Fan Zhanga,
Károly Németh*b,
Javier Bareñoc,
Fulya Doganc,
Ira D. Bloomc and
Leon L. Shaw*a
aMechanical, Materials and Aerospace Engineering Department, Illinois Institute of Technology, Chicago, IL 60616-3793, USA. E-mail: lshaw2@iit.edu; Tel: +1-312-567-3844
bDepartment of Physics, Illinois Institute of Technology, Chicago, IL 60616-3793, USA. E-mail: nemeth@agni.phys.iit.edu; Tel: +1-630-632-2382
cChemical Sciences and Engineering Division, Argonne National Laboratory, Argonne, IL 60439, USA
First published on 3rd March 2016
The feasibility of synthesizing functionalized h-BN (FBN) via the reaction between molten LiOH and solid h-BN is studied for the first time and its first ever application as an electrode material in Li-ion batteries is evaluated. Density functional theory (DFT) calculations are performed to provide mechanistic understanding of the possible electrochemical reactions derived from the FBN. Various materials characterizations reveal that the melt-solid reaction can lead to exfoliation and functionalization of h-BN simultaneously, while electrochemical analysis proves that the FBN can reversibly store charges through surface redox reactions with good cycle stability and coulombic efficiency. DFT calculations have provided physical insights into the observed electrochemical properties derived from the FBN.
Although functionalized BN (FBN) materials have never been studied for electrode applications, oxygen-functionalized graphite nanoplatelets (OfGNP-s) have been investigated previously by Liu, et al.2 as cathode materials for Li-ion batteries, and have demonstrated extremely high specific power in the order of 10–100 kW kg−1 combined with specific energy of 100–1000 W h kg−1. Stable operation with little capacity loss has been demonstrated for 1000 cycles. A similar approach using a sodium anode and OfGNP cathode has been studied by Kim, et al.3 demonstrating specific energies of 100–500 W h kg−1 and specific powers of 50–100 kW kg−1 with 300 cycles at a current rate of 1 A g−1. In the OfGNP cathodes the reactive surface species are either oxo (CO) or epoxide (C–O–C) functional groups that become reduced to C–O–(Li/Na) during discharge.3 The capacity and energy density of an OfGNP cathode depend on the concentration of oxygen on the surface of the nanoplatelets. The greater the oxygen concentration, the higher the capacity and energy density.2,3 Unfortunately, the high concentration of oxygen increases the risk of thermal runaway processes in which the OfGNP may explosively decompose to carbon monoxide,6 questioning whether OfGNP will ever be safe enough for mass-applications. Liu, et al.2 also discuss other than oxo or epoxide functionalizations of GNP-s, such as –COOH, –OH, –NH2, –OR, and –COOR functionalization. However, these functionalizations have not been tested and thus the reactivity and cycle-stability of these species are uncertain.
A potentially safer, thermally more resistant alternative to OfGNP may be FBNs, as proposed by Nemeth.4,5 Monolayer h-BN is resistant to oxidation up to 850 °C (ref. 6) at which oxidation and cleavage of h-BN monolayers happen. In contrast, monolayer graphene can be oxidized at as low as about 200 °C in a stream of oxygen containing about 1% water2 and graphene oxide explosively decomposes at about 300 °C.7 No explosive decomposition of oxidized h-BN has been observed.6 Therefore, it is worthwhile to explore the possibility of utilizing FBNs as electrodes for Li-ion battery applications. No such studies have ever been conducted before, although layered BN has been used previously as an additive to graphene oxide electrodes to prevent re-stacking of graphene oxide.8
The covalent functionalization of h-BN with –OH,9–12 –NH2,13 CBr2,14 –F,15 and substituted phenyl16 functional groups has been realized through radicals9–11,13–16 or nucleophilic/electrophilic attack2,17,18 of anions/cations on the boron/nitrogen atoms. Both boron and nitrogen atoms can be covalently functionalized. Radicals and anions based nucleophilic attacks create bonds to the electron-deficient boron atoms (Lewis acid), while molecular cations may functionalize the electron-rich nitrogen atoms (Lewis base). Molten alkaline salts have been used for decades to etch h-BN whereby they covalently functionalize and dissolve h-BN.17,18 Non-covalent functionalization of h-BN is also possible, for example, by molten amines that utilize a boron-amine Lewis acid–base interaction.19,20 To the best of our knowledge, although many functionalized h-BN species exist, so far only a single one is known in both alkalinated (reduced) and non-alkalinated (oxidized) forms, namely hydroxyl functionalized boron nitride, for which both the BN(OH)x9–11 and BN(OH)xNax (where 0 < x ≤ 1)12 forms have been synthesized.
In this study, we investigate, for the first time, the synthesis of FBN powder via the reaction between molten LiOH and solid h-BN and its first ever application as an electrode material in Li-ion batteries. Various materials characterizations reveal that the melt-solid reaction can lead to exfoliation and functionalization of h-BN simultaneously, while electrochemical analysis and density functional theory (DFT) calculations prove that FBN can reversibly store charges through surface redox reactions.
X-ray photoelectron spectroscopy (XPS) was carried out with a Physical Electronics 5000 VersaProbe II system equipped with a monochromatic aluminum Kα (15 kV) X-ray source. The excitation beam size employed was 100 μm and the power was 25 W. The background pressure in the analysis chamber was around 1.0 × 10−7 Pa. XPS spectra were recorded in fixed analyzer transmission mode. A survey spectrum was acquired using a pass energy of 117.4 eV, a step size of 1 eV, and an acquisition time of 150 ms per step. Individual element regions were acquired using a pass energy of 23.5 eV, a step size of 0.2 eV, and an acquisition time of 3.0 s per step. In all cases Ar+ and e− neutralization were employed. Adventitious carbon on the sample was used to calibrate the binding energy (B.E.) scale with respect to C–H environments at 284.8 eV. Individual element regions were further deconvoluted into components by least square fitting to a sum of Gaussian or Gaussian–Lorentzian components and a Shirley background using the Multipack package by Physical Electronics, provided with the system.
1H, 7Li and 11B solid state NMR experiments were performed on a 500 MHz Bruker Advance III spectrometer (11.7 Tesla superconducting magnet) with a 2.5 mm MAS probe operating at 30 kHz spinning speed. A rotor synchronized echo pulse sequence (90°τ–180°τacq), where τ = 1/νr (spinning frequency) was used for all the acquisitions. A π/2 pulse width of 2.0 μs and pulse delay of 2 s was used for 1H experiments and the chemical shift was referenced to TMS at 0 ppm. 7Li experiment was performed with a π/2 pulse width of 2.5 μs and pulse delay of 5 s and was referenced with 1 M LiCl at 0 ppm. For 11B MAS NMR, a π/2 pulse width of 2.6 μs and a pulse delay of 2 s were used. 0.1 M boric acid was used as a secondary reference at 18.8 ppm. The simulation and the fit of spectrum was performed using a shape analysis package included in Topspin software. All the experiments were performed at room temperature and the sample was packed under an inert atmosphere in a glovebox.
Since the amount of CB in the electrode was high (50 wt%), pure CB half cells (i.e., 90 wt% CB + 10 wt% PVDF versus a lithium metal anode) were also fabricated in order to take the charge storage effect of CB in the calculation of the specific capacity of FBN. In making these half cells, the procedure of making FBN half cells was followed precisely. Furthermore, the mass loading of CB in each electrode was set to be ∼2 mg cm−2, similar to that used in the FBN electrodes. The CB half cells were charged and discharged between 1.0 and 3.3 V with a current density of 28 mA g−1 CB which again was similar to that used for charging/discharging FBN half cells.
The peak at ∼1155 cm−1 is associated with the B–O deformation vibrations, similarly to an analogous band observed in (HO)xBN.11,12 While ref. 11 reports a very broad and shallow peak between 3000 and 3600 cm−1 that is assigned to hydrogen bonds between and within (HO)xBN layers, we could not identify its analogue in our product. This suggests that the FBN formed through the melt-solid reaction may contain little Li(HO)BN or (HO)BN because some inter- and intra-molecular hydrogen bonds are expected from Li(HO)BN and (HO)BN. However, as we will see later, XPS and NMR data indicate the presence of OH-functional group in the as-synthesized FBN. Thus, we hypothesize that one of the possible reasons for the lack of intra-FBN hydrogen bonding in the FTIR data might be due to the fact that only about 1/3 of B atoms are functionalized (as indicated by XPS and NMR analyses to be discussed later) and thus most of the surface –OH groups are too far from each other to form hydrogen bonds. In addition, the presence of the –OLi group on the surface (identified using XPS and to be discussed later) can also prevent the formation of intra-FBN hydrogen bonding. The lack of inter-FBN hydrogen bonding is likely due to the relatively random orientations of FBN monolayers which does not allow for the formation of a sufficient number of inter-FBN hydrogen bonds to be detected by FTIR.
Fig. 1(b) compares the XRD patterns of h-BN, LiOH, and the as-synthesized FBN. Note that all the peaks of h-BN and LiOH are absent in the melt-solid reaction product, indicating that exfoliation of h-BN has taken place during the melt-solid reaction. This is consistent with the expectation because the completely exfoliated h-BN has no XRD peaks,26 similarly to graphene and graphene oxide.3 The complete disappearance of XRD peaks can be expected for any monolayer FBNs as well, such as for Li(HO)BN, Li2(O)BN, Li(O)BN, and (O)BN. Another possibility for disappearance of h-BN peaks is amorphization. However, the FTIR analysis above reveals that the h-BN bonding is still present after the melt-solid reaction. Therefore, disappearance of h-BN peaks cannot be attributed to amorphization, but is due to exfoliation of h-BN into monolayers with functionalization on the surface of monolayers. In spite of functionalization and exfoliation, it is noted that accompanied with disappearance of h-BN peaks and LiOH peaks, several new yet small intensity peaks appear. The intensities of these peaks are very small since the XRD patterns in Fig. 1(b) are collected under the same conditions and presented in the same scale. Thus, these small peaks suggest the formation of a small quantity of crystalline phase(s) along with FBNs. The nature of these trace amounts of crystalline phases is unknown at this stage, but they are not the possible compounds of crystalline Li3BO3, LiBO2, Li2BNO and BNO since these compounds do not match the crystalline peaks of these compounds as shown in Fig. 1(c) where the peak intensities of these trace amounts of crystalline phases have been magnified for easy observation and their peak positions are marked as well.
Fig. 2 shows XPS spectra of the as-synthesized sample. Fig. 2(a) contains a survey spectrum showing that, other than small amounts of F and C (∼1 at% and ∼5 at%, respectively), the sample does not contain detectable amounts of any atomic species not expected in the as-synthesized FBN. Fig. 2(b) through Fig. 2(f) contain detailed XPS scans of the B1s, N1s, O1s, Li1s, and C1s regions of the same sample. Each of these figures contains the experimental data (black dots) after subtraction of a Shirley background, as well as fit and individual components (continuous lines) resulting from a least squares deconvolution of the data. Reported binding energies (B.E.), taken from references listed in the NIST XPS Database,27 are also indicated by discontinuous vertical lines in Fig. 2(b) through Fig. 2(f), for reference. Table 1 summarizes the sample composition estimated from the XPS data presented in Fig. 2(b) through Fig. 2(f), while Table 2 lists the fit parameters resulting from the least squares deconvolution. Note that both Tables 1 and 2 do not contain H element because XPS cannot detect H element directly.
Element | B | N | O | Li | C | F |
---|---|---|---|---|---|---|
Concentration (at%) | 33 | 29 | 21 | 12 | 5 | 1 |
Region | Centroid (eV) | FWHM (eV) | Area (% of region) | Gaussian ratio (%) | χ2 |
---|---|---|---|---|---|
B1s | 187.22 | 2.63 | 4.64 | 100 | 1.02 |
188.71 | 1.63 | 9.08 | 100 | ||
190.60 | 1.96 | 71.75 | 100 | ||
192.66 | 1.58 | 14.53 | 100 | ||
N1s | 395.89 | 2.43 | 13.9 | 77 | 2.50 |
397.86 | 2.00 | 65.95 | 87 | ||
398.35 | 1.43 | 20.15 | 77 | ||
O1s | 530.42 | 4.23 | 28.72 | 92 | 0.73 |
531.44 | 2.43 | 49.11 | 100 | ||
532.99 | 2.01 | 22.17 | 100 | ||
Li1s | 55.70 | 2.89 | 100 | 60 | 1.25 |
C1s | 284.90 | 2.71 | 77.10 | 78 | 1.25 |
288.70 | 3.25 | 22.90 | 68 | ||
F1s | 685.38 | 2.88 | 100 | 57 | 1.13 |
The B1s spectral region, as shown in Fig. 2(b), is dominated by a Gaussian contribution at B.E. ∼190.6 eV, close to the reported 190.5 eV value of B in BN. Two additional components centered at 192.66 eV and 188.71 eV can be respectively attributed to electron-deprived and electron-enriched B environments. The component centered at 192.66 eV is shifted from BN toward B(OH)3 and B2O3 environments, reported at 193.6 eV and 193.7 eV, respectively, consistent with electron withdrawing OH-functionalization or O-functionalization of B atoms resulting in HO–BN and O–BN environments, respectively. Note that LiO–BN will also result in a similar shift as HO–BN, and thus the peak at 192.66 eV is assigned to HO–BN, LiO–BN and O–BN (Table 3). In contrast, the component centered at 188.71 eV is shifted toward B10H14 at 187.8 eV, consistent with H functionalized B in H–BN environments. Finally, a minor, broad component centered at 187.22 eV can be mainly attributed to metallic B, perhaps with some contributions of boron hydrides and carbides.
Region | Centroid (eV) | Concentration (at% of sample) | Assignment |
---|---|---|---|
B1s | 187.22 | 1.5 | B |
188.71 | 3.0 | H![]() ![]() |
|
190.60 | 23.7 | BN | |
192.66 | 4.8 | HO![]() ![]() ![]() ![]() ![]() ![]() |
|
N1s | 395.89 | 4.0 | BN![]() ![]() |
397.86 | 19.1 | BN | |
398.35 | 5.8 | ||
O1s | 530.42 | 6.0 | Organic |
531.44 | 10.2 | Li2CO3/Li2O/LiOH | |
532.99 | 4.6 | HO![]() ![]() ![]() ![]() ![]() ![]() |
|
Li1s | 55.70 | 12.0 | Li2CO3/Li2O/LiOH/LiF/LiO![]() ![]() |
C1s | 284.90 | 3.4 | C–H/Organic |
288.70 | 1.0 | Li2CO3 | |
F1s | 685.38 | 1.0 | LiF |
The N1s spectral region, Fig. 2(c), is well described by three Gaussian–Lorentzian peaks centered at 395.89 eV, 397.86 eV, and 398.35 eV. The main peak (at 397.86 eV), which accounts for ∼65% of the signal in the N1s region, is very close to the reported N1s B.E. of BN (398.1 eV) and is broad enough (FWHM – 2 eV) to comfortably cover the BN range and to encompass the narrower peak (FWHM – 1.4 eV) centered at 398.35 eV, which itself accounts for ∼20% of the signal in the N1s region. This secondary peak is shifted towards N environments with lower electron density with respect to BN, such as lithium azide and ammonia, reported at 398.7 eV. Conversely, the peak centered at 395.89 eV is shifted towards higher electron density environments, consistent with H functionalization of the N atom in BN; i.e., BN–H environments. The sum of the atomic concentrations associated with the two higher B.E. components of the N1s region (Table 3) accounts for ∼25% of the observed atoms in the sample. This is in good agreement with the amount of atoms accounted for by the B1s contribution at 190.6 eV (∼24%) and BN stoichiometry. However, the breadth of the N1s contributions and their lack of clear resolution suggest that, rather than representing three clearly distinguished N environments in BN (i.e., BN and two different N functionalizations), they constitute a mathematical approximation to a more complicated N1s peak shape stemming from a distribution of N environments in the sample. For example, it is not difficult to imagine that an HO–BN functionalized B atom will pull some additional charge from the neighboring N atoms to partially compensate for the charge donated to the OH group. In this case, the N atoms neighboring this B atom will show up at higher B.E. than in pure BN (398.1 eV). This interpretation is somewhat supported by the fact that the total amount of N in the sample detected by XPS (∼29 at%) is close to the total amount of B attributed to BN, H–BN, HO–BN, LiO–BN and O–BN-like environments (∼31 at%). The small discrepancy from overall BN stoichiometry is within standard estimates of XPS accuracy of a few at%, but could be indicative of a slight preference for B termination at the edges of the BN sheets, or stabilization of N vacancies due to the functionalization.
The O1s spectral region, Fig. 2(d), is described by three Gaussian–Lorentzian components centered at 530.42 eV, 531.44 eV, and 532.99 eV. The main component, which is centered at 531.44 eV and accounts for half of the area of the region, is close to the reported B.E. of lithium carbonate (531.5 eV), lithium oxide (531.3 eV) and lithium hydroxide (531.2 eV). This peak probably represents remaining unreacted LiOH and/or contamination from its reaction with adventitious atmospheric hydrocarbons. The highest B.E. contribution in the O1s region, centered at 532.99 eV, is shifted towards H2O (533.2 eV) and B(OH)3 (533.4 eV) positions, consistent with O in HO–BN, O–BN and LiO–BN environments. This assignment is supported by the close match between the number of O atoms (4.6 at%) and B atoms (4.8 at%) accounted for this and the corresponding B1s contribution (at 192.66 eV) assigned to B atoms in HO–BN, O–BN and LiO–BN moieties (see Table 3). A final O1s contribution, centered at 530.42 eV and accounting for ≲30% of the O1s signal, is close to the reported B.E. of aminobutyric acid (530.7 eV) and several metal oxides. Lacking evidence of the presence of metals other than boron and lithium, we attribute this peak to organic oxygen.
The Li1s spectral region, Fig. 2(e), can be described by a single, broad (FWHM – 2.89 eV) Gaussian–Lorentzian contribution centered at 55.7 eV. The Li1s signal spans reported B.E. values for several lithium compounds, indicated in the figure, including lithium carbonate (55.2 eV), oxide (55.6 eV), hydroxide (54.9 eV) and fluoride (55.7 eV). Similarly, the F1s region (not shown here because of its low concentration, 1 at% only) can be described by a single broad (FWHM = 2.88 eV) Gaussian–Lorentzian component centered at 685.38 eV, close to the reported 685.1 eV value for LiF. Note that there are only 1 at% LiF and 10 at% Li2CO3/Li2O/LiOH, while the total amount of Li is 12 at%. This suggests that LiO–BN is about 1 at%. Here, the LiOH could be a remnant of the precursor used to functionalize the BN, and the Li2O a product of its reaction to donate H to H–BN and BN–H; however, the origin of the fluorine is not clear at this point.
The C1s region, Fig. 2(f), contains two Gaussian–Lorentzian components, centered at 284.9 eV and 288.7 eV. The smaller of these two components (at 288.7 eV) accounts for only ∼1% of the atoms in the sample and spans the B.E. reported for lithium carbonate (289.8 eV). Although it is possible that a fraction of its intensity corresponds to Li2CO3, it is more likely that most of the C in the sample corresponds to adventitious atmospheric hydrocarbons and, potentially, N-containing organic products of its reaction with BN.
Fig. 3 shows 1H, 7Li and 11B solid state NMR data for the as-synthesized FBN sample. 1H MAS NMR reveals the presence of at least four different proton peaks at around −2, 0.6, 2.8 and 6.2 ppm. The lower frequency peak at −2.0 ppm is due to the presence of LiOH, whereas 0.6 and 2.8 ppm peaks might be due to the presence of H–BN or HO–BN environments and NH–groups, respectively. Higher frequency proton peaks (6 ppm and above) are generally due to either hydrogen bonded OH-groups or the presence of protonated amines. 7Li MAS NMR shows one relatively broad peak at ∼0 ppm. Due to low and close chemical shift range of diamagnetic 7Li NMR peaks, different lithium bearing species (such as Li2O, LiF, LiOH and LiO–BN) are expected to overlap and give one broad resonance.
11B MAS NMR data of FBN sample shows at least 4 different boron environments (Fig. 3(c)). 11B is a quadrupolar nuclei and it can give rise to significant broadening and distortions of the NMR resonances of compounds with asymmetrical boron environments. Boron containing oxo-compounds typically contain tetrahedral BO4 or trigonal BO3 configurations. The former has a relatively narrow single resonance with δiso of 2 to −4 ppm and a CQ of 0 to 0.5 MHz whereas the latter has larger CQ of around 2.5 MHz and δiso values ranging from 12 to 19 ppm. For boron nitrides, the cubic phase (c-BN) generally shows a symmetric 11B NMR resonance at around 1.6 ppm and h-BN gives a single quadrupolar lineshape peak with δiso of ∼30 ppm and CQ of 2.8 MHz.28,29
As seen in Fig. 3(c), the FBN sample shows at least two distorted boron environments with isotropic chemical shift of 18.3 and 29.7 ppm mostly due to the presence of 3-coordinated B–O and/or B–N environments. The presence of the two peaks at 1.4 and 6.8 ppm suggests the formation of 4-coordinate B–O or B–O–N environments as these peaks are not expected to be present in pristine h-BN sample and a cubic BN phase is unlikely. The relative intensities of these peaks are 100, 69.9, 40.6 and 57.6 at 29.7, 18.3, 6.8 and 1.4 ppm, respectively. This means that the 4-coordinated B has a relative amount of 36% of all boron. A trivial explanation of the 4-coordinated B would be the presence of glassy B2O3 which could be in an amorphous state and will not show up in the XRD pattern. However, if this 4-coordinated B was due to B2O3, it would have produced significant summation and overtone vibrations in the FTIR spectrum, between 2000 and 3000 cm−1.30 Since the FTIR spectrum is completely flat at this region, we have assigned this 4-coordinated B to O atoms attached to the B-s in the basal plane of h-BN, i.e., O–BN as revealed in the XPS analysis.
In summary, FTIR, XRD, XPS and NMR analyses have indicated unambiguously the functionalization of h-BN and the products of the functionalization are HO–BN, O–BN, LiO–BN and H–BN. However, about two-third h-BN do not appear to be functionalized even though its long-range order has disappeared (as evidenced by the disappearance of its crystalline peaks in the XRD pattern). A small amount of LiOH may remain unreacted in the sample. Furthermore, the XPS analysis reveals that the melt-solid reaction condition (500 °C) has resulted in some loss of LiOH, most likely due to the vaporization of LiOH during synthesis since the melting point of LiOH is only 462 °C.
SEM images of the as-synthesized FBN particles and the starting h-BN powder are shown in Fig. 4. Many particles of the as-synthesized FBN are agglomerated and the particle sizes range from 200 nm to 3 μm. This morphology is quite different from that of the starting h-BN powder which contains a large number of h-BN sheets with some particles. Further, the starting h-BN particles have larger sizes (∼2 to 20 μm) than that of the final product, suggesting that molten LiOH not only exfoliates h-BN, but also reacts with h-BN in the direction perpendicular to the basal plane, thereby breaking large h-BN sheets into multiple pieces with smaller basal planes. This is quite different from the work by Li, et al.18 who have formed transparent h-BN nanosheets (∼4 nm thick) through exfoliation of h-BN using molten NaOH/KOH mixture at 180 °C for 2 h. The exact reason for the different morphologies is not clear at this stage, but it is likely due to the higher temperature used in this study (500 °C in order to create molten LiOH). One of the possible mechanisms is that high temperature could result in several reactions more than just exfoliation and substantial agglomeration of FBN sheets, leading to the formation of small particles rather than nanosheets.
The representative charge/discharge behavior of CB half cells is shown in Fig. 5(a). The open circuit voltage (OCV) is 2.55 V, and the cell is charged first to 3.3 V and then discharged to 1 V. The first discharge gives rise to two plateaus with one at ∼2.2 V and the other below 1.3 V. The specific capacity for the first discharge is large (∼123 mA h g−1), but it decreases drastically to 38 mA h g−1 in the second discharge and continues to decrease gradually beyond the second cycle. However, the capacity decrease becomes extremely small after the tenth cycle and the specific capacity stays at ∼28 mA h g−1. The two plateaus in the first discharge are attributed to oxide impurities in CB which react with Li+ ions irreversibly. In a recent study on activated carbon31 we have found that activated carbon half cells with a lithium metal anode also exhibit such plateaus in the first discharge. The EDS analysis31 reveals that the main impurities in activated carbon are SiO2, CaO, Al2O3 and FeO (see Table S1 in ESI†). All of these oxide impurities can be reduced by Li+ ions. For example, SiO2 reacts with Li+ ion irreversibly to form Li4SiO4 alloy according to the reaction of 2SiO2 + 4Li+ + 4e− = Li4SiO4 + Si.32 After the first cycle the plateau near 2.2 V disappears completely (Fig. 5(a)), clearly indicating that this plateau is caused by irreversible reactions. In contrast, the plateau below 1.3 V is still present but with much smaller capacities in the subsequent cycles.
![]() | ||
Fig. 5 Voltage profiles of (a) a CB half cell and (b) a FBN half cell, charged and discharged from 1.0 to 3.3 V at a current density of 28 mA g−1 CB and FBN, respectively. |
Fig. 5(b) shows the representative charge/discharge behavior of FBN half cells. The specific capacity in this figure is calculated through subtracting the total capacity of the entire electrode by the capacity derived from the given amount of CB in the electrode (e.g., 28 mA h g−1 of 50 wt% CB in the electrode when the cell voltage is at 1.0 V). The resulting capacity after the subtraction is then divided by the weight of FBN in the electrode to obtain the specific capacity of FBN. Note that the first discharge profile has a plateau at ∼2.2 V which disappears in the following cycles. This plateau is ascribed to the effect of 50 wt% CB in the electrode because it has the same behavior as the one displayed by the CB half cell (Fig. 5(a)). Similarly, the second plateau below 1.6 V in the first cycle is partially due to CB, but it is also partially from some irreversible reactions of FBN because CB cannot account for all of the irreversible capacity in the first discharge. However, the most interesting phenomenon in Fig. 5(b) is that FBN can offer a specific capacity at ∼78 mA h g−1 with a good cycle stability and coulombic efficiency (Fig. 6(a)).
![]() | ||
Fig. 6 (a) The specific capacity and coulombic efficiency of the FBN half cell in Fig. 5(b) as a function of charge/discharge cycles, and (b) the differential capacity vs. voltage of the charge/discharge profiles of CB and FBN half cells shown in Fig. 5(a) and (b), respectively. The vertical dashed lines indicate the location where dQ/dV starts to deviate from constants. |
To better understand the charge/discharge behavior of FBN, Fig. 5 has been re-plotted as the dQ/dV vs. V curves (where Q is the quantity of charge and V is the voltage), as shown in Fig. 6(b). It can be seen that dQ/dV is a constant between 1.85 and 2.80 V for FBN and between 1.60 and 2.80 V for CB, indicating that both half cells store charges through electrical double layer (EDL) mechanism in these voltage ranges.33,34 However, redox reactions do take place below 1.85 V and 1.60 V for FBN and CB, respectively, because dQ/dV is not a constant below these respective voltages.
To further understand the nature of redox reactions, cyclic voltammetry (CV) tests were conducted. As shown in Fig. 7, there are two reduction reactions in the first scan, but the reduction reaction R1 at ∼2.0 V potential disappears in the subsequent scans, clearly indicating that R1 is not a reversible reaction. This behavior is in excellent agreement with charge/discharge behavior of CB half cells and FBN half cells shown in Fig. 5, and this reduction has been attributed to oxide impurities in CB. It is interesting to note that R2 reduction at ∼0.6 V potential is reversible since a corresponding oxidation peak O2 is present. Furthermore, in the first scan R2 reduction contains some irreversible components because the current of R2 reduction decreases significantly in the subsequent scans. Again, these phenomena match the charge/discharge behavior of FBN half cells very well (Fig. 5(b)), unambiguously proving that redox reactions are present at the cell voltage below ∼1.8 V.
One of the possible mechanisms for redox reactions of FBN is Li reacting with surface-bound oxygen radicals of Ox–BN, as shown in Fig. 8, with the electrochemical reaction shown below.
Ox–BN + 2xLi+ + 2xe− → Li2xOx–BN (x ≤ 1) | (1) |
![]() | ||
Fig. 8 A fragment of oxidized h-BN (O–BN) 3 × 3 supercell, as cut out from an infinite 2D monolayer. Color code: B – pink, N – blue, O – red, Li – magenta. |
Our DFT calculations predict that this redox reaction gives an electrochemical potential of 2.08 V (vs. Li/Li+) for x = 1. In this reaction, the h-BN sheet may be viewed as a 2D monolayer that captures ·O· radicals and binds them to one B and one N atom through an epoxide bond on the surface. Therefore, the cell reaction may be simplified to
·O· + 2Li+ + 2e− → 2Li+ + O2− | (2) |
The proposed surface intercalation/conversion reactions of FBN-s are only distantly related to Li intercalation in bulk h-BN, even though electron absorption by electron deficient valence shells of B atoms happen in both processes. The intercalation of pure Li (without –O– and –OH groups) into bulk h-BN has been experimentally studied in the literature36 and has been theoretically modeled as well.37 However, the intercalation of pure Li into bulk h-BN is a very different process, as the presence of –O– and –OH groups in our case accelerates the exfoliation of h-BN, resulting in functionalized exfoliated monolayers. In contrast, pure Li intercalation between h-BN layers results in 3D crystals instead of exfoliated and functionalized monolayers. In spite of their difference, the absorption of electrons into the electron-deficient valence shell of B atoms of h-BN monolayers also happens even by electron donating pure Li and other alkali metals, e.g. when they are dissolved in 1,2-dimethoxyethane (DME) at room temperature.38 In this case the alkali metal generates solvated electrons in the DME solvent (similarly to liquid ammonia), the solvated electrons are absorbed by the layers of bulk h-BN, and the negatively charged h-BN layers exfoliate due to electrostatic repulsion and will be surrounded by solvated alkali cations in the solution/suspension, as described recently in ref. 38. This latter experiment is another demonstration that h-BN monolayers are capable of absorbing electrons without decomposing, even though when functionalized, the electron absorption happens on sp3 hybridized B atoms, instead of sp2 ones (the latter is the case for pure h-BN).
This charge storage mechanism is unique for h-BN and cannot be found in graphene, as the absorption of radicals on graphene leads to the rearrangement of double and single bonds and in general does not preserve the radical character of the absorbed species. Strong magnetism is expected to be characteristic for most radical functionalized h-BNs, and it has been detected in fluorinated h-BN.15
Since XPS and NMR analyses have indicated the presence of HO–BN in the as-synthesized FBN, we have also performed DFT calculations for hydroxyl (–OH) functionalization and found that this functionalization offers a greater electrochemical potential (∼2.97 V vs. Li/Li+) with the following redox reaction.
HO–BN + Li+ + e−→ Li(HO)–BN | (3) |
This reaction will provide a theoretical specific capacity of 550 mA h g−1 and a specific energy of 1632 W h kg−1 when coupled with a Li anode. However, experimentally no redox reactions at 2.97 V are observed. The discrepancy may be due to the significant agglomeration of the as-synthesized particles while the DFT calculations assure isolated monolayers.
Experimentally, we have only observed sloping charge/discharge curves from 1.85 to 1.0 V (Fig. 5), which is closer to the predicted voltage of 2.08 V for the redox reaction of OxBN indicated in eqn (1). The lacking of a voltage plateau at 2.08 V may be attributed to several possible mechanisms. These are (i) multiple surface functional groups (e.g., –O and =O), (ii) different patterns of –O and O functional groups on the surface of h-BN, (iii) partial surface functionalization, (iv) agglomeration of FBN particles, and (v) impurities in the original h-BN. All of these can result in redox reactions at the surface of FBN particles with different electrochemical potentials and thus sloping voltage profiles in the charge/discharge cycles. This is consistent with many recent studies on carbon and graphene with O- and N-containing functional groups2,3,39–41 which have shown sloping voltage profiles with a wide voltage range, e.g., from 0.01 to 2.0 V39–41 or from 1.0 to 4.0 V,2,3 for Li-ion and Na-ion reactions at the surface of the functionalized carbon and graphene.
It should be pointed out that when the charge/discharge voltage range is expanded from 1.0 to 3.3 V to from 0.4 to 3.3 V, the reversible discharge specific capacity of the FBN half cell is increased from ∼78 mA h g−1 to ∼145 mA h g−1 with good cycle stability and coulombic efficiency (Fig. 9). If the discharge voltage is further reduced to end at 0.01 V, the reversible discharge specific capacity is increased further to ∼400 mA h g−1. This value is compatible with the specific capacity of the state-of-the-art graphite anode (∼350 mA h g−1). Fig. 10(a) shows how the voltage profile varies with the charge/discharge voltage window for the 4th charge/discharge cycle, whereas Fig. 10(b) compares the discharge specific capacities of the 3rd and 4th cycles for three charge/discharge voltage windows. It is very clear from Fig. 9 that the increased specific capacity for the charge/discharge voltage window from 0.4 to 3.3 V is due to redox reactions. This is also the case for the charge/discharge voltage window from 0.01 to 3.0 V (Fig. 10). These phenomena indicate that the FBN synthesized from the melt-solid reaction has the potential for anode applications if nanosheets with little agglomeration can be produced.
It is worthy of mentioning that the valences of atoms in eqn (3) are as follows: B + 3, N − 3, C + 4, H + 1, and Li + 1 while the O valence changes from −1 to −2 during lithiation in both equations. These are idealized valences that do not take into account that the change of the O valence is partially distributed over the N atoms, as discussed above. The simplified valences are based on a model in which the reversible reduction of hydroxyl radicals happens without the involvement of the atoms of the h-BN substrate to which these radicals are covalently bound. This charge storage mechanism is also supported by the fact that both hydroxyl radical9–11 and hydroxide ion12 functionalization of h-BN have been described in the literature. In the case of the hydroxyl radical the oxygen valence is −1, whereas the oxygen valence is −2 in the case of the hydroxide ion.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c6ra03141b |
This journal is © The Royal Society of Chemistry 2016 |