2D-Layered Lithium Carboxylate Based on ... - ACS Publications

Dec 16, 2016 - Laboratoire de Réactivité et Chimie des Solides (LRCS), UMR CNRS 7314, Université de Picardie Jules Verne (UPJV), Amiens, 33...
0 downloads 0 Views 2MB Size
Subscriber access provided by UNIV OF NEWCASTLE

Article

2D-layered lithium carboxylate based on biphenyl core as negative electrode for organic lithium-ion batteries Lionel Fédèle, Frédéric Sauvage, Sébastien Gottis, Carine Davoisne, Elodie Salager, Jean-Noël Chotard, and Matthieu Becuwe Chem. Mater., Just Accepted Manuscript • DOI: 10.1021/acs.chemmater.6b03524 • Publication Date (Web): 16 Dec 2016 Downloaded from http://pubs.acs.org on December 19, 2016

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

Chemistry of Materials is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

2D-layered lithium carboxylate based on biphenyl core as negative electrode for organic lithium-ion batteries Lionel Fédèlea,b,d, Frédéric Sauvagea,b,d, Sébastien Gottisa,b,d, Carine Davoisnea,b,d, Elodie Salagerc,d, Jean-Noël Chotarda,b,d, Matthieu Becuwe*a,b,d (a) Laboratoire de Réactivité et Chimie des Solides (LRCS), UMR CNRS 7314, Université de Picardie Jules Verne (UPJV), Amiens, 33 rue Saint-leu, 80039 Amiens, France (b) Institut de Chimie de Picardie (ICP), FR CNRS 3085, Amiens, France (c) CNRS Conditions Extrêmes et Matériaux : Haute Température et Irradiation (CEMHTI), UPR 3079, Univ Orléans, 1D avenue de la Recherche Scientifique, 45071 Orléans, France (d) Réseau sur le Stockage Electrochimique de l’Energie (RS2E), FR CNRS 3459, France

Table of content

Abstract A 2D-layered conjugated organic lithium carboxylate based on 4,4’-biphenyl core has been synthesized and characterized by combining powder x-ray diffraction, two-dimensional solidstate

13

C NMR and high-resolution transmission electron microscopy. The combination of

these complementary technics are identifying both molecular packing and structural organization within the solid. The electrochemical performances of this new organic negative electrode have been evaluated. The specific 2D-layered offered by the 4,4’-biphenyl

ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

framework confers to the material an unprecedented rate capability, maintaining as high as 180 mAh.g-1 over 25 cycles at 2e- / hour rate without any means of electrode formulation. Such electrochemical performances are ascribed at once to the enhancement of lithium conduction inside the mesoscopic crystal structure, promoted by the greater distance between the inorganic LiO4 layers and to the greater structural robustness upon cycling sustaining no structural amorphization during lithium insertion by contrast to the other members of the series of carboxylate-based organic electrodes.

Keywords: 2-D layered Organic anode material, lithium-ion battery, lithium carboxylate, high rate capability

Introduction Environmental considerations are now at the center of our prerogative to prompt the ecoconception and the design of new materials able to efficiently store electrical energy from the grid or renewable resources. It requires the emergence of new paradigm on sustainable materials and devices dedicated to renewable energy production and storage.1 One main stream, initiated for a couple of years to limit the environmental outcome from the synthetic chemistry activities, is to promote the “chimie douce” to bring synthetic temperatures closer to

2–4

or to room-temperatures.5–8 Another illustration stems from the bio-assisted synthesis

which have demonstrated their beneficial impact not only in terms of low-temperature transformation but also on electrochemical performances when included into lithium-ion batteries thanks to the nano-texturation of the material formed.9–12 More recently, organic materials have gained credit to replace the inorganic electrodes which are currently built around a 3d transition metal associated to an oxide or a polyanionic framework.13–15 Organic electrodes promise to combine eco-conception from the cradle-to-grave, lower environmental

ACS Paragon Plus Environment

Page 2 of 26

Page 3 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

repercussion, limit metal depletion and decrease of weight and cost of full assembly through the replacement of actual copper current by aluminum,16 a fully recyclable element. It also fully exploits the richness of its chemistry to design the materials of tomorrow.

17–21

Nevertheless, regardless of whether it is as a positive or a negative electrode, challenges remain to convey their performances and stability to the current level of the benchmark LiCoO2 or LiFePO4/graphite. One first issue, already addressed by the community, is the solubility of the organic positive electrode based on quinone,22,23 TCNQ,24 nitroxide,25 or phenothiazine derivatives26 into the electrolyte. A solution to this impediment has been proposed by replacing the benchmark liquid electrolytes based on cyclic and linear alkyl carbonates by gel polymer27,28 or by integrating a functional group into the molecular backbone to render it insoluble; an example emphasizing on the organic chemistry richness.29– 31

A second important drawback lies in the lack of electrical conductivity requiring high

carbon content and specific formulation steps to ensure efficient electronic percolation across the electrode.32,33 For carboxylate-based negative electrode, such as the firstly proposed dilithium terephthalate,34 a drastic amelioration of the power properties can be obtained by extending the π-conjugation in the central core unit through naphtalen35 and then perylen moieties.36 These studies contributed to reveal one role played by the central core unit other than simply a spatial spacer between the redox active carboxylate groups. For this sake of both understanding and designing superior materials, this work continues to focus on the central core unit in which we introduced successfully a 4,4’-bi-phenyl unit leading to the dilithium 4,4’-BiPhenylDiCarboxylate (Li2-BPDC). This new molecule affords to increase the separation distance of the two end carboxylate units compared to the closely-related 2,6naphthalene ande 1,4-phenyl counterparts (Scheme 1). The results, discussed in details, give credit to the importance to spatially separate the two electrochemically active units where Li2-

ACS Paragon Plus Environment

Chemistry of Materials

BPDC offers an improved rate capability compared to the closely-related shorter linear di-

10.30 Å

8.314Å

lithium carboxylate-based molecules.

5.69Å

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 26

Scheme 1. Evolution of the distance between the two-end carboxylates in different centro-symmetric aromatic organic carboxylic acid used as a spacer in metal-organic structure (distances determined using Avogrado molecular editor).

2. Materials and methods 2.1 Reagents and Chemicals 4,4’-biphenyldicarboxylic acid (4,4’-BPDCA) and lithium hydroxide monohydrate (LiOH.H2O) were purchased from Alfa Aesar and Sigma-Aldrich, respectively. All the solvents used in this study were purchased from Fischer scientific. The deuterated dimethylsulfoxide (DMSO-d6) were purchased from Euriso-top. All solvents and reagents were used as received. 2.2 Instrumentation X-Ray Powder Diffraction patterns (XRPD) were acquired using a Bruker D8 diffractometer equipped with a Cu anti-cathode (Kα radiation, operating at 40 kV–40 mA). Patterns were collected in the 2θ range of 10–60° with a step size of 0.03°. Indexing of the powder pattern was performed using DICVOL37 and the final refinement was done using the Rietveld method as implement in the Fullprof Suite package.38 The particles morphology and their size were estimated by Scanning Electron Microscopy (SEM) using an environmental

ACS Paragon Plus Environment

Page 5 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

FEI Quanta 200 FEG-microscope. The microstructural and structural investigations were performed using a Tecnai F20-STWIN through bright field imaging, high resolution transmission electron microscopy (HRTEM) and selected area electron diffraction (SAED). The solid-state

13

C-NMR experiments were carried out at room temperature on a Bruker

Avance I wide bore spectrometer operating at a 1H Larmor frequency of 300 MHz, with a Bruker 4 mm double resonance CPMAS probe-head. The magic angle spinning rate was 12.5 kHz. For the 1D 1H-13C CPMAS spectrum 408 transients were acquired with acquisition times of 18 ms, using a recycle delay of 10 s, a CP contact time of 3.5 ms and a SPINAL-64 hetereonuclear decoupling at a 1H radio-frequency field of 90 kHz. The total acquisition time was 1 hour. For the 2D refocused

13

C-13C CP-INADEQUATE spectrum, similar conditions

were used. Sixty-four t1 points were acquired using STATES with 1536 transients each. The maximum t1 delay was 5.1 ms. Delays for the conversion to and from the double-quantum coherence were 4 ms. The total acquisition time was 11.5 days. The processing was performed so as to display a single quantum-single quantum correlation made using a homemade script. The quantitative

13

C spectrum was recorded using 60 transients in 8 days. The

excitation pulse was set to π/8 to reduce the recycle time to 11875 s. All 13C chemical shifts presented in this manuscript were referenced with respect to the CH2 resonance of adamantane at 38.48 ppm with respect to neat TMS. Electrochemical performances were determined using 2 electrodes Swagelok-type cell assembled in an argon-filled glovebox. For this, a typical working electrode was composed of ca. 10 mg of active material (60%) manually mixed with 40 % of Super P carbon. It was separated from a lithium foil playing both the role of counter and reference electrode by two Whatman fiberglass sheets soaked with LiPF6 (1M) in ethylene carbonate-dimethyl carbonate (EC-DMC) 1/1 (v/v) mixture (LP30 electrolyte, certified battery purity grade, Merck). The different galvanostatic

ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 6 of 26

measurements were recorded between 3 V and 0.4 V (vs. Li+/Li) using a Biologic VMP potentiostat/galvanostat.

2.3 Dilithium salt synthesis To a suspension of 0.5 g of 4,4’-biphenyl dicarboxylic acid (4,4’-BPDCA) in 20 mL of methanol was added 3 equivalents of LiOH. The mixture was then heated under stirring at reflux for overnight to yield to the crystallization of the di-lithium salt as a white powder (yield 82 %). The solid was retrieved by centrifugation and washed successively with pure methanol, acetone and diethylether. The powder was finally dried under vacuum at 80°C and transferred into an argon-filled glovebox for storage. 1H liquid NMR, δ in ppm (300MHz; DMSO-d6): 8.07 (4H, d, J=8Hz, H-3, H-3’, H-5, H-5’), 7.88 (4H, d, J = 8Hz, H-2, H-2’, H-6, H-6’). The surface area of the powder is 10.5 m2.g-1 based on nitrogen adsorption (CBET = 86). Its lithium stoichiometry was determined to 1.98 lithium (+/- 0.03) per formula unit by atomic absorption spectrometry. Water content in the obtained material was estimated to ca. 21 ppm using Karl Fischer titration method.

3 LiOH.H20 (3eq.)

Li+ MeOH

-

2

2’ 1

4 5

6

3’ 4’

1’ 6’

-

Li+

5’

Scheme 2: Synthesis of dilithium 4,4’-BiPhenyl DiCarboxylate (Li2-BPDC).

3. Results and discussion The organization of molecular packing in Li2-BPDC was evidenced by powder x-ray diffraction. After lithiation, the four intense peaks at 16.8°; 24.3°; 25.3° and 26.8° in the acid form of BPDC (ESI figure S1) are entirely disappearing in favor to new intense diffraction peaks at 20°; 20.8°; 21.7°; 22.6° and 28.6° attributed to the lithiated salt Li2-BPDC (Figure 1a).

ACS Paragon Plus Environment

Page 7 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Figure 1: (a) Refined powder X-ray diffraction pattern of crystallized Li2-BPDC with Scanning Electron Micrograph in inset and (b) illustration of the crystal structure of Li2-BPDC 2D-network. (c and d) HighResolution Transmission Electron Microscopy micrograph of the sample before electron irradiation. (c) HRTEM micrograph of the sample before electron irradiation with in inset from right to left the associated electron diffraction pattern along the [3-41] zone axis and an enlarged filtered view of the plane layout and (d) the same area after electron irradiation with the white arrow showing the preferential loss of matter and in inset the crystal structure projection according the [3-41] zone axis.

Li2-BPDC crystallizes within a monoclinic lattice cell with P21/c space group. Rietveld refinement yields to lattice cell parameters of a = 12.7977(16) Å, b = 5.1453(7) Å, c = 8.6293(11) Å and β = 96.425(7)°. These values are in good agreement with those reported in the literature by Parise et al.39 As it is commonly observed for crystallized organic lithiumcarboxylate materials40,41, the crystal structure of Li2-BPDC can be described as a twodimensional layered Lithium Organic Framework (LOF) where in this case the biphenyl units are separated by inorganic sheets of LiO4. In these layers, LiO4 tetrahedra are joined together

ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

via one edge, forming Li2O6 entities, these latter sharing their corners to form an infinite layer along (100) plane. The crystal structure shows that each lithium is coordinated by four carboxylates and vis-et-versa. The LiO4 layers are linked together, similarly to a traditional 3D-metal organic framework, through 4,4’-biphenyl dicarboxylate unit giving an inter-layer distance (Lithium-Lithium) of 12.79 Å, smaller than the one expected of 13.38 Å (Figure 1b). This difference stems from the biphenyl units which are tilted by 74° with respect to the LiO4 layers giving an intrinsic distance between carbon from the carboxylate of 9.84 Å instead of 10.30 Å (Figure 1b). This tilt induces a loss of molecular plan symmetry while preserving the centro-symmetry of the structure by an inversion center; centro-symmetry seemingly required in carboxylate-based electrode to maintain a good reversibility in the lithium insertion / deinsertion process.42 As a result, the structure bears seven types of non-equivalent carbon as shown in the inset of figure 2a. By contrast to the pristine acid form for which the particles morphology is sub-micrometric without clearly defined morphology and heterogeneously distributed in size (ESI Figure S1), the lithiated salt adopts a kind of layered morphology with particles forming platelets or rods of low aspect ratio, being representatively less than 1 µm wide and few micrometers long (SEM micrograph inset figure 1a). Such morphology and crystallinity can be also well visualized by TEM (Fig. 1c). The sample is oriented according to the [3-41] zone axis (right inset in figure 1c). The interplanar spacing dhkl = 4.04 Å is in agreement with the (111) plane (left inset in figure 1c).The increasing dose of electron to the sample induces the creation of whiter area in the sample, as indicated by the series of arrows in the Figure 1d. They are all parallel to each other and have an angle of ~65° with respect to the (111) plane, signifying a braking and preferential loss of matter along such direction which corresponds to the [010] direction (see inset in figure 1d). This orientation is not totally surprising since the electron irradiation will favorably break weaker bonds, in the present case the Li-O from the

ACS Paragon Plus Environment

Page 8 of 26

Page 9 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

tetrahedral framework. It induces then a loss of the Li-O which consequently leaves the carbon framework free to move out from the sample. Consistently with x-ray diffraction which shows the complete disappearance of the diffraction peaks related to the acid, complementary FT-IR and NMR spectroscopic studies were investigated to ensure the quantitative deprotonation of the acid in favor to the lithiation (ESI Figure S2). An important effort has been herein led to resolve the NMR peak assignment for an accurate characterization of the atomic structure of the molecular material. To the best of our knowledge, this approach was never undertaken so far for organic-based Li-ion electrode material. Specific NMR sequences were used to assign precisely all the carbon signals. In a first step, a quantitative spectrum based on natural abundance of 13C was acquired to estimate the number of carbon atoms contributing to each peak (Figure 2a). Owing to the symmetry of the molecule, we would expect carbons 2 and 6 to be equivalent and give raise to a single peak of relative intensity 2, as well as carbons 3 and 5. Besides those two peaks, we expect three more peaks of intensity 1, respectively, for carbons 1, 4 and 7. So a liquid-state NMR spectrum would exhibit five peaks (neglecting J-splitting) with relatives signal intensities of 2:2:1:1:1 In the solid-state, the equivalence between carbon 3 and 5 is broken by molecular packing, as evidenced by X-rays (7 non-equivalent carbons). However, five peaks are observed on the CPMAS spectrum (ESI Figure S2c). Integration obtained from the quantitative spectrum (figure 2a) also indicates five main peaks with intensity ratios 1.4:1:2:1:2, in good agreement with what is expected if the symmetry was conserved in the crystal. A broad peak (carbon black in the electrode formulation) and small contributions from potential decomposition/amorphization of the material are also detected in the quantitative spectrum. The 1D spectrum, in disagreement with the X-ray data, prompted us to perform a thorough assignment of the carbon spectrum. This approach, following the NMR crystallography framework,43–45 is performed here for the first time on organic-based Li-ion

ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 26

electrode material. A solid-state NMR 13C-13C CP-INADEQUATE experiment 46 is combined with a periodic DFT calculation of the carbon chemical shifts on the solid-state structure of Li2BPDC (Figure 2b).

Figure 2: (a)

Natural abundance

13

C quantitative spectrum of Li2-BPDC and corresponding

deconvolution. (b) Two-dimensional solid-state NMR spectrum of Li2-BPDC material using

13

C-13C CP-

INADEQUATE single-quantum/single-quantum plot (Double-quantum/single-quantum available in ESI figure S3). (c) Calculation of the

13

C NMR chemical shift in crystallized Li2-BPDC using the CASTEP

program and comparison with experimental values.

We recall that the INADEQUATE experiment uses the J-coupling between pairs of carbons13 to create double-quantum coherences. As a result, a cross-peak appears between peaks that correspond to directly bonded carbons. The assignment is performed starting from peak A (178 ppm) assigned to carboxylate 7 owing to the specific chemical shit of this function. The A-C correlation indicates unambiguously that peak C corresponds to carbon 4 directly bonded to 7. Contrary to what suggested the 1D spectrum, C is also correlated to two other peaks, E and D. This result highlights a loss of isochrony between carbons 3 and 5 that would be

ACS Paragon Plus Environment

Page 11 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

equivalent in the molecule but not anymore in the crystal due to the molecular packing, in agreement with the X-ray diffraction results. One would be tempted to assign carbon 3 to peak E and carbon 5 to peak D (or vice versa). The D-E cross-peak casts doubt on this interpretation as carbon 3 and 5 are not bonded. We can therefore conclude that peak C or peak E result from the accidental overlap of two non-equivalent carbons. At this point, calculations are needed to assist the precise assignment of the carbon spectrum. Periodic DFT calculations of the chemical shifts in the solid-state47–49 were conducted with the CASTEP software on the structure provided from the X-ray diffraction study (Procedure details available in ESI). The calculated chemical shifts for the 7 non-equivalent carbons in the crystallographic cell are reported in figure 2c. The peaks already assigned using the experimental INADEQUATE are well estimated with a small constant shift. Peak E stems from carbon 2 and carbon 6, as expected for similar bonding, while peak B arises from carbon 1. More interestingly, the calculation lifts the veil on peak C that contains contributions from carbon 4 and carbon 5. These two carbons, although not equivalent in the molecule, have similar environments in the crystal and therefore similar chemical shifts. Finally, the assignment for carbons 3 and 5 can now be refined, with carbon 3 assigned to the peak at 131.8 ppm (peak D) and carbon 5 to the peak at 134.3 ppm (peak C). The INADEQUATE experiment confirms the results of the calculation. Cross-peaks between peaks C and D (C4-C3 bond), peaks D and E (C3-C2 bond), peaks E and B (C2-C1 bond), peaks B and E (C1-C6 bond) and peaks E and C (C6-C5) all confirm the calculation. Note that we don’t get a cross-peak for the C4-C5 bond, most probably because of their extremely close chemical shifts. This type of experiment, performed for the first time on battery material, opens new perspectives of investigations which could, in conjunction with other high performance tools like FT-IR or EPR, allow to identify intermediate species formed during electrochemical

ACS Paragon Plus Environment

Chemistry of Materials

process

and

to

gain

better

insight

on

the

exact

mechanism

during

lithium

insertion/desinsertion. As commonly intuited, the reactivity of centro-symetrical dilithium carboxylate is based on the double insertion of electron on carbonyl group inducing generation of a negative charge and a radical species. It is followed by an uptake of two lithium by formula unit (one by lithium carboxylate) for charge compensation in the carbonyl. In the case of Li2-BPDC, this two-electron process offers a theoretical gravimetric capacity of 211 mAh.g-1. This electrochemical reaction induces a swaying of electronic density leading to the formation of a carbonaceous quinonic-type entity as represented in scheme 3. -

Li+

- Li

+1e -, +1Li+

Li+

+

-1e-, -1Li+

.

-

+1e-, +1Li+ -1e-, -1Li+

Scheme 3: Expected electrochemical lithium insertion/desinsertion process in Li2-BPDC.

The electrochemical performances of Li2-BPDC were evaluated in galvanostatic mode at different rates from 0.2 e-/h to 2 e-/h (Figure 3). 3

a)

3

b)

0.2e-/h

0.5e-/h 2.5

Potential (V)

Potential (V)

2.5 2 1.5

1.5

0.5

0.5 0

0.5

1

1.5 2 2.5 3 3.5 e- in Li2-C14H8O4

4

0

0.5

1

1.5

2

2.5

3

3.5

4

e- in Li2-C14H8O4

3

c)

2

1

1

d)

1e-/h 2.5

3

2e-/h 2.5

Potential (V)

Potential (V)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 12 of 26

2 1.5 1

2 1.5 1

0.5

0.5 0

0.5

1

1.5

2

2.5

3

e- in Li2-C14H8O4

3.5

4

0

0.5

1

1.5

2

2.5

3

e- in Li2-C14H8O4

ACS Paragon Plus Environment

3.5

4

Page 13 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Figure 3: Galvanostatic charge/discharge curves between 0.4 - 3.0 V (vs. Li+/Li) for the ten first cycles of Li2-BPDC mixed with 40% of carbon Super P at a regime of 0.2e-/h (a) , 0.5e-/h (b) , 1e-/h (c) , 2e-/h (d) rate.

Li2-BPDC inserts lithium at a reduction potential of 0.7 V vs Li+/Li. This is comparatively lower than the di-lithium terephthalate and related counterparts owing to the π-conjugation enhancement contributing to the energetic destabilization of LUMO orbitals of the molecule.50–53 The total capacity of the first discharge shows a greater value (3.3 electrons per formula unit) than the expected theoretical value of 2. This additional capacity stems mainly from the formation of the SEI layer. However, we argue that it also results from the electrode formulation since it can be reduced by lower carbon content and even avoided by a specific electrode formulation.33 After this first discharge, Li2-BPDC displays reversibly full theoretical capacity, ca. 2.1 e- extracted corresponding to a gravimetric capacity of 221 mAh.g1

. This electro-activity is associated to a low polarization value of 50 mV at 0.2 e-/h rate. It is

remarkable to underline that increasing this rate by ten-fold (2 e-/h), it has relatively low impact on both the polarization value and on the reversible capacity obtained. Interestingly, a peculiar phenomenon of this material, a bump at the end of the oxidation is systematically experienced regardless of the rate or on the number cycles. Since this phenomenon occurs independently on whether lithium or sodium cation is used,54 we argue that this occurrence stems from the de-intercalation of solvent molecules from the structure, in analogy with such a similar behavior encountered on graphite-based electrode in which the solvent co-intercalates/de-intercalates between the graphene sheets.55 This co-insertion of solvent molecules in the case of Li2-BPDC does not totally come as a surprise since the important distance between the LiO4 layers (as it will be discussed in the following) which can easily accommodate the Li+ cation with its solvation shell. In addition, if similarly to the graphite, a greater mobility of the solvated lithium cation inside the microporous network of

ACS Paragon Plus Environment

Chemistry of Materials

Li2-BPDC is expected.56,57 This could explain the superior rate capability of this material as it is discussed below. These electrochemical performances can be maintained at least over 25 cycles regardless of the regime as illustrated in Fig. 4 (thus on the resident time in the cell, see ESI Fig S4). b) 400

c) 1200

100

1400

95

300 90

250 200

85

150

80

100 75

50

1000

1200

0 800

2 600 400

-Im(Z)/Ohm

350

-Im(Z)/Ohm

Gravimetric capacity (mAh.g-1)

a)

Coulombic efficiency (%)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 26

1000 800 600 400

1 200

3

200

2

Oxidized 0

0

5

10

15

20

70 25

0

0

0

200

Cycle number

400

600

800

1000

1200

Reduced

1 0

500

1000

1500

Re(Z)/Ohm

Re(Z)/Ohm

Figure 4: (a) Evolution of the gravimetric capacity of Li2-BPDC as a function of the number of cycles at a rate of 0.2 e-/h (black curve), 0.5 e-/h (green curve) , 1 e-/h (blue curve) , 2 e-/h (red curve). Evolution of Nyquist plots of Li2-BPDC electrode/Li cells upon cycling at 1Li+/5h rate at oxidized and reduced state (b and c respectively).

The evolution of the gravimetric capacities systematically remain closed to the theoretical value (Cth) after 25 cycles: 200 mAh.g-1 (94.8 % of Cth), 183.4 mAh.g-1 (87.0 % of Cth), 186.5 mAh.g-1 (88.4 % of Cth) and 182.2 mAh.g-1 (86.4 % of Cth) at a cycling rate of 0.2 e-/h, 0.5 e/h, 1 e-/h and 2 e-/h, respectively. This denotes the robustness of the electrode material during lithium intercalation / de-intercalation processes despite a modification of charge transfer resistance upon cycling more important at oxidized state (Figure 4 b and c). This reasonable increase suggests that interfacial reactions occur (for both lithium metal and composite) thus hampering lithium insertion / de-intercalation. The

advantage

of

Li2-BPDC

compared

to

Li2-NDC

(di-lithium

2,6-naphthalene

dicarboxylate)35 or Li2-BDC (di-lithium 1,4-benzene dicarboxylate)34 can be seen in Figure 6. Rigorously compared under equivalent conditions, namely 40% Csp, same electrode area and similar loading, Li2-BPDC shows superiority in terms of electrochemical capacity, a lower

ACS Paragon Plus Environment

Page 15 of 26

polarization (100 mV in these conditions) and cycling performances (Fig. 5a-b). By comparison, the phenyl core unit can uptake 0.7 lithium / formulation unit after the second discharge whereas it increases to 1.6 lithium for the naphthalene counterpart. Li2-NDC exhibits a reversible capacity of 1.5 lithium cation (Q = 176 mAh.g-1) with a higher polarization than Li2-BPDC of 380 mV. As a result, Li2-BPDC maintains an unprecedented gravimetric capacity for an organic-based carboxylate material, namely 183 mAh.g-1 (87% of theoretical capacity) after 25 cycles at 2 e- / h rate. For comparison, Li2-BDC and Li2-NDC are maintaining capacities of only 54 mAh.g-1 (18 % of theoretical capacity) and 134 mAh.g-1 (57% of theoretical capacity), respectively. Note that for this comparison, the surface area for the different materials is also relatively comparable, 10.5 m².g-1 for Li2-BPDC, 13.2 m².g-1 for Li2-NDC, and 14.7 m².g-1 for Li2-BDC based on nitrogen physisorption.

Gravimetric capacity (mA.h.g-1)

3

Potential (V)

2.5

2

1.5

1

0.5

300

100

250

80

200 60 150 40 100

0 0 0.5 1 1.5 2 2.5 e- in Li2-BDC e- in Li2-NDC e- in Li2-BPDC

20

50

2e-/h 0

5

10

15

20

Coulombic efficiency (%)

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

0 25

Cycle number

Figure 5: (a) Galvanostatic charge/discharge curves between 0.4 V and 3.0 V (vs. Li+/Li) during second cycle and (b) Evolution of gravimetric capacity in discharge over 25 cycles (full line) including the corresponding coulombic efficiency (dash line) for Li2-BDC (blue curve), Li2-NDC (green curve) and Li2BPDC (red curve) at 2e-/h rate.

Obviously, the electronic delocalization in the core unit plays a role to enhance the power rate capability of the material as discussed previously.35,36 However, beside this π-conjugation enhancement, we also argue that the size itself of the organic spacer, which modifies the distance between the electro-active carboxylate units, plays a noticeable role to improve the

ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

electrode performances by improving lithium-ion conduction within the structure. This is supported by the Figure 6 comparing the three crystallographic structures in the (100) plane. All materials are crystallizing in a low symmetry monoclinic cell of space group: P21/c. The structures can be described similarly as a 2D-layered structure consisting of LiO4 zig zag slabs separated by the organic core unit. Although disconnected along [100] direction, this later provides an electronic link between the two redox units. It also plays the role of spacer between the inorganic layers (Figure 6a). Going from phenyl, naphthyl to biphenyl core unit, the inter-layer distance increases from 8.36 Å, 10.31 Å to 12.79 Å, respectively. The angle formed between the organic spacer and the LiO4 layer goes from 63.9°, 66.3° to 73.5° from benzene, naphthyl to biphenyl, respectively. Consequently, the theoretical crystal density decreases from 1.622, 1.598 to 1.494 coming with a free space increases from 79.1 %, 80.6 to 82.3 % for Li2-BPDC (Fig. 6b).

ACS Paragon Plus Environment

Page 16 of 26

Page 17 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

Figure 6: (a) Crystal structure comparison of Li2-BDC, Li2-NDC and Li2-BPDC in the (100) plan (top) and along the a axis (bottom). (b) Evolution/comparison of crystallographic parameters as function of spacer size from benzene ring to biphenyl (note Z = 2 for all materials).

In addition to the π-conjugation enhancement, the distance between inorganic layers are also responsible for the improvement of the electrochemical properties versus lithium and the enhancement of the power rate capability of the material. Solvent intercalation in Li2-BPDC is the result from this noticeable increase of volume cell, interlayer distance and free space. In situ / in operando X-ray diffraction experiments have been carried out to gain better insight

on the impact of the lithium insertion (and de-insertion) on the evolution of the crystal structure. The series of diffractograms collected at a rate of 0.05 lithium / hour is gathered in Figure 7. Commonly to the carboxylate-based organic materials investigated so far,34 the reduction of Li2-BPDC (from pattern 1 to 49) occurs via a biphasic reaction between Li2BPDC and Li4-BPDC. Indeed, we observe the gradual fading of the Li2-BPDC reflections while new sets of reflections appears. However, by contrast to other carboxylate-based organic materials for which a strong amorphization was experienced,

29

the discharged Li4-

BPDC phase (pattern 49) remains crystallized until 0.5 V (vs. Li+/Li) (Figure 7a and 7b). Upon recharge (from pattern 50 to 75), Li2-BPDC is reformed and the pattern obtained at the end of the charge (pattern 75) is perfectly matching with the pristine materials, demonstrating the full reversibility of the process. Moreover, no significant peaks’ broadening, a common observation for the other materials from the series, was observed. The absence of amorphization and electrochemical milling of the crystallites in the case of Li2-BPDC suggests the excellent flexibility of the crystal structure which can accommodate the structural strength of the subsequent Li4-BPDC / Li2-BPDC nucleation.

ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 18 of 26

Figure 7: (a and b) In situ powder X-ray diffraction realized on the second discharge of Li2-BPDC mixed with 40% of Csp at 0.05e-/h rate between 0.5 to 3V (vs. Li+/Li). Full acquisition (one pattern by hour) available in ESI figure S5. In the x-ray powder patterns, the reflection marked with a * is due to the Beryllium used as a current collector in the in situ cell. (c) 2D contour view of the in situ X-ray diffraction during the second cycle between 20 and 23°.

Looking closely to the in situ experiment on the 2D contour view, one can observe the disappearing of three peaks at 20.1°, 21.7° and 22.5° to the benefit of new ones at 20.6°, 21.9° and 22.2° at the end of reduction. Interestingly, one can also distinguish a new peak appearing exactly between Li2-BPDC and Li4-BPDC at 21.3°. We think that lithium reduction from Li2BPDC to Li4-BPDC proceeds with the formation of an intermediate phase of composition Li3BPDC. This intermediate phase, not observed so far on the other materials of the series, may also aid to better accommodate the structural strain until the fully discharge composition.

4. Conclusions This work describes the synthesis and characterization of a new biphenyl-based lithium carboxylate by combining powder x-ray diffraction, solid-state

13

C NMR and

transmission electron microscopy. For first time dedicated to an analysis of battery material, NMR 13C-13C CP-INADEQUATE enabled unambiguous assignment of the carbon signals to the crystal. This opens new perspectives, which are under development in our lab, to better

ACS Paragon Plus Environment

Page 19 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

understand the electrochemical mechanism of organic electrodes and to identify possible intermediate phases formed upon cycling. The integration of a larger organic spacer compared to terephthalate- and naphthalene-based materials offers enhancement in the electrode’s rate capability without any means of specific electrode formulation. It also provide a more flexible structural framework which can better accommodate the subsequent structural strains resulting from the successive lithium insertion / de-insertion processes, therefore likely contributing also to the more robust structure upon cycling. This is what suggests the in situ x-ray diffraction study during lithium insertion / de-insertion, showing the maintaining of crystallinity during lithium insertion until the fully discharged Li4-BPDC composition, and, the possible occurrence of an intermediate phase of composition Li3-BPDC. Supporting Information Experimental procedures, synthesis protocols, spectroscopic characterization, further electrochemical tests and complete in-situ XRD experiment are described in details in the electronic supporting information, available free of charge on the ACS publications website. Acknowledgements The French National Research network on electrochemical energy storage (RS2E) is gratefully acknowledged for supporting this project through the funding of the Ph-D grant of Lionel Fédèle. The ANR funding agency is also gratefully acknowledged for financial support through the grant accorded for the projects “VOLTA” and “Store-ex”. The authors wish to acknowledge Matthieu Courty and Dr. Dominique Cailleu for their helpful assistance for ThermoGravimetric Analysis and Solid-State NMR experiments, respectively. References (1)

Larcher, D.; Tarascon, J.-M. Towards greener and more sustainable batteries for electrical energy storage. Nat. Chem. 2015, 7, 19–29.

ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(2)

Tarascon, J.-M.; Recham, N.; Armand, M.; Barpanda, P.; Walker, W.; Dupont, L. Hunting for Better Li-Based Electrode Materials via Low Temperature Inorganic Synthesis †. Chem. Mater. 2010, 22, 724–739.

(3)

Li, N.; Martin, C. R. A High-Rate , High-Capacity , Nanostructured Sn-Based Anode Prepared Using Sol-Gel Template Synthesis. J. Electrochem. Soc. 2001, 148, 164–170.

(4)

Hu, Y.; Doeff, M. M.; Kostecki, R.; Fin, R. Electrochemical Performance of Sol-Gel Synthesized LiFePO 4 in Lithium Batteries. J. Electrochem. Soc. 2004, 151, A1279– A1285.

(5)

Patra, S.; Davoisne, C.; Bruyère, S.; Bouyanfif, H.; Cassaignon, S.; Taberna, P.-L.; Sauvage, F. Room-Temperature Synthesis of High Surface Area Anatase TiO 2 Exhibiting a Complete Lithium Insertion Solid Solution. Part. Part. Syst. Charact. 2013, 30, 1093–1104.

(6)

Sauvage, F.; Bodenez, V.; Vezin, H.; Albrecht, T. A.; Tarascon, J. M.; Poeppelmeier, K. R. Ag4V2O6F2 (SVOF): A high silver density phase and potential new cathode material for implantable cardioverter defibrillators. Inorg. Chem. 2008, 47, 8464–8472.

(7)

Andriamiadamanana, C.; Laberty-robert, C.; Sougrati, M. T.; Casale, S.; Davoisne, C.; Patra, S.; Sauvage, F. Room-Temperature Synthesis of Iron-Doped Anatase TiO2 for Lithium-Ion Batteries and Photocatalysis. Inorg. Chem. 2014, 53, 10128–10139.

(8)

Sauvage, F.; Bodenez, V.; Tarascon, J.-M.; Poeppelmeier, K. R. Room-Temperature Synthesis Leading to Nanocrystalline. J. Am. Chem. Soc. 2010, 21, 6778–6782.

(9)

Moradi, M.; Kim, J. C.; Qi, J.; Xu, K.; Li, X.; Ceder, G.; Belcher, A. M. A biofacilitated synthetic route for nano-structured complex electrode materials. Green Chem. 2016, 18, 2619–2624.

(10)

Rosant, C.; Avalle, B.; Larcher, D.; Dupont, L.; Friboulet, A.; Tarascon, J.-M. Biosynthesis of Co3O4 electrode materials by peptide and phage engineering:

ACS Paragon Plus Environment

Page 20 of 26

Page 21 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

comprehension and future. Energy Environ. Sci. 2012, 5, 9936–9943. (11)

He, W.; Zhang, X.; Du, X.; Zhang, Y.; Yue, Y. Bio-assisted synthesis of mesoporous Li 3 V 2 ( PO 4 ) 3 for high performance lithium-ion batteries. Electrochim. Acta 2013, 112, 295–303.

(12)

Nam, K. T.; Kim, D.-W.; Yoo, pil j.; Chiang, C.-Y.; Meethong, N.; Hammond, P. T.; Chiang, Y.-M.; Belcher, A. M. Virus-enabled synthesis and assembly of nanowires for lithium ion battery electrodes. Science, 2006, 312, 885–888.

(13)

Kim, J. C.; Li, X.; Kang, B.; Ceder, G. High-rate performance of a mixed olivine cathode with off-stoichiometric composition. Chem. Commun. 2015, 51, 13279–13282.

(14)

Kim, J. C.; Seo, D.; Ceder, G. Theoretical capacity achieved in a LiMn0.5Fe0.4Mg0.1BO3 cathode by using topological disorder. Energy Environ. Sci. 2015, 8, 1790–1798.

(15)

Gaberscek, M.; Küzma, M.; Janez, J. Electrochemical kinetics of porous, carbondecorated LiFePO4 cathodes: separation of wiring effects from solid state diffusio. Phys. Chem. Chem. Phys. 2007, 9, 1815–1820.

(16)

Ogihara, N. Nonaqueous secondary battery electrodes, nonaqueous secondary battery including the same, and assembled battery. PCT 2014, US 8,795,8.

(17)

Tarascon, J. M. Towards sustainable and renewable systems for electrochemical energy storage. ChemSusChem 2008, 1 , 777–779.

(18)

Poizot, P.; Dolhem, F. Clean energy new deal for a sustainable world: from non-CO2 generating energy sources to greener electrochemical storage devices. Energy Environ. Sci. 2011, 4, 2003.

(19)

Liang, Y.; Zhang, P.; Chen, J. Function-oriented design of conjugated carbonyl compound electrodes for high energy lithium batteries. Chem. Sci. 2013, 4, 1330.

(20)

Song, Z.; Zhou, H. Towards sustainable and versatile energy storage device - an

ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

overview of organic electrode materials. Energy Environ. Sci. 2013, 6 , 2280. (21)

Häupler, B.; Wild, A.; Schubert, U. S. Carbonyls : Powerful Organic Materials for Secondary Batteries. Adv. Energy Mater. 2015, 5, 1402034.

(22)

Dunningt, J. S.; Bennion, D. N. The Electrochemical Reduction of MetaDinitrobenzene in Dimethylsulfoxide. J. Electrochem. Soc. 1970, 117, 485–491.

(23)

Alt, H.; Binder, H.; Klempert, G.; Köhling, A.; Sandstede, G. Evaluation of organic battery electrodes: Voltammetric study of the redox behaviour of solid quinones. J. Appl. Electrochem. 1972, 2, 193–200.

(24)

Nishida, S.; Yamamoto, Y.; Takui, T.; Morita, Y. Organic rechargeable batteries with tailored voltage and cycle performance. ChemSusChem 2013, 6, 794–797.

(25)

Nishide, H.; Iwasa, S.; Pu, Y. J.; Suga, T.; Nakahara, K.; Satoh, M. Organic radical battery: Nitroxide polymers as a cathode-active material. Electrochim. Acta 2004, 50, 827–831.

(26)

Golriz, A. A.; Suga, T.; Nishide, H.; Rüdiger, B.; Gutmann, J. S. Phenothiazinefunctionalized redox polymers for a new cathode-active material. Rsc Adv. 2015, 5, 22947–22950.

(27)

Lécuyer, M.; Gaubicher, J.; Barrès, A.; Dolhem, F.; Deschamps, M.; Guyomard, D.; Poizot, P. Electrochemistry Communications Short communication A rechargeable lithium / quinone battery using a commercial polymer electrolyte. Electrochem. commun. 2015, 55, 22–25.

(28)

Zhu, Z.; Hong, M.; Guo, D.; Shi, J.; Tao, Z.; Chen, J. All-Solid-State Lithium Organic Battery with Composite Polymer Electrolyte and Pillar[5]quinone Cathode. J. Am. Chem. Soc. 2014, 136, 16461–16464.

(29)

Shimizu, A.; Kuramoto, H.; Tsujii, Y.; Nokami, T.; Inatomi, Y.; Hojo, N.; Suzuki, H.; Yoshida, J. I. Introduction of two lithiooxycarbonyl groups enhances cyclability of

ACS Paragon Plus Environment

Page 22 of 26

Page 23 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

lithium batteries with organic cathode materials. J. Power Sources 2014, 260, 211–217. (30)

Deunf, E.; Moreau, P.; Quarez, E.; Guyomard, D.; Dolhem, F.; Poizot, P. Reversible anion intercalation in a layered aromatic amine : a high-voltage host structure for organic. J. Mater. Chem. A 2016, 4, 6131–6139.

(31)

Wang, S.; Wang, L.; Zhang, K.; Zhu, Z.; Tao, Z.; Chen, J. Organic Li4C8H2O6 nanosheets for lithium-ion batteries. Nano Lett. 2013, 13, 4404–4409.

(32)

Renault, S.; Mihali, V. A.; Brandell, D. Optimizing the electrochemical performance of water-soluble organic Li-ion battery electrodes. Electrochem. commun. 2013, 34, 174– 176.

(33)

Oltean, V. A.; Renault, S.; Brandell, D. Electrochemistry Communications Enhanced performance of organic materials for lithium-ion batteries using facile electrode calendaring techniques. Electrochem. commun. 2016, 68, 45–48.

(34)

Armand, M.; Grugeon, S.; Vezin, H.; Laruelle, S.; Ribière, P.; Poizot, P.; Tarascon, J. Conjugated dicarboxylate anodes for Li-ion batteries. Nat. Mater. 2009, 8, 120–125.

(35)

Fédéle, L.; Sauvage, F.; Bois, J.; Tarascon, J.-M.; Bécuwe, M. Lithium Insertion / DeInsertion Properties of π-Extended Naphthyl-Based Dicarboxylate Electrode Synthesized by Freeze-Drying. J. Electrochem. Soc. 2014, 161, A46.

(36)

Fédèle, L.; Sauvage, F.; Bécuwe, M. Hyper-conjugated lithium carboxylate based on a perylene unit for high-rate organic lithium-ion batteries. J. Mater. Chem. A 2014, 2, 18225–18228.

(37)

Boultif, A.; Loue, D. Powder pattern indexing with the dichotomy method. J. Appl. Crystallogr. 2004, 37, 724–731.

(38)

Rodriguez-carvajal, J. Recent advances in magnetic structure determination by neutron powder diffraction. Phys. B 1993, 192, 55–69.

(39)

Banerjee, D.; Borkowski, L. A.; Kim, S. J.; Parise, J. B. Synthesis and Structural

ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Characterization of Lithium-Based Metal−Organic Frameworks. Cryst. Growth Des. 2009, 9, 4922–4926. (40)

Banerjee, D.; Parise, J. B. Recent advances in s-block metal carboxylate networks. Cryst. Growth Des. 2011, 11, 4704–4720.

(41)

Kaduk, J. a. Terephthalate salts: salts of monopositive cations. Acta Crystallogr. Sect. B Struct. Sci. 2000, 56, 474–485.

(42)

Burkhardt, S. E.; Bois, J.; Tarascon, J. M.; Hennig, R. G.; Abruna, H. D. Licarboxylate anode structure-property relationships from molecular modeling. Chem. Mater. 2013, 25, 132–141.

(43)

Bonhomme, C.; Gervais, C.; Babonneau, F.; Coelho, C.; Aza, T.; Ashbrook, S. E.; Gri, J. M.; Yates, J. R.; Mauri, F.; Pickard, C. J. First-Principles Calculation of NMR Parameters Using the Gauge Including Projector Augmented Wave Method : A Chemist ’ s Point of View ́ e r. Chem. Rev. 2012, 112, 5733–5779.

(44)

Harris, R. K.; Joyce, A.; Pickard, C. J.; Cadars, S.; Emsley, L. Assigning carbon-13 NMR spectra to crystal structures by the INADEQUATE pulse sequence and first principles computation : a case study of two forms of testosterone w. Phys. Chem. Chem. Phys. 2006, 8, 137–143.

(45)

Salager, E.; Stien, R. S.; Pickard, C. J.; Elena, B.; Emsley, L. Powder NMR crystallography of thymol. Phys. Chem. Chem. Phys. 2009, 11, 2610–2621.

(46)

Lesage, A.; Bardet, M.; Emsley, L. Through-Bond Carbon - Carbon Connectivities in Disordered Solids by NMR. J. Am. Chem. Soc. 1999, 121, 10987–10993.

(47)

Clark, S. J.; Segall, M. D.; Chris J Pickard; Hasnip, P. J.; Prober, M. I. J.; Refson, K. First principles methods using CASTEP. Zeitschrift für Krist. - Cryst. Mater. 2005, 220, 567–570.

(48)

Pickard, C. J.; Mauri, F. All-electron magnetic response with pseudopotentials: NMR

ACS Paragon Plus Environment

Page 24 of 26

Page 25 of 26

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Chemistry of Materials

chemical shifts. Phys. Rev. B 2001, 63, 1–13. (49)

Yates, J. R.; Pickard, C. J.; Mauri, F. Calculation of NMR chemical shifts for extended systems using ultrasoft pseudopotentials. Phys. Rev. B 2007, 76, 24401–24412.

(50)

Walker, W.; Grugeon, S.; Armand, M.; Tarascon, J. Electrochemical characterization of lithium 4 , 4 0 -tolane-dicarboxylate for use as a negative electrode in Li-ion batteries †‡. J. Mater. Chem. 2011, 21, 1615–1620.

(51)

Walker, W.; Grugeon, S.; Vezin, H.; Laruelle, S.; Armand, M.; Tarascon, J. M.; Wudl, F. Electrochemistry Communications The effect of length and cis / trans relationship of conjugated pathway on secondary battery performance in organolithium electrodes. Electrochem. commun. 2010, 12, 1348–1351.

(52)

Mihali, V. A.; Renault, S.; Nyholm, L.; Brandell, D. RSC Advances Benzenediacrylates as organic battery electrode materials : Na versus Li †. RSC Adv. 2014, 4, 38004–38011.

(53)

Ogihara, N.; Yasuda, T.; Kishida, Y.; Ohsuna, T.; Miyamoto, K.; Ohba, N. Organic Dicarboxylate Negative Electrode Materials with Remarkably Small Strain for HighVoltage Bipolar Batteries. Angew. Chemie Int. Ed. 2014, 53, 11467–11472.

(54)

Choi, A.; Kim, Y. K.; Kim, T. K.; Kwon, M.-S.; Lee, K. T.; Moon, H. R. 4,4′Biphenyldicarboxylate sodium coordination compounds as anodes for Na-ion batteries. J. Mater. Chem. A 2014, 2, 14986.

(55)

Santhanam, R.; Noel, M. Influence of polymeric binder on the stability and intercalation / de-intercalation behaviour of graphite electrodes in non-aqueous solvents. J. Power Sources 1996, 63, 1–6.

(56)

Yamada, Y.; Iriyama, Y.; Abe, T.; Ogumi, Z. Kinetics of Lithium Ion Transfer at the Interface between Graphite and Liquid Electrolytes : Effects of Solvent and Surface Film. Langmuir 2009, 25, 12766–12770.

ACS Paragon Plus Environment

Chemistry of Materials

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

(57)

Abe, T.; Fukuda, H.; Iriyama, Y.; Ogumi, Z. Solvated Li-Ion Transfer at Interface Between Graphite and Electrolyte. J. Electrochem. Soc. 2004, 151, A1120–A1123.

ACS Paragon Plus Environment

Page 26 of 26