross-section of a Li/amorphous Li3PO4/LCO thin-film microbatteries. (b) Discharge curves for various current densities in the voltage range of 3.0 to 4.5 V vs. Li+/Li. (Reproduced with permission from [88]. Copyright 2014 Elsevier).
Figure 5. (a) SEM image cross-section of a Li/amorphous Li3PO4/LCO thin-film microbatteries. (b) Discharge curves for various current densities in the voltage range of 3.0 to 4.5 V vs. Li+/Li. (Reproduced with permission from [88]. Copyright 2014 Elsevier).
Coatings 09 00386 g005
Figure 6. Phase diagram of microstructure development in PLD LiNi0.8Co0.2O2 films as a function of the growth temperature. (Reproduced with permission from [100]. Copyright 2006 American Chemical Society).
Figure 6. Phase diagram of microstructure development in PLD LiNi0.8Co0.2O2 films as a function of the growth temperature. (Reproduced with permission from [100]. Copyright 2006 American Chemical Society).
Coatings 09 00386 g006
Figure 7. The specific capacity of LiNi0.8Co0.15Al0.05O2 thin films deposited onto Ni foil, Si wafer, and ITO-coated glass as a function of the substrate temperature.
Figure 7. The specific capacity of LiNi0.8Co0.15Al0.05O2 thin films deposited onto Ni foil, Si wafer, and ITO-coated glass as a function of the substrate temperature.
Coatings 09 00386 g007
Figure 8. (a) SEM images of LiMn2O4 films deposited at Ts = 300 °C and Ts = 600 °C. (b) The grain size of PLD thin films as a function of Ts. (Reproduced with permission from [133]. Copyright 2007 Springer).
Figure 8. (a) SEM images of LiMn2O4 films deposited at Ts = 300 °C and Ts = 600 °C. (b) The grain size of PLD thin films as a function of Ts. (Reproduced with permission from [133]. Copyright 2007 Springer).
Coatings 09 00386 g008
Figure 9. (a) Surface morphology and (b) cross-sectional picture of PLD-grown LMO thin film deposited on an Si(001) substrate covered by a 0.2 µm SiO2 layer at 575 °C under a 13 Pa oxygen partial pressure. (Reproduced with permission from [135]. Copyright 2008 Elsevier).
Figure 9. (a) Surface morphology and (b) cross-sectional picture of PLD-grown LMO thin film deposited on an Si(001) substrate covered by a 0.2 µm SiO2 layer at 575 °C under a 13 Pa oxygen partial pressure. (Reproduced with permission from [135]. Copyright 2008 Elsevier).
Coatings 09 00386 g009
Figure 10. Cyclic voltammogram recorded at a 0.5 mV·s−1 scan rate of a Li/Li3PO4/LiCoMnO4 thin film battery. The LiCoMnO4 film cathode was grown under PO₂ = 100 Pa. (Reproduced with permission from [171]. Copyright 2014 Elsevier).
Figure 10. Cyclic voltammogram recorded at a 0.5 mV·s−1 scan rate of a Li/Li3PO4/LiCoMnO4 thin film battery. The LiCoMnO4 film cathode was grown under PO₂ = 100 Pa. (Reproduced with permission from [171]. Copyright 2014 Elsevier).
Coatings 09 00386 g010
Figure 11. Variation of the grain size in V2O5 thin films as a function of the substrate temperature. (Reproduced with permission from [205]. Copyright 2005 American Chemical Society).
Figure 11. Variation of the grain size in V2O5 thin films as a function of the substrate temperature. (Reproduced with permission from [205]. Copyright 2005 American Chemical Society).
Coatings 09 00386 g011
Figure 12. PO₂ vs. 1000/Ts phase diagram for films of vanadium oxides laser-pulse deposited on silicon substrates. (Reproduced with permission from [213]. Copyright 2015 AIP Publishing).
Figure 12. PO₂ vs. 1000/Ts phase diagram for films of vanadium oxides laser-pulse deposited on silicon substrates. (Reproduced with permission from [213]. Copyright 2015 AIP Publishing).
Coatings 09 00386 g012
Figure 13. Discharge profiles vs. lithium uptake for Li//V6O13 thin-film microbatteries. Active cathode films were grown with: (a) Ts = 250 °C, as-deposited; (b) Ts = 250 °C, annealed at 300 °C in Ar; (c) Ts = 25 °C, annealed at 300 °C in Ar.
Figure 13. Discharge profiles vs. lithium uptake for Li//V6O13 thin-film microbatteries. Active cathode films were grown with: (a) Ts = 250 °C, as-deposited; (b) Ts = 250 °C, annealed at 300 °C in Ar; (c) Ts = 25 °C, annealed at 300 °C in Ar.
Coatings 09 00386 g013
Figure 14. Temperature of the in-plane ionic for conductivity epitaxial Li0.33La0.56TiO3 solid electrolyte thin films (36-nm tick) deposited on NdGaO3(110) by PLD at Ts higher than 900 °C under PO₂ = 5 Pa. (Reproduced with permission from [272]. Copyright 2012 Elsevier).
Figure 14. Temperature of the in-plane ionic for conductivity epitaxial Li0.33La0.56TiO3 solid electrolyte thin films (36-nm tick) deposited on NdGaO3(110) by PLD at Ts higher than 900 °C under PO₂ = 5 Pa. (Reproduced with permission from [272]. Copyright 2012 Elsevier).
Coatings 09 00386 g014
Figure 15. SEM cross-sectional images of solid-state thin film lithium batteries, (a) SnO/LVSO/LCO and (b) SnO/LVSO/LMO. (Reproduced with permission from [121]. Copyright 2006 Elsevier).
Figure 15. SEM cross-sectional images of solid-state thin film lithium batteries, (a) SnO/LVSO/LCO and (b) SnO/LVSO/LMO. (Reproduced with permission from [121]. Copyright 2006 Elsevier).
Coatings 09 00386 g015
Figure 16. (a) SEM cross-section image of LTO film (410 nm thick) heat treated at 800 °C. (b) Charge–discharge profiles recorded at 20 µA cm−2 (i.e., ~1.15 C) current density in the voltage range of 1 to 2 V vs. Li+/Li of PLD films heated at various temperatures. (Reproduced with permission from [308]. Copyright 2009 Elsevier).
Figure 16. (a) SEM cross-section image of LTO film (410 nm thick) heat treated at 800 °C. (b) Charge–discharge profiles recorded at 20 µA cm−2 (i.e., ~1.15 C) current density in the voltage range of 1 to 2 V vs. Li+/Li of PLD films heated at various temperatures. (Reproduced with permission from [308]. Copyright 2009 Elsevier).
Coatings 09 00386 g016
Figure 17. Ragone plot (normalized by the active cell area) for lithium thin-film microbatteries fabricated with crystalline LiCoO2 (black lines), crystalline LiMn2O4 (blue lines), and nanocrystalline LixMn2−yO4 (red lines) cathode materials with different thicknesses (in µm).
Figure 17. Ragone plot (normalized by the active cell area) for lithium thin-film microbatteries fabricated with crystalline LiCoO2 (black lines), crystalline LiMn2O4 (blue lines), and nanocrystalline LixMn2−yO4 (red lines) cathode materials with different thicknesses (in µm).
Coatings 09 00386 g017
Table 1. Typical laser beams for PLD films of transition-metal (TM) oxides for energy storage.
Table 1. Typical laser beams for PLD films of transition-metal (TM) oxides for energy storage.
LaserWavelength (nm)Pulse Width (ns)Frequency (Hz)Laser Fluence (J·cm−2)Ref.
Excimer KrF24820112[10]
Excimer ArF1931052[11]
Excimer XeCl308[12]
Nd:YAG532 (a)8106[13]
Nd:YAG266 (b)1.6[14]
(a) Frequency doubled; (b) Fourth harmonic.
Table 2. Industrial developments of thin film microbatteries.
Table 2. Industrial developments of thin film microbatteries.
ManufacturerElectrochemical ChainSpecificationsRef.
Cymbet Co.EnerChipTM LiCoO2/Li60 µAh·cm−2·µm−1/5000 cycles[34]
Infinite Power SolutionsLiCoO2 or V2O5/LiPON/Li“Thinergy” 40 µAh·cm−2·µm−1[36]
Front Edge TechnologyLiCoO2/LiPON/Li“Nanoenergy” 0.9 mAh cm−2[37]
Ulvac Inc.LiCoO2/Li3PO4/Li50 µAh·cm−2·µm−1[38]
STMicroelectronicsLiCoO2/LiPON/Li700 µAh/discharge at 5 mA[39]
ExcellatronLiCoO2-LiMnO2-LiPON-Sn3N40.3 µm thick/0.1 mAh/2000 cycles[35]
EnfucellMnO2-based cellvoltage rating >3 V[33]
GS Caltexn/a300 µm thick/3.9 V/ 8000 cycles[40]
Table 3. Solid-state lithium microbatteries fabricated by the PLD technique.
Table 3. Solid-state lithium microbatteries fabricated by the PLD technique.
Electrochemical ChainCharacteristicsRef.
Li/Li3PO4/LiCoO29.5 µAh·cm−2[14]
Li/Li4SiO4/LiCoO210 µAh·cm−2[44]
SnO/Li6.1V0.61Si0.39O5.36/LiCoO29.5 Ah·cm−2 at 44 µA·cm−2[41]
In/80Li2S–20P2S5/LiCoO295 mAh·g−1 at 0.13 mA·cm−2[42]
Li/Li3PO4/LiMnPO412 mAh·g−1[45]
Li/LiPON/Li4Ti5O1232 µAh·cm−2 at 3.5 µA·cm−2[43]
Table 4. Characteristics of the oxide materials used as a positive electrode in Li batteries. Δxm is the quantity of electrons transferred (or Li uptake).
Table 4. Characteristics of the oxide materials used as a positive electrode in Li batteries. Δxm is the quantity of electrons transferred (or Li uptake).
TMO MaterialDensity (g/cm3)Cm (mAh g−1) for ΔxmCf (µA·cm−2·µm−1)Cf (mC·cm−2·µm−1)
V2O53.35294 (Δxm = 2)98.5354
MoO34.69279 (Δxm = 2)130.8471
LiCoO25.03 a137 (Δxm = 0.5)68.9248
LiNi0.5Co0.5O24.90 a274 (Δxm = 1)134.2483
LiMn2O44.32 a148 (Δxm = 1)64.0274
a after Ozuku, T.; Ueda, A. J. Electrochem. Soc. 1994, 141, 2972.
Table 5. Diffusion coefficients of Li+ ions in PLD LMO thin film frameworks.
Table 5. Diffusion coefficients of Li+ ions in PLD LMO thin film frameworks.
Growth ConditionsTs (°C)MethodDiffusion Coefficient (cm−2·s−1)Ref.
PO₂ =13 Pa800GITT2.5 × 10−11[47]
PO₂ = 13 Pa, P = 108 W·cm−2300GITT10−11–10−12[7]
PO₂ = 30 Pa, Φ = 1.0 mJ·cm−2600EIS/PITT/GITT10−11–10−14[151]
PO₂ = 26 Pa, Ppulse = 160 mJ400PITT10−10–10−12[152]
PO₂ = 20 Pa, Φ = 1.1 mJ·cm−2500Li radiotracer1.4 × 10−10 (300 °C)[153]
PO₂ = 0.2 Pa, Φ = 3.5 mJ·cm−2350Li radiotracer1.8 × 10−12 (350 °C)[154]
PO₂ = 10 Pa, Φ = 2.0 mJ·cm−2650Chronoamperometry8.4 × 10−10[155]
Table 6. Electrochemical properties of PLD-prepared LMO thin film electrodes. J is the current density, δ is the film thickness, and ΔCc is the capacity fading per cycle.
Table 6. Electrochemical properties of PLD-prepared LMO thin film electrodes. J is the current density, δ is the film thickness, and ΔCc is the capacity fading per cycle.
PLD Conditions Ts/PO₂Specific CapacityElectrochemical ParametersRef.
300 °C/13 Pa/0.5 J·cm−2120 mC·cm−2·µm−1J = 5 µA·cm−2; δ = 200 nm; ΔCc = 0.14%[7]
25 °C/0.2 Pa/3.5 mJ·cm−21.6 µAh·cm−2J = 4.4 µA·cm−2; δ = 100 nm; ΔCc = 0.4%[121]
500 °C/20 Pa/2 Jn/aΔCc = 0.05%/cycle[122]
500 °C/20 Pa/4.3 J·cm−242 µAh·cm−2·µm−1J = 13 µA·cm−2 (1C); δ = 0.3 µm; ΔCc = 0.025%[127]
13 Pa/2.3 J·cm−2130 mAh·g−1I =5 µA (3.3C); δ = 110 nm; ΔCc = 0.14%[129]
575 °C/13 Pa/160 mJ111 mAh·g−1J = 50 µA·cm−2; δ = 0.3 µm; ΔCc = 1.5%[135]
600 °C/30 Pa/1 J·cm−296 mAh·g−1J = 6 µA·cm−2; δ = 100 nm;[151]
600 °C/30 Pa/1 J·cm−2170 mC cm−2·µm−1I = 1 mA (180C); δ = 120 nm[156]
400 °C/26 Pa/160 mJ142 mAh·g−1J = 20 µA·cm−2; δ = 0.4 µm; ΔCc = 0.06%[157]
450 °C/26.7 Pa/4.6 J·cm−2120 mAh·g−1J = 330 mA·g−1; δ = 140 nm; ΔCc = 0.2%[163]
650 °C/2.6 Pa/457 mJ·cm−226 µAh·cm−2·µm−1J = 1.88 mA·cm−2 (348C); δ = 90 nm[164]
Table 7. Electrochemical properties of PLD-prepared vanadium oxide thin film electrodes. J is the current density, δ is the film thickness, and ΔCc is the capacity fading per cycle.
Table 7. Electrochemical properties of PLD-prepared vanadium oxide thin film electrodes. J is the current density, δ is the film thickness, and ΔCc is the capacity fading per cycle.
PLD Conditions Ts/PO₂Specific CapacityElectrochemical ParametersRef.
200 °C/2.6 Pa/12 J·cm−2200 mAh·g−1Capacity loss of <2% over 200 cycles[193]
300 °C/133 Pa/0.1 J·cm−2235 mC cm−2·µm−1J = 5 µA·cm−2; ΔCc = 1.1%[195]
100 °C/6 Pa/200 mJ240 mAh·g−10.1C current rate; δ = 650 nm; ΔCc = 0.1%[223]
200 °C340 mAh·g−1Capacity loss <2% after 100 cycles[201]
200 °C/10 Pa/2 J·cm−2340 mC cm−2·µm−1δ = 100 nm; 8000 cycles (electrochromic)[216]
Table 8. Electrical properties of PLD-grown solid-electrolyte thin films.
Table 8. Electrical properties of PLD-grown solid-electrolyte thin films.
ElectrolyteIonic Conductivity (S cm−1)Electronic Conductivity (S cm−1)Ref.
Li3.3PO3.9N0.17 (LiPON)1.6 × 10−6>10−14[259]
Li3.4V0.6Si0.4O4 (LVSO)2.5 × 10−77.4 × 10−13[41]
Li3.25Ge0.25P0.75S4 (thio-LISICON)2.2 × 10−31.5 × 10−7[260]
Li0.5La0.5TiO3 (LLTO)2.2 × 10−53.5 × 10−11[261]
Li2O-Si2O (LSO)2.2 × 10−4 (200 °C)[262]
β-LiAlSiO44.0 × 10−7 (225 °C)[263]
Li6BaLa2Ta2O12 (garnet-type)2.0 × 10−6 (25 °C)2.9 × 10−13[264]
Table 9. Literature review of garnet-like solid electrolyte thin films grown via PLD.
Table 9. Literature review of garnet-like solid electrolyte thin films grown via PLD.
MaterialSubstrateσi at 25 °C (S cm−1)Ea (eV)Ref.
Li6BaLa2Ta2O12MgO(100)2 × 10−6 at 25 °C0.42[278]
Li7La3Zr2O12Si3N4/Si6.3 × 10−3 at 400 °C0.60[279]
Li7La3Zr2O12SrTiO3, sapphire7.4 × 10−7 at 25 °C0.32[280]
Li7La3Zr2O12Si, SiO2, MgO1.6 × 10−6 at 25 °C0.35[281]
Al-doped Li7La3Zr2O12Gd3Ga3O122.5 × 10−6 at 25 °C0.52[275]
Al-doped Li7La3Zr2O12MgO8.3 × 10−4 at 300 °C0.60[264]
Table 10. Growth conditions and electrochemical properties of new anode materials.
Table 10. Growth conditions and electrochemical properties of new anode materials.
Anode MaterialTs (°C)Ambient (Pa)Laser Energy (J cm−2)Electrochemical PropertiesRef.
C-monocolumn25(N2) 6.70.08315 mAh·g−1 at 0.1 A·g−1 (50th cycle)[342]
NiFe2O4650(O2) 2.02.5370 mAh·g−1 at 25 µA·cm−2[343]
Co2O3700(O2) 1002.0750 mAh·g−1 at 2C (350th cycle)[344]
Cu2Sb25(Ar) 1.33–4>135 mAh·g−1 at 35 µA·cm−2 (100th cycle)[345]
Fe2O325(O2) 0.30.5905 mAh·g− 1 at 0.1 A·g− 1 (200th cycle)[346]
Co3O4600(O2) 402600 mAh·g−1 at 10 µA·g− 1 (2th cycle)[347]
Co3O4/Co(OH)2250(O2/H2) 30360 mAh·g−1 at 32 A·g− 1 (1000th cycle)[348]
In2O3200(O2) 52500 mAh·g− 1 at 10 µA·g− 1 (30th cycle)[349]
Sb2Se3200(Ar) 52530 mAh·g− 1 at 5 µA·g− 1 (100th cycle)[350]
FeOF25(Ar) 52110 mAh·g− 1 at 5 µA·g− 1 (100th cycle)[351]
SrLi2Ti6O14700(O2) 22130 mAh·g− 1 at C/20 (40th cycle)[352]
MnO500(Ar) 205425 mAh·g− 1 at C/8 (25th cycle)[352]
VP2400(Ar) 102987 mAh·g− 1 at 5 µA·cm−2 (50th cycle)[354]

Share and Cite

MDPI and ACS Style

Julien, C.M.; Mauger, A. Pulsed Laser Deposited Films for Microbatteries. Coatings 2019, 9, 386. https://0-doi-org.brum.beds.ac.uk/10.3390/coatings9060386

AMA Style

Julien CM, Mauger A. Pulsed Laser Deposited Films for Microbatteries. Coatings. 2019; 9(6):386. https://0-doi-org.brum.beds.ac.uk/10.3390/coatings9060386

Chicago/Turabian Style

Julien, Christian M., and Alain Mauger. 2019. "Pulsed Laser Deposited Films for Microbatteries" Coatings 9, no. 6: 386. https://0-doi-org.brum.beds.ac.uk/10.3390/coatings9060386

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