ArticlePDF Available

Abstract and Figures

Two‐dimensional, Knight‐shifted, T2‐contrasted ²³Na magnetic resonance imaging (MRI) of an all‐solid‐state cell with a Na electrode and a ceramic electrolyte is employed to directly observe Na microstructural growth. A spalling dendritic morphology is observed and confirmed by more conventional post‐mortem analysis; X‐ray tomography and scanning electron microscopy. A significantly larger ²³Na T2 for the dendritic growth, compared with the bulk metal electrode, is attributed to increased sodium ion mobility in the dendrite. ²³Na T2‐contrast MRI of metallic sodium offers a clear, routine method for observing and isolating microstructural growths and can supplement the current suite of techniques utilised to analyse dendritic growth in all‐solid‐state cells.
This content is subject to copyright. Terms and conditions apply.
Batteries
Imaging Sodium Dendrite Growth in All-Solid-State Sodium Batteries
Using 23Na T2-Weighted Magnetic Resonance Imaging
Gregory J. Rees, Dominic Spencer Jolly, Ziyang Ning, T. James Marrow,
Galina E. Pavlovskaya,* and Peter G. Bruce*
Abstract: Two-dimensional, Knight-shifted, T2-contrasted
23Na magnetic resonance imaging (MRI) of an all-solid-state
cell with a Na electrode and a ceramic electrolyte is employed
to directly observe Na microstructural growth. A spalling
dendritic morphology is observed and confirmed by more
conventional post-mortem analysis; X-ray tomography and
scanning electron microscopy. A significantly larger 23Na T2
for the dendritic growth, compared with the bulk metal
electrode, is attributed to increased sodium ion mobility in
the dendrite. 23Na T2-contrast MRI of metallic sodium offers
a clear, routine method for observing and isolating micro-
structural growths and can supplement the current suite of
techniques utilised to analyse dendritic growth in all-solid-state
cells.
All-solid-state batteries (ASSB) with a ceramic electrolyte
and an alkali metal anode could deliver a step-change in
energy storage and safety. The use of solid-state electrolytes
has numerous advantages over the conventional organic
electrolytes such as the ability to use metal anodes, removal of
volatile and flammable electrolyte organics, and they open up
the possibility of Li-Air and Li-Sulphur cathodes (which have
higher volumetric density).[1] These advantages are protracted
when coupled with sodium anodes which allow the use of
aluminium current collectors (whereas more-expensive Cu is
required for Li), sodium also has a significantly higher natural
abundance (2.36% abundance in the earths crust) compared
to that of conventionally used lithium (<0.002%) and,
therefore, offers more security against a volatile Li
market.[1,2] One of the greatest barriers to the progress of
ASSBs is the formation of dendrites (filaments of alkali
metal) on charging that penetrate the ceramic leading to
a short-circuit and cell failure.[3] Dendritic growths in ASSB
systems have different morphologies to their solution coun-
terparts, which can be correlated to the electrochemistry.
These growths have been categorized into four discrete
morphologies; straight, branching, spalling, and diffuse.[4, 5]
Imaging such dendrites is essential to understand their
growth and to develop mechanisms to prevent their forma-
tion. Magnetic resonance imaging (MRI) can provide non-
destructive, isotope specific, structural, time-resolved, and
quantifiable multi-dimensional information.[6] Both 1H and
7Li MRI and magnetic resonance spectroscopy (MRS) have
been utilised to probe dendrites in liquid electrolyte electro-
chemical cells,[7] and 7Li chemical shift imaging has explored
Li microstructural growth in ASSBs.[8] Recent work by Bray
et al. shows that in-operando 23Na MRI and MRS studies on
sodium cells with organic liquid electrolytes were able to
determine the sodium speciation upon galvanostatic cycling.[9]
In-situ 23Na nuclear magnetic resonance (NMR) during
electro-deposition of Na, shows that reversible high-surface-
area mossy and/or dendritic structures can be observed and
attributed to a nucleation mechanism.[12] Direct T2contrast
MRI on any battery material (1H, 6/7Li and 23Na) has never
been investigated. However, NMR relaxometry measure-
ments (spin-lattice relaxation; T1, spin-spin relaxation; T2)
have been vital in understanding ion dynamics in a range of
battery systems.[13]
Imaging solid-state electrolytes is technically more diffi-
cult than their solution counterparts as the NMR linewidths
are substantially broader, this causes the T2/T2*to be short
and hence the signal to dephase during the application of the
imaging gradients. Conventional 1H MRI is not possible in
ASSBs due to the lack of protons in the system and if the
[*] Dr. G. J. Rees, D. Spencer Jolly, Z. Ning, Prof. T. J. Marrow,
Prof. P. G. Bruce
Department of Materials, University of Oxford
Parks Road, Oxford, OX1 3PH (UK)
E-mail: peter.bruce@materials.ox.ac.uk
Prof. P. G. Bruce
Department of Chemistry, University of Oxford
South Parks Road, Oxford, OX1 3QZ (UK)
and
The Henry Royce Institute
Parks Road, Oxford, OX1 3PH (UK)
Dr. G. J. Rees, Prof. P. G. Bruce
The Faraday Institution
Harwell Campus, Didcot, OX11 0RA (UK)
Prof. G. E. Pavlovskaya
Sir Peter Mansfield Imaging Centre, School of Medicine, University of
Nottingham
Nottingham, NG7 2RD (UK)
and
NIHR Nottingham Biomedical Research Centre, University of
Nottingham
Nottingham, NG7 2RD (UK)
E-mail: galina.pavlovskaya@nottingham.ac.uk
Supporting information and the ORCID identification number(s) for
the author(s) of this article can be found under:
https://doi.org/10.1002/anie.202013066.
2020 The Authors. Angewandte Chemie International Edition
published by Wiley-VCH GmbH. This is an open access article under
the terms of the Creative Commons Attribution License, which
permits use, distribution and reproduction in any medium, provided
the original work is properly cited.
A
ngewandte
Chemi
e
Communications
How to cite: Angew. Chem. Int. Ed. 2021,60, 21102115
International Edition: doi.org/10.1002/anie.202013066
German Edition: doi.org/10.1002/ange.202013066
2110 2020 The Authors. Angewandte Chemie International Edition published by Wiley-VCH GmbH Angew. Chem. Int. Ed. 2021,60, 2110 –2115
system did contain protonated groups these would have broad
intrinsic linewidths (kHz) which would dephase during the
application of gradients. Therefore, work in this field has
focussed on the narrower linewidths of 7Li and its distribution
within the solid-electrolyte.[14] Likewise, 23Na MRI is consid-
ered more challenging than 7Li, as it has a significant chemical
shift, greater quadrupole moment, and lower sensitivity than
7Li.[15] However, these larger properties can be utilised to
achieve greater image contrast. Specifically, with a time-
incremented series of MRI experiments, spin-spin (T2)
relaxation can be measured at each spatial element of the
object and an MRI image based on relaxation rather than spin
density can be produced. The achieved spatial contrast
supplies greater information to the images and allows one
to comment on the dynamics or local site symmetry of
individual pixels. Here we report for the first time, T2contrast
23Na MR images of metallic Na electrodes in a pristine state
and after short-circuiting. The T2maps of the electrochemical
cells allow us to comment on the Na-ion dynamics of the
formed features. As 23Na is significantly more difficult than
7Li, this methodology can readily be adapted to more
conventional Li-ASSBs and liquid electrolyte systems.
Two-dimensional spin-echo Knight-shift 23Na MRI images
of the symmetrical all-solid-state Na jNa-b’’-Alumina jNa
cell are shown for the pristine cell in Figure 1c and a cell after
passing a current in Figure 1d, with the dendrite apparent in
the latter. A thorough description of the experimental setup is
given in the SI (Figure S1). These intensity images show that
the amount of signal originating from the dendritic feature is
limited due to the deficiency of 23Na nuclear spins in this
structure, despite this, the dendrite is still observable in
Figure 1 d. Herein lies the major obstacle with materials MRI;
to increase the resolution (or to produce 3D representations)
a smaller area of space is needed to be sampled, however, this
means that the number of nuclear spins contributing to the
signal, in this given area, is reduced and, therefore, the
experimental times increase dramatically. MRI has circum-
vented this issue by employing contrast driven sequences
which can isolate the area of interest and, thus, does not
require such high-resolution images.
The microstructural growth observed in Figure 1d follows
a spalling morphology previously observed by Kazyak and co-
workers, other dendritic morphologies which form in ASSBs
are also discussed.[4] A spalling morphology is formed when
a dendritic crack propagates back to the surface, forming
a conical surface fracture which often precedes the dendrite
fully traversing the solid-electrolyte and causing a short
circuit. Although the dendrite must have extended from
counter to working electrode at the moment of short-circuit
(voltage =0 V in Figure 1a), at the time of imaging there
remains no observable dendrite transversing the electrodes
and only the spallation section of the dendrite is observed.
This is due to the high current passing through the thinnest
parts of the dendrite causing Joule heating, which burns away
the thinnest sections of the dendrite. Growth of a metallic Na
dendrite parallel to the magnetic field causes the NMR signal
to shift to a higher frequency by 5 ppm (from 1126 up to
1131 ppm),[9,12] this is due to orientation dependant nature of
the Knight shift.[16] The shift is smaller than that observed in
7Li as Na has a reduced bulk magnetic susceptibility.[2] The
spalling nature of the microstructural growth seen here has no
distinct directionality, with respect to the magnetic field (B0,
Figure 1b), therefore no obvious
secondary peak formation is
observed in the MRS spectra
(Figure S2b).
The skin depth of sodium is
11 mm at 105 MHz (9.4 T), there-
fore, only the top and bottom
11 mm of metallic Na is observed
in 2D projections. At 9.4 T the
lithium-7 skin depth is compara-
ble, 12 mm. As dendrites are
narrow thin filaments of metallic
species, their formation should
increase the number of 23Na
nuclei visible in the MRS and
MRI experiments as more nuclei
are moved into areas below the
skin-depth and are, therefore,
accessible to the radio-frequency
pulse. There are two possible
reasons for not notably detecting
an increase in 23Na signal, the
first is that large spherical spal-
ling morphology of the dendritic
growth does not reduce the skin
depth and secondly the size of
the parallel component of the
dendrite may be below the limit
Figure 1. a) Galvanostatic cycling of a Na jNa-b’’-Alumina jNa cell under 1 MPa stack-pressure and
a current density 0.5 mA cm2. A charge of 9.5 mAh cm2is passed on the first 1
=
2cycle during which
increased polarisation is observed, consistent with the formation of voids at the electrode as Na is
stripped, as shown previously.[10] On current reversal, the cell short-circuits rapidly consistent with the
formation of dendrites. b) The arrangement of the cell (imaged in the z,yplane) within the magnetic
field (B0), this orientation is conserved throughout all the images. The traditional, intensity, Knight shift
23Na MRI images of the c) pristine cell and d) cell after short-circuit containing a dendritic growth
(highlighted) at the edge of the cell, which is consistent with ion migration along the electric field
lines.[11] The equivalent gradient-echo intensity images are given in Figure S4.
A
ngewandte
Chemi
e
Communications
2111Angew. Chem. Int. Ed. 2021,60, 2110 –2115 2020 The Authors. Angewandte Chemie International Edition published by Wiley-VCH GmbH www.angewandte.org
of detection. The lack of chemical shift change and/or
increase in signal intensity means that conventional 23Na
NMR cannot be utilized to observe dendritic growth in these
ASSBs.
More established ex-situ post-mortem dendrite character-
isation techniques, X-ray computed tomography and scanning
electron microscopy (SEM), were completed to confirm the
nature of the dendrites morphology. The X-ray computed
tomography image for the pristine cell is shown in Figure 2a
and the cell after shorting is given in Figure 2b. A spalling
morphology dendritic formation is observed in the shorted
cell, which is in good agreement with the 23Na MRI. The pixel
size of the tomography images is 4.66 mm and no clear
evidence of a dendrite or crack is observed within the
electrolyte. Likewise, the lack of dendrite observed penetrat-
ing the length of the electrolyte agrees with the MRI results
and suggests that Joule heating has burnt away the dendrite.
The corresponding SEM images are presented in Figure 2c
and an increased resolution image focusing on part of the
spalling feature is given in Figure 2d, with the corresponding
EDX images in Figure 2e. The post-mortem cross-sectional
images show the formation of a spallation crack. This crack is
filled with a metallic material assumed to be Na, however, this
was not confirmed with certainty as although there is a weak
Na signal from the dendritic area, the response is dominated
by a strong C signal, likely due to the reaction of adventitious
carbon species with the highly reactive and newly exposed Na
surfaces.
These images highlight the
difficulties with definitively
imaging dendrites in ASSBs.
The nano-scale of the crack and
dendrite makes visualizing these
formations challenging by X-ray
tomography and although higher
resolution can be achieved, these
are not routinely accessible for
post-mortem cell-failure analysis
or industrial use. In addition,
SEM/EDX is often imperfect as
its cross-sectional nature makes
it a destructive technique, only
able to show a 2D area of a 3D
dendrite and EDX is often not
able to achieve the desired spe-
cificity of chemical information.
The effect of T2weighting the
MR images is shown in Figure 3 a
for the pristine and Figure 3c for
the cell after short-circuit, with
their respective T2histograms
given in Figure 3 b and d. With
current experimental parame-
ters, a full T2map (5 echo incre-
ments) can be achieved in
11.5 hours, using a spin-echo
acquisition scheme. Although
this timescale is too long for
in-situ or in-operando measurements of dendritic growth,
making this method only applicable for post-mortem analysis,
one needs to recall that these measurements were completed
on commercially available equipment (suitable for medical
and biological applications) and optimisation of the coil size,
probe power handling, gradient strength, and magnetic field
strength will all significantly reduce the experiment time. The
superior contrast of the dendrite (Figure 3 c, red region) is due
to a longer T2of the Na nuclei in the growth, this is an isolated
peak in the corresponding distribution histogram (Figure 3 d;
labelled, yellow peak).
The T2exponential decay curves for a range of pixels in
the bulk electrode (Figure 4 c) and the dendrite (Figure 4 a),
with respect to the background noise (Figure 4; given in dark
blue), illustrates the appreciable differences between the T2
relaxation characteristics of the dendrite and the bulk metal
electrodes. The noise level in these images is minor even at
extended echo delays. In the dendrite pixels, a signal to noise
ratio of 4 is achieved despite the lack of 23Na nuclei in these
sites. The T2relaxation of the electrodes varies from 5–10 ms
in both the pristine and cell after short-circuiting, with an
increased full width at half maximum height (FWHM,
Figure 3 b,d) being observed in the cell after short-circuiting.
This FWHM increase is attributed to increased local disorder
of the 23Na nuclei due to the formation of voids in the
electrode.[10,17] The T2relaxation times in the dendritic growth
are significantly longer (>12 ms) with a cluster of T2s at
16 ms (Figure 4 b).
Figure 2. The X-ray tomography images (4.66 mm resolution) of the a) pristine cell and b) cell after
short circuit showing the spalling morphology. Scanning electron micrographs illustrating the dendritic
crack formation with increasing resolution, c) 100 and d) 10 mm scale. e) The corresponding energy-
dispersive X-ray (EDX) images highlighting the distributions of (i) C, (ii) O, (iii) Na, and (iv) Al.
A
ngewandte
Chemi
e
Communications
2112 www.angewandte.org 2020 The Authors. Angewandte Chemie International Edition published by Wiley-VCH GmbH Angew. Chem. Int. Ed. 2021,60, 2110 –2115
Spin-spin (T2) relaxation is a time constant which
describes the signal decay in the transverse (x,y) plane and
is routinely utilized in conventional in-vivo medical 1H MRI
to generate image contrast.[18] The T2relaxation mechanism,
in solids, is attributed to numerous factors such as dipole-
dipole coupling, anisotropies, quadrupole effects, and local
motion. The dominant T2relaxation effect in metals is Pauli
paramagnetism coupled with strong dipole-dipole coupling.[19]
The 100% natural abundance 23Na spins directly couple
through their nuclear magnetic moments, and indirectly
through the intermediary of the conduction electron spins,
the so-called pseudo-dipolar and pseudo-exchange coupling,
respectively.[19] The T2weighted map of the cell after short-
circuit shows significant contrast between the Na electrodes
(T2span =5–12 ms) and the increased relaxation of the
dendrite (T2span =16–17 ms). As the dendrite growth has
reduced dimensions with reduced local symmetry, compared
to the electrode, then one would expect the T2to be shorter
than the bulk Na metal electrode, due to increased quad-
rupolar/anisotropic effects. In metallic NMR, the signals are
broadened when particle sizes are lower than 10 nm.[20] Below
10 nm the local symmetry experienced by the 23Na nuclei will
distort away from cubo-octahedral, causing systematic line-
width broadening as the particle size decreases. This dis-
tortion of the local symmetry will also affect the difference
between the highest occupied molecular orbital (HOMO) and
the lowest unoccupied molecular orbital (LUMO), resulting
in a change in the Knight shift.[21] As there is no observed shift
or broadening in the 23Na peak position from either the
electrode and the dendrite (Figure S2b), then it may be
assumed the dendrite is larger than 10 nm. Incidentally,
tomography images were col-
lected with a 4.66 mm pixel size
and no dendrite was observed in
these images. Therefore, we pro-
pose that the dendrite width
ranges from tens to hundreds of
nanometres. The spalling micro-
structural growth observed in the
23Na images is 23–46 mm in size
(one—two pixels along the z-
axis).
The absence of broadened
features in the NMR suggests
that the increased spin-spin relax-
ation must be dominated by
another effect. We attribute the
substantially increased T2to
increased mobility of the 23Na
nuclei in the dendritic filament,
this mobility would reduce the
dipole-dipole coupling and,
hence, increase the spin-spin
relaxation. The mechanisms gov-
erning self-diffusion in locally
disordered materials are driven
by their defects. Point defects,
such as vacancies or interstitials,
give increased cation self-diffu-
sion.[22] Dislocations, grain boundaries, phase boundaries, and
free surfaces are other types of defects found in solids cause
significant clusters of local defects and act as a diffusion
barrier.[23] Therefore, the observed increase in T2within the
dendrite could be due to a higher concentration of point
defects or a lower concentration of dislocations, grain
boundaries, phase boundaries or free surfaces within the
dendrite compared to the bulk metal electrode, likely due to
the unique conditions of plating Na at a fast rate into a thin
crack with a small aperture.
Regardless of the mechanism, the effect of the different T2
values between the bulk electrode and dendrite is the ability
to resolve one from another and to discretely determine
where a dendrite begins and the electrode stops. The isolation
of the minor peak at a T2of 16 ms (Figure 3d) of the cell
after short-circuit shows that the 23Na nuclei in the dendrite is
in a significantly different dynamic environment to the bulk
metal in the electrode, and offers the opportunity to develop
T2resolved experiments to solely image the dendrite. A
broader T2distribution (Figure 3d) in the electrode region
(5–12 ms) of the shorted cell is evident and can be attributed
to a greater range of local Na site symmetries and the
formation of voids.
The application of T2weighted 23Na MRI is a promising
technique to directly observe the formation and determine
the structural dynamics of dendrites in ASSBs. The drive for
contrast imaging is to remove resolution limits and here,
despite the dendrite being smaller than the resolution of both
23Na MRI and tomography, one can still observe the dendrite
by MRI. The T2of the dendrite is significantly longer than the
T2of the bulk metal electrode. Due to a lack of broadening
Figure 3. The T2weighted contrast maps (with the same orientation and scale as Figure 1c and d) of
a) the pristine symmetrical cell and c) the cell after short-circuit. The dendrite is highlighted and has
a significantly increased T2. The respective total T2distributions are given in the histograms (b) and
(d), with the isolated dendrite distribution highlighted. The individual 23Na images which were used to
produce the T2weighted maps are given in Figure S3.
A
ngewandte
Chemi
e
Communications
2113Angew. Chem. Int. Ed. 2021,60, 2110 –2115 2020 The Authors. Angewandte Chemie International Edition published by Wiley-VCH GmbH www.angewandte.org
observed in the corresponding NMR spectra, the longer T2
must be attributed to increased Na dynamics in the dendrite.
This contrast driven methodology can also be utilized on
liquid electrolyte cells, which have different dendritic mor-
phologies to ASSBs.[24] One limitation of 23Na MRI is the
inability to image the crack development in ASSBs, therefore
we recommend a multimodal imaging approach, combining
X-ray computed tomography to track morphological changes
in the cells, high-resolution elemental content can be achieved
from ex-situ SEM, and dynamic information can be attained
from contrast-driven MRI. This multimodal approach allows
one to image crack formation, microstructural growth, ion-
dynamics, and any dendritic formations.[25]
Acknowledgements
P.G.B. is indebted to the Engineering and Physical Sciences
Research Council (EPSRC), including Next Generation
Solid-State Batteries [EP/P003532/1], Henry Royce Institute
for Advanced Materials [EP/R00661X/1, EP/S019367/1, EP/
R010145/1], and the Faraday Institution All-Solid-State
Batteries with Li and Na Anodes [FIRG007, FIRG008] for
financial support. G.E.P. thanks the Medical Research
Council for funding (Grant # MC_PC_15074) of sodium
imaging methodology development. The X-ray tomography
facilities were funded by an EPSRC Grant [EP/M02833X/1];
“University of Oxford: experimental equipment upgrade”.
Conflict of interest
The authors declare no conflict of interest.
Keywords: all-solid-state electrolytes · batteries ·
magnetic resonance imaging · NMR spectroscopy · sodium
[1] C. Vaalma, D. Buchholz, M. Weil, S. Passerini, Nat. Rev. Mater.
2018,3, 18013.
[2] W. M. Haynes, D. R. Lide, T. J. Bruno, CRC Handbook of
Chemistry and Physics: A Ready-Reference Book of Chemical
and Physical Data, 97th Ed., CRC Press, Boca Raton, FL, 2016.
[3] J. Janek, W. G. Zeier, Nat. Energy 2016,1, 16141.
[4] E. Kazyak, R. Garcia-Mendez, W. S. LePage, A. Sharafi, A. L.
Davis, A. J. Sanchez, K.-H. Chen, C. Haslam, J. Sakamoto, N. P.
Dasgupta, Matter 2020,2, 1025 1048.
[5] J. Garche, C. K. Dyer, P. T. Moseley, Z. Ogumi, D. A. J. Rand, B.
Scrosati, Encyclopedia of Electrochemical Power Sources, Elsev-
ier Science, Amsterdam, 2013.
[6] B. Blumich, NMR imaging of materials, Oxford University Press,
Oxford, 2000.
[7] S. Chandrashekar, N. M. Trease, H. J. Chang, L.-S. Du, C. P.
Grey, A. Jerschow, Nat. Mater. 2012,11, 311 –315 ; A. J. Ilott, M.
Mohammadi, H. J. Chang, C. P. Grey, A. Jerschow, Proc. Natl.
Acad. Sci. USA 2016,113, 10779 10784.
[8] L. E. Marbella, S. Zekoll, J. Kasemchainan, S. P. Emge, P. G.
Bruce, C. P. Grey, Chem. Mater. 2019,31, 2762 2769.
[9] J. M. Bray, C. L. Doswell, G. E. Pavlovskaya, L. Chen, B.
Kishore, H. Au, H. Alptekin, E. Kendrick, M.-M. Titirici, T.
Meersmann, M. M. Britton, Nat. Commun. 2020,11, 2083.
[10] D. Spencer Jolly, Z. Ning, J. E. Darnbrough, J. Kasemchainan,
G. O. Hartley, P. Adamson, D. E. J. Armstrong, J. Marrow, P. G.
Bruce, ACS Appl. Mater. Interfaces 2020,12, 678 685.
[11] R. Failla, M. Bologna, B. Tellini, J. Power Sources 2019,433,
126675.
[12] P. M. Bayley, N. M. Trease, C. P. Grey, J. Am. Chem. Soc. 2016,
138, 1955 1961.
[13] L. Zhou, M. Leskes, T. Liu, C. P. Grey, Angew. Chem. Int. Ed.
2015,54, 14782 14786 ; Angew. Chem. 2015,127, 14995 14999;
A. Grenier, P. J. Reeves, H. Liu, I. D. Seymour, K. Mrker, K. M.
Wiaderek, P. J. Chupas, C. P. Grey, K. W. Chapman, J. Am.
Chem. Soc. 2020,142, 7001 7011; A. Dunst, V. Epp, I. Hanzu,
S. A. Freunberger, M. Wilkening, Energy Environ. Sci. 2014,7,
2739 2752.
[14] S. Chandrashekar, O. Oparaji, G. Yang, D. Hallinan, J. Electro-
chem. Soc. 2016,163, A2988 A2990 ; P.-H. Chien, X. Feng, M.
Tang, J. T. Rosenberg, S. ONeill, J. Zheng, S. C. Grant, Y.-Y. Hu,
J. Phys. Chem. Lett. 2018,9, 1990 1998.
Figure 4. The spin-echo decay curves showing the determined T2
parameters for a range of pixels from a) the dendritic growth and
c) the bulk electrode. The curves are colour coded to the pixels shown
in the expanded 2D T2-contrast image presented in (b). Each pixel
intensity normalised curve is fitted to a mono-exponential with
a constant offset determined by the noise level (given in dark blue).
A
ngewandte
Chemi
e
Communications
2114 www.angewandte.org 2020 The Authors. Angewandte Chemie International Edition published by Wiley-VCH GmbH Angew. Chem. Int. Ed. 2021,60, 2110 –2115
[15] E. G. Sadykhov, Y. A. Pirogov, N. V. Anisimov, M. V. Gulyaev,
G. E. Pavlovskaya, T. Meersmann, V. N. Belyaev, D. V. Fomina,
Appl. Magn. Reson. 2018,49, 925 957.
[16] R. Bhattacharyya, B. Key, H. Chen, A. S. Best, A. F. Hollen-
kamp, C. P. Grey, Nat. Mater. 2010,9, 504 510.
[17] M.-C. Bay, M. Wang, R. Grissa, M. V. F. Heinz, J. Sakamoto, C.
Battaglia, Adv. Energy Mater. 2020,10, 1902899.
[18] G. B. Chavhan, P. S. Babyn, B. Thomas, M. M. Shroff, E. M.
Haacke, RadioGraphics 2009,29, 1433 1449.
[19] C. P. Slichter, Surf. Sci. 1981,106, 382 396.
[20] G. J. Rees, S. T. Orr, L. O. Barrett, J. M. Fisher, J. Houghton,
G. H. Spikes, B. R. C. Theobald, D. Thompsett, M. E. Smith, J. V.
Hanna, Phys. Chem. Chem. Phys. 2013,15, 17195 17207.
[21] J. J. van der Klink, H. B. Brom, Prog. Nucl. Magn. Reson.
Spectrosc. 2000,36, 89 201.
[22] S. Kumarakrishnan, N. L. Peterson, T. O. Mason, J. Phys. Chem.
Solids 1985,46, 1007 1014.
[23] H. Mehrer, S. V. Divinski, Defect Diffus. Forum 2009,289 292,
15 38.
[24] L. Frenck, G. K. Sethi, J. A. Maslyn, N. P. Balsara, Front. Energy
Res. 2019,7, 115.
[25] Supporting research data has been deposited in the Oxford
Research Archive and is available under this DOI: https://doi.
org/10.5287/bodleian:1amymAzYZ.
Manuscript received: September 27, 2020
Accepted manuscript online: October 6, 2020
Version of record online: November 24, 2020
A
ngewandte
Chemi
e
Communications
2115Angew. Chem. Int. Ed. 2021,60, 2110 –2115 2020 The Authors. Angewandte Chemie International Edition published by Wiley-VCH GmbH www.angewandte.org
... Besides, some basic issues surrounding the sodium dendrite propagation paths, charge transfer across the interphase, and the failure mechanism of the solid electrolytes remain unclear. Although advanced techniques such as X-ray computer tomography (XCT), 39,40 time-of-ight secondary ion mass spectrometry (ToFSIMS), 41,42 nuclear magnetic resonance (NMR), 43 in-situ transmission electron microscopy (TEM) apparatus, [44][45][46] etc. have been developed to reveal the internal mechanisms in solid-state batteries. However, these techniques are too precious for the mechanism analysis and the battery examinations in a large scale. ...
Preprint
Full-text available
Dendrite growth in solid-state sodium batteries (SSBs) is one of the most concerned issues that critically affect the battery efficiency and cycling performance. Herein, a fluorescence tomography technology is developed to observe the sodium dendrite growth in SSBs by designing a fluorescent Eu³⁺-doped Na3Zr2Si2PO12 solid electrolyte (SE). Under the Eu³⁺-fluorescence contrast, three-dimensional optical images of the sodium dendrites are obtained by using a confocal laser scanning microscopy. In this way, in-depth sodium dendrite observation during charge/discharge cycles is performed, showing the dendrite initiating stage near the surface and subsequent propagation along the grain boundaries of the SE. Further, a grain-boundary-doping method is promoted and the corresponding Na//Na symmetric cell achieves a record-high cycling stability for more than 1 year (415 d, ongoing) at 25 ℃. This work demonstrates an optical tomography method observing dendrite growth in SSBs and provides an insightful guidance for the design of high-performance SEs.
Article
Sodium batteries (SBs) emerge as a potential candidate for large‐scale energy storage and have become a hot topic in the past few decades. In the previous researches on electrolyte, designing electrolytes with the solvation theory has been the most promising direction is to improve the electrochemical performance of batteries through solvation theory. In general, the four essential factors for the commercial application of SBs, which are cost, low temperature performance, fast charge performance and safety. The solvent structure has significant impact on commercial applications. But so far, the solvation design of electrolyte and the practical application of sodium batteries have not been comprehensively summarized. This review first clarifies the process of Na ⁺ solvation and the strategies for adjusting Na ⁺ solvation. It is worth noting that the relationship between solvation theory and interface theory is pointed out. The cost, low temperature, fast charging, and safety issues of solvation are systematically summarized. The importance of the de‐solvation step in low temperature and fast charging application is emphasized to help select better electrolytes for specific applications. Finally, new insights and potential solutions for electrolytes solvation related to SBs are proposed to stimulate revolutionary electrolyte chemistry for next generation SBs.
Article
Full-text available
Sodium secondary batteries have gained accolades as future energy storage devices due to their low costs and environmental benignity, but are heavily impeded by the poor anodic stabilities of most electrolytes, including solid‐state electrolytes (SSE), that render them incompatible with high‐voltage positive materials. This study reports, for the first time, a new synthesis technique using a fluorohydrogenate ionic liquid (IL)precursor to prepare a [DEME][PF6] ([DEME]⁺: N,N‐diethyl‐N‐methyl‐N‐(2‐methoxyethyl) ammonium) with high yield and high purity. Herein, a Na[PF6]‐[DEME][PF6] IL is formulated and subjected to a series of electrochemical tests to validate its performance in battery applications. The present IL harbors a strong oxidative stability (up to 5.2 V on Pt and >4.5 V on conductive carbon electrodes) that aids in the suppression of oxidative decompositions of one typical SSE, for example, beta alumina solid electrolyte (BASE), thereby extending their electrochemical window in hybrid SSE systems. A hybrid solid‐state Na secondary battery employing a high voltage positive electrode, Na3V2(PO4)2F3, is assembled using the BASE/IL configuration, and features energy density and superior cycling performance. This work demonstrates that sandwiching an SSE between the oxidatively stable [PF6]⁻ IL can be an effective design for high voltage operation Na secondary batteries.
Article
In order to develop new electrode and electrolyte materials for advanced sodium-ion batteries (SIBs), it is crucial to understand a number of fundamental issues. These include the compositions of the bulk and interface, the structures of the materials used, and the electrochemical reactions in the batteries. Solid-state NMR (SS-NMR) has unique advantages in characterizing the local or microstructure of solid electrode/electrolyte materials and their interfaces-one such advantage is that these are determined in a noninvasive and nondestructive manner at the atomic level. In this review, we provide a survey of the recent advances in the understanding of the fundamental issues of SIBs using advanced NMR techniques. First, we summarize the applications of SS-NMR in characterizing electrode material structures and solid electrolyte interfaces (SEI). In particular, we elucidate the key role of in-situ NMR/MRI in revealing the complex reactions and degradation mechanisms of SIBs. Next, the characteristics and shortcomings of SS-NMR and MRI techniques in SIBs are also discussed in comparison to similar Li-ion batteries. Finally, an overview of SS-NMR and MRI techniques for sodium batteries are briefly discussed and presented.
Article
Full-text available
Atomic scale imaging revealed that the deposition/stripping of Na from the β′′-Al 2 O 3 solid electrolyte causes delamination cracks and the closure of the conduction planes.
Article
Full-text available
Sodium-ion batteries are a promising battery technology for their cost and sustainability. This has led to increasing interest in the development of new sodium-ion batteries and new analytical methods to non-invasively, directly visualise battery chemistry. Here we report operando 1H and 23Na nuclear magnetic resonance spectroscopy and imaging experiments to observe the speciation and distribution of sodium in the electrode and electrolyte during sodiation and desodiation of hard carbon in a sodium metal cell and a sodium-ion full-cell configuration. The evolution of the hard carbon sodiation and subsequent formation and evolution of sodium dendrites, upon over-sodiation of the hard carbon, are observed and mapped by 23Na nuclear magnetic resonance spectroscopy and imaging, and their three-dimensional microstructure visualised by 1H magnetic resonance imaging. We also observe, for the first time, the formation of metallic sodium species on hard carbon upon first charge (formation) in a full-cell configuration. Na-ion batteries offer multiple advantages, but there is a critical need for improved materials and understanding of sodiation mechanisms. Here the authors deploy operando 23Na magnetic resonance imaging and spectroscopy to observe sodium battery chemistry and dendrite formation, enabling new insight.
Article
Full-text available
Solid-state electrolytes (SSEs) have attracted substantial attention for next-generation Li-metal batteries, but Li-filament propagation at high current densities remains a significant challenge. This study probes the coupled electrochemical-morphological-mechanical evolution of Li-metal-Li7La3Zr2O12 interfaces. Quantitative analysis of synchronized electrochemistry with operando video microscopy reveals new insights into the nature of Li propagation in SSEs. Several different filament morphologies are identified, demonstrating that a singular mechanism is insufficient to describe the complexity of Li propagation pathways. The dynamic evolution of the structures is characterized, which demonstrates the relationships between current density and propagation velocity, as well as reversibility of plated Li before short-circuit occurs. Under deep discharge, void formation and dewetting are directly observed, which are directly related to evolving overpotentials during stripping. Finally, similar Li penetration behavior is observed in glassy Li3PS4, demonstrating the relevance of the new insights to SSEs more generally.
Article
Full-text available
All‐solid‐state batteries with an alkali metal anode have the potential to achieve high energy density. However, the onset of dendrite formation limits the maximum plating current density across the solid electrolyte and prevents fast charging. It is shown that the maximum plating current density is related to the interfacial resistance between the solid electrolyte and the metal anode. Due to their high ionic conductivity, low electronic conductivity, and stability against sodium metal, Na‐β″‐alumina ceramics are excellent candidates as electrolytes for room‐temperature all‐solid‐state batteries. Here, it is demonstrated that a heat treatment of Na‐β″‐alumina ceramics in argon atmosphere enables an interfacial resistance <10 Ω cm² and current densities up to 12 mA cm⁻² at room temperature. The current density obtained for Na‐β″‐alumina is ten times higher than that measured on a garnet‐type Li7La3Zr2O12 electrolyte under equivalent conditions. X‐ray photoelectron spectroscopy shows that eliminating hydroxyl groups and carbon contaminations at the interface between Na‐β″‐alumina and sodium metal is key to reach such values. By comparing the temperature‐dependent stripping/plating behavior of Na‐β″‐alumina and Li7La3Zr2O12, the role of the alkali metal in governing interface kinetics is discussed. This study provides new insights into dendrite formation and paves the way for fast‐charging all‐solid‐state batteries.
Article
Full-text available
Lithium metal is a promising anode material for next-generation rechargeable batteries, but non-uniform electrodeposition of lithium is a significant barrier. These non-uniform deposits are often referred to as lithium “dendrites,” although their morphologies can vary. We have surveyed the literature on lithium electrodeposition through three classes of electrolytes: liquids, polymers and inorganic solids. We find that the non-uniform deposits can be grouped into six classes: whiskers, moss, dendrites, globules, trees, and cracks. These deposits were obtained in a variety of cell geometries using both unidirectional deposition and cell cycling. The main result of the study is a figure where the morphology of electrodeposited lithium is plotted as a function of two variables: shear modulus of the electrolyte and current density normalized by the limiting current density. We show that specific morphologies are confined to contiguous regions on this two-dimensional plot.
Article
Full-text available
All-solid-state batteries potentially offer safe, high-energy-density electrochemical energy storage, yet are plagued with issues surrounding Li microstructural growth and subsequent cell death. We use 7Li NMR chemical shift imaging and electron microscopy to track Li microstructural growth in the garnet-type solid electrolyte, Li6.5La3Zr1.5Ta0.5O12. Here, we follow the early stages of Li microstructural growth during galvanostatic cycling, from the formation of Li on the electrode surface to dendritic Li connecting both electrodes in symmetrical cells, and correlate these changes with alterations observed in the voltage profiles during cycling and impedance measurements. During these experiments, we observe transformations at both the stripping and plating interfaces, indicating heterogeneities in both Li removal and deposition. At low current densities, 7Li magnetic resonance imaging detects the formation of Li microstructures in cells before short-circuits are observed and allows changes in the electrochemical profiles to be rationalized.
Article
Full-text available
Sodium is a key element in a living organism. The increase of its concentration is an indicator of many pathological conditions. ²³Na magnetic resonance imaging (MRI) is a quantitative method that allows to determine the sodium content in tissues and organs in vivo. This method has not yet entered clinical practice widely, but it has already been used as a clinical research tool to investigate diseases such as brain tumors, breast cancer, stroke, multiple sclerosis, hypertension, diabetes, ischemic heart disease, osteoarthritis. The active development of the ²³Na MRI is promoted by the growth of available magnetic fields, the expansion of hardware capabilities, and the development of pulse sequences with ultra-short echo time.
Article
Substituted Li-layered transition-metal oxide (LTMO) electrodes such as Li x Ni y Mn z Co1-y-zO2 (NMC) and Li x Ni y Co1-y-zAl z O2 (NCA) show reduced first cycle Coulombic efficiency (90-87% under standard cycling conditions) in comparison with the archetypal Li x CoO2 (LCO; ∼98% efficiency). Focusing on Li x Ni0.8Co0.15Al0.05O2 as a model compound, we use operando synchrotron X-ray diffraction (XRD) and nuclear magnetic resonance (NMR) spectroscopy to demonstrate that the apparent first-cycle capacity loss is a kinetic effect linked to limited Li mobility at x > 0.88, with near full capacity recovered during a potentiostatic hold following the galvanostatic charge-discharge cycle. This kinetic capacity loss, unlike many capacity losses in LTMOs, is independent of the cutoff voltage during delithiation and it is a reversible process. The kinetic limitation manifests not only as the kinetic capacity loss during discharge but as a subtle bimodal compositional distribution early in charge and, also, a dramatic increase of the charge-discharge voltage hysteresis at x > 0.88. 7Li NMR measurements indicate that the kinetic limitation reflects limited Li transport at x > 0.86. Electrochemical measurements on a wider range of LTMOs including Li x (Ni,Fe) y Co1-yO2 suggest that 5% substitution is sufficient to induce the kinetic limitation and that the effect is not limited to Ni substitution. We outline how, in addition to a reduction in the number of Li vacancies and shrinkage of the Li-layer size, the intrinsic charge storage mechanism (two-phase vs solid-solution) and localization of charge give rise to additional kinetic barriers in NCA and nonmetallic LTMOs in general.
Article
3-electrode studies coupled with tomographic imaging of the Na/Na-β”-alumina interface reveal that voids form in the Na metal at the interface on stripping and they accumulate on cycling, leading to increasing interfacial current density, dendrite formation on plating, short circuit and cell failure. The process occurs above a critical current for stripping (CCS) for a given stack pressure, which sets the upper limit on current density that avoids cell failure, in line with results for the Li/solid-electrolyte interface. The pressure required to avoid cell failure varies linearly with current density, indicating that Na creep rather than diffusion per se dominates Na transport to the interface and that significant pressures are required to prevent cell death, > 9 MPa at 2.5 mA·cm⁻².
Article
In this paper, we present a new model of dendrite growth in battery cells through theoretical and numerical analysis. We use a statistical model based on the competition between a deterministic electric field and a stochastic force, which both drive the movement of the particles inside the battery cell. The simulation of the dendrite growth is modeled via Diffusion Limited Aggregation. As a major aspect, we point out the key role played by the intense electric field close to the edges of the electrodes, which statistically drives the particles along the electric field lines.