Sie sind auf Seite 1von 9

Visualizing electromagnetic fields in metals by MRI

Chandrika Sefcikova Chandrashekar, Annadanesh Shellikeri, S. Chandrashekar, Erika A. Taylor, and


Deanne M. Taylor

Citation: AIP Advances 7, 025310 (2017); doi: 10.1063/1.4977700


View online: http://dx.doi.org/10.1063/1.4977700
View Table of Contents: http://aip.scitation.org/toc/adv/7/2
Published by the American Institute of Physics

Articles you may be interested in


Phonon band structures of the three dimensional latticed pentamode metamaterials
AIP Advances 7, 025309025309 (2017); 10.1063/1.4977715
AIP ADVANCES 7, 025310 (2017)

Visualizing electromagnetic fields in metals by MRI


Chandrika Sefcikova Chandrashekar,1 Annadanesh Shellikeri,2
S. Chandrashekar,3,a Erika A. Taylor,4 and Deanne M. Taylor5,6,7
1 LincolnHigh School, 3838 Trojan Trail, Tallahassee, Florida 32311, USA
2 Aeropropulsion, Mechatronics and Energy Center, Florida State University,
2003 Levy Ave., Tallahassee, Florida 32310, USA
3 National High Magnetic Field Laboratory (NHMFL) and Florida State University,

1800 E. Paul Dirac Drive, Tallahassee, Florida 32310, USA


4 Department of Chemistry, Wesleyan University, 52 Lawn Ave., Hall-Atwater Labs,

Middletown, Connecticut 06459, USA


5 Department of Pediatrics, Perelman School of Medicine, University

of Pennsylvania, Philadelphia, Pennsylvania 19104, USA


6 Department of Biomedical and Health Informatics, The Childrens Hospital

of Philadelphia, Philadelphia, Pennsylvania 19041, USA


7 Department of Genetics, Rutgers University, Piscataway, New Jersey 08854, USA

(Received 29 July 2016; accepted 9 February 2017; published online 27 February 2017)

Based upon Maxwells equations, it has long been established that oscillating electro-
magnetic (EM) fields incident upon a metal surface, decay exponentially inside the
conductor, leading to a virtual absence of EM fields at sufficient depths. Magnetic
resonance imaging (MRI) utilizes radiofrequency (r.f.) EM fields to produce images.
Here we present a visualization of a virtual EM vacuum inside a bulk metal strip by
MRI, amongst several findings. At its simplest, an MRI image is an intensity map of
density variations across voxels (pixels) of identical size (=x y z). By contrast in
bulk metal MRI, we uncover that despite uniform density, intensity variations arise
from differing effective elemental volumes (voxels) from different parts of the bulk
metal. Further, we furnish chemical shift imaging (CSI) results that discriminate dif-
ferent faces (surfaces) of a metal block according to their distinct nuclear magnetic
resonance (NMR) chemical shifts, which holds much promise for monitoring surface
chemical reactions noninvasively. Bulk metals are ubiquitous, and MRI is a premier
noninvasive diagnostic tool. Combining the two, the emerging field of bulk metal MRI
can be expected to grow in importance. The findings here may impact further develop-
ment of bulk metal MRI and CSI. 2017 Author(s). All article content, except where
otherwise noted, is licensed under a Creative Commons Attribution (CC BY) license
(http://creativecommons.org/licenses/by/4.0/). [http://dx.doi.org/10.1063/1.4977700]

I. INTRODUCTION
MRI is a household name as a diagnostic tool in the medical field,1,2 with an impressive resume
in many other fields, including the study of materials,36 corrosion of metals, monitoring batteries
and supercapacitors.7,8
However, historically, MRI of bulk metals has been very rare (as detailed in supplementary
material Section S2), due to challenges posed by the physics of propagation of radiofrequency (r.f.)
magnetic field (B1 ) in bulk metals. Chief among them are the rapidly decaying B1 inside the metal
(Fig.1), and orientation of the bulk metal surface relative to incident B1 .911 Recent studies1218 have
overcome these difficulties to demonstrate and highlight bulk metal NMR and MRI, albeit, primarily
applied to batteries and electrochemical cells (section S2).
Here we present several findings on bulk metal MRI and CSI:

a
Corresponding author: cshekar@fsu.edu

2158-3226/2017/7(2)/025310/8 7, 025310-1 Author(s) 2017


025310-2 Chandrashekar et al. AIP Advances 7, 025310 (2017)

FIG. 1. Skin depth , and effective subsurface depth eff . (a) Exponential decay of magnitude of B1 (y), as a function of the
depth from the metal surface (see Eq. (S1) in section S3). (b) Effective subsurface depth eff , without B1 decay, that would
account for the MR signal (see Eq. (S7) in section S3).

During a systematic noninvasive thickness measurement of bulk metal strips by MRI15 (section
S4), we come across unexpected regions of intensity, and assign them to two mutually orthogonal
pairs of faces of the strip.
To explain the peculiar ratios of intensities from corresponding subsurfaces in bulk metal
MRI and CSI, we derive formulae from first principles, unveiling a surprising underlying
reason for the observed intensity patterns: Effective elemental volumes are different for these
subsurfaces.
In the process, the images enable a visualization of a virtual absence of EM fields inside the
bulk metal via an MRI tunnel.
Additionally, we demonstrate that the bulk metal CSI distinguishes different faces (surfaces) of
a metal block according to their distinct NMR chemical shifts.
We attained these by employing three phantoms (samples) P0, P1, P3, depicted schematically
in Fig.S1 and described in Methods section S1.1. All phantoms were derived from the same stock
of 0.75 mm thick lithium (Li) strip. Phantom P3 is a super strip composed of three Li strips pressed
together, forming an effective single strip three times thicker than the individual strips in phantoms
P0 and P1.
The setup of phantoms, r.f. coil and the gradient assembly ensures that the imaging directions x,
y, z fulfill the condition that x k a k B1 and z k B0 (the static main magnetic field), with the possibility
to reorient the phantom about the x-axis; a, b, c are the three sides of the strips.
Since all MRI and CSI images to follow were acquired with the given phantoms bc faces normal
(, perpendicular) to B1 , the images have no contribution from these faces to the magnetic resonance
(MR) signal,1316 since B1 penetration into the metal is maximal when it is parallel (k) to the metal
surface, and minimal when k metal surface.9,13,14
For details on the MRI experiments, including the nomenclature, the reader is referred to Methods
section S1.2.

II. MRI
Fig. 2 furnishes stackplots (intensity along the vertical axis) of 7 Li 2d MRI (without slice
selection) of phantom P3. Panel (a) displays MRI(xy). Panel (b) displays MRI(yz).
It is straightforward to infer that the walls of high intensity regions in either image emanate from
ac faces of the P3 strip,1315 as we did while measuring the thickness of metal strips (Figs.S2 and S3,
section S4). In either image, contributions from spatially unresolved dimensions sum up to yield the
high intensity walls.
However, the unexpected intensity between the two ac faces of the super strip, in both the images,
is perplexing. The 2d MRI(xy) in Fig. 2a exhibits a low intensity plateau spanning the walls. The 2d
MRI(yz) in Fig. 2b displays low intensity ridges bridging the walls.
025310-3 Chandrashekar et al. AIP Advances 7, 025310 (2017)

FIG. 2. Surprising regions of intensity in bulk metal MRI. 7 Li 2d MRI (sans slice selection) stack plots of phantom P3 (Fig.S1).
Vertical axis denotes intensity (arbitrary units), resulting from the sum total of spin density along non imaged dimension. (a)
2d MRI(xy); same data as in blue, labeled P3, regions of Fig.S3b. (b) 2d MRI(yz). In either image, there is no contribution from
the bc faces to the MR signal since they are B1 (section I). In either image, the high intensity walls are easily associated
with the ac faces (see Figs.S2, S3 and section S4). The intensity between the two ac faces of the metal strip is evident in both
images. Note the low intensity plateau and the ridges between the walls, respectively in the MRI(xy) and MRI(yz) images.
See section II and Fig. S4.

To understand better these unexpected regions of intensity, we acquired 7 Li 3d MRI(xyz) of


phantom P3, shown in Fig. 3. In addition to the ac faces (separated along y), the ab faces (separated
along z) are revealed (not previously reported).
As noted earlier, bc faces (being k to B1 ) are absent. The hollow region in MRI(xyz) arises
from the skin depth phenomenon,914 restricting the EM fields to effectively access only a limited
subsurface underneath the ac and ab faces (Fig. 1 and section S3). The presence of faces k B1 ,
coupled with the conspicuous absence of faces k B1 , in combination with the hollow region, imparts
the 3d image an appearance of an MRI tunnel, supplying a visualization of a virtual absence of
EM fields in the interior of a metallic conductor, hitherto depicted only schematically in literature
(for e.g., Ref.12).
With the aid of 3d MRI in Fig. 3, the intensity regions in Fig. 2a and Fig. 2b, can be easily
interpreted as simply regions resulting respectively from projections along z and x axes of the 3d
image. It is clear that the intensity between ac faces (either the plateau or the ridges), is due to the
pair of ab faces of the superstrip P3.
Still, the basis for the relative intensity values remains elusive at this stage.
For the 2d MRI(xy) in Fig. 2a, it can be argued that, for the ac face the entire length of side
c = 7 mm (Fig.S1) contributes to the signal, while for the ab face, only a subsurface depth eff 9.49 m

FIG. 3. Peering at a virtual electromagnetic vacuum inside metals, through an MRI tunnel. 7 Li 3d MRI(xyz) of phantom
P3 (Fig.S1). In addition to the already identified ac faces in previous images Figs. (2, S2 and S3), the ab faces are revealed
(hitherto not reported), accounting for the low intensity regions in Fig. 2. The bc faces are conspicuous by absence, being
B1 . The resulting MRI tunnel supplies a visual of peering at a virtual absence of EM fields in the interior of a metallic
conductor, since the center of the strip is at least (2.25mm/2)/(6.9m) = 163 skin-depths away from the surface (and hence
any EM fields and the MR signal would be virtually buried in noise; see Fig. 1 and section II).
025310-4 Chandrashekar et al. AIP Advances 7, 025310 (2017)

contributes (Eq.(S2), Eq.(S7), section S3 and Fig. 1). This would lead to an expected (from skin-
depth arguments) ratio of the corresponding intensities, S ac /S ab , to be of the order of c/(2eff ) 368
(Fig.S4a), in obvious disagreement with the observed ratio (of maxima of S ac and S ab ) of 6.6.
For the 2d MRI(yz) in Fig. 2b, the expected intensity pattern in a stack plot would be one with
equal intensities from ab and ac faces, since they share the same side, a, along x (non imaged) axis
(Fig.S4b). This again, is in stark contrast with the observed ratio (of maxima of S ac and S ab ) of 10.
For the MRI(xyz), naively, uniform intensity would be expected from both ab and ac faces,
resulting in a ratio of unity. Instead, the observed ratio (of maxima of S ac and S ab ) is found to be 3.8.
Thus, the MRI images bear peculiar intensity ratios from comfortably identified (from 3d MRI)
regions of the bulk metal. We will return to this topic later.

III. CSI
The 7 Li NMR spectrum of phantom P3 (Fig. 4 inset) contains two distinct peaks in the Knight
shift region for metallic 7 Li (see Methods section S1.2), centered at 1 = 256.4 and 2 = 266.3 ppm.
At first sight it might seem odd that a metallic strip, of regular geometry and uniform density, that is
entirely composed of identical Li atoms, gives rise to two NMR peaks instead of the expected single
peak.
To gain additional insight as to the spatial distribution of the Li metal species with different NMR
shifts, we turn to CSI, which combines an NMR chemical shift (CS) dimension with one or more
imaging(I) dimensions.3,1921
The 2d CSI(y) shown in Fig.S5 comprises of two bands separated along y located at 2 , along
the CS dimension, while a low intensity band spans them along y at a CS of 1 , strongly hinting at,
the two bands (at 2 ) being associated with ac faces.
This observation called for adding an additional imaging dimension along z, leading to 3d
CSI(yz), which is realized in Fig. 4, where y and z are the imaging dimensions, accompanied by
the CS dimension. The bands separated along z, occur at 1 . The bands separated along y, occur at
2 . In conjunction with the 3d MRI image in Fig. 3, it is evident that the pairs of bands at 1 and 2
arise from ab and ac faces respectively, completing the spatio-chemical assignment.

FIG. 4. Chemical shift imaging of a bulk metal strip. 7 Li 3d CSI(yz) of phantom P3: chemical shift, y and z axes comprise
the three dimensions. Despite being composed of identical Li atoms, 7 Li NMR spectrum (inset) of phantom P3 exhibits two
peaks, instead of the expected single peak (section III). Short and tall peaks are centered respectively at 1 = 256.4 and 2 =
266.3 ppm. In the CSI, bands separated along z, occur at CS 1 . The bands separated along y, occur at CS 2 . In conjunction
with the 3d MRI (Fig. 3) and P3 schematics (Fig.S1), the pair of bands at 1 are assigned to ab faces, and pair of bands at 2
are assigned to ac faces (thus completing the assignment of both the 2d CSI(y) in Fig.S5, and the NMR spectrum itself).
025310-5 Chandrashekar et al. AIP Advances 7, 025310 (2017)

These assignments readily carry over to 2d CSI(y) in Fig.S5, with the pair of bands at 2 and the
low ridge spanning them at 1 , being respectively identified with ac and ab faces. Similarly, in the
NMR spectrum, the short and tall peaks respectively at 1,2 are assigned to ab and ac faces, consistent
with the reported12,14 experiments and simulations.
Thus the chemical (Knight) shift for a given face of a bulk metal strip depends on its orientation
relative to B0 , due to bulk magnetic susceptibility (BMS) effect.1214,17,2224
Interestingly, the 3d CSI sheds new light on previous bulk metal NMR studies. For instance, in
an earlier study,12 a similar shift difference between NMR peaks was observed at kand orientations
(relative to B0 ) of the major faces of a thinner metal strip, by carrying out two separate experiments.
Here, phantom P3 furnishes these two orientations in a single experiment, via ac and ab faces
(Fig.S1). The present work provides physical insight into another previous14 observation. It was
found that, under rotation about z k c k B0 axis, the intensity of NMR peak was invariant only for
ab faces. Consulting the 3d MRI (Fig. 3) and 3d CSI (Fig. 4), it is easy to see why: Under this
rotation, only for ab faces, orientation of B1 is unchanged. Since NMR peak intensity for a given
face depends on relative orientation to B1 ,13,14 it remains constant only for the ab faces. On the other
hand, the orientation of the faces relative to B0 , is invariant under this rotation: ac and bc faces remain
k B0 , while ab faces remain B0 . Hence the chemical shifts remain constant, since they depend on
the B0 orientation relative to the faces.
Like for MRI, the basis for the ratio of intensities from the ac and ab faces (S ac /S ab 2.8), is
not immediately intuitively obvious and will be explored next.

IV. INTENSITY RATIO FORMULAE FOR BULK METAL MRI AND CSI
The peculiar intensity ratios in MRI and CSI, of signals S ab and S ac , arising respectively from ab
and ac faces of phantom P3 (Fig.2, sections II and III), could be due to gradient switching involved
in the MRI experiments (the resultant eddy currents could be different for ab and ac faces). However,
as shown in section S5, this can be ruled out on the basis of 2d MRI(yz) and MRI(zy) at mutually
orthogonal orientations (related by a rotation about x k a k B1 ), shown in Fig.S6.
And yet, it is possible to derive, from first principles, expressions for the ratios of MRI and CSI
intensities from ab and ac faces.
For the 2d MRI(xy), in Fig. 2a, the signal intensity from the ab faces can be written as (see
Eq.(S8))

Sab (x, y) dx dy dz = dx dy 2eff (1)

with dx dy dz denoting elemental volume of the metal, and eff is the effective subsurface depth that
would account for the MR signal in the absence of B1 decay (see Eq.(S7) and Fig.1). Above, the
integral over z, is replaced by eff underneath the two ab faces separated along z.
Similarly, for the signal intensity from either of the ac faces,

Sab (x, y) dx dy dz = dx eff c (2)

since the subsurface now is y.


Eq.(1) and Eq.(2) reveal differing effective elemental volumes (voxels) underneath these faces:
ab
dVeff = dx dy eff (3)

ac
dVeff = dx eff dz (4)
From Eq.(1) and Eq.(2),
Sac c
= (5)
Sab 2y
where we have replaced dy by y, the resolution along y k b. Consulting the Methods section S1.2
and Fig.S1, c = 7 mm, y = 0.25 mm and Eq.(5) yields a calculated ratio of S ac /S ab = 14 (as illustrated
025310-6 Chandrashekar et al. AIP Advances 7, 025310 (2017)

in Fig.S7a), within an order of magnitude of the observed ratio (section II, Fig. 2a) and a 25 fold
improvement relative to the expected ratio (Fig. S4a).
Continuing in the same vein, for the 2d MRI(yz), in Fig. 2b, it can be easily shown that
(Section S6)
Sac z
= (6)
Sab y
with, y, z, being the respective resolutions along y, z (see also Fig. S7b). Since y , z, Eq. (6)
explains the departure from the expected equality S ac = S ab (Fig. S4a).
As detailed in Section S6 and Table S1, the S ac /S ab ratios calculated from Eqs.(5), (6) and (S11),
differ by less than an order of magnitude from the observed values for 2d MRI(xy), 2d MRI(yz),
2d MRI(zy), MRI(xyz), 3d CSI(yz) and 2d CSI(y). An interesting feature is that, for 2d MRI(xy),
Eq.(5) reveals that S ac /S ab increases with increasing resolution along y, affirmed experimentally
in Fig.S8. Similarly, for 3d MRI(xyz), the nonintuitive fact that S ab , S ac is visually demonstrated
in Fig.S9.
In summary, the formulae unveil the underlying reason (differing effective elemental volumes
underneath these faces) for the observed bulk metal MRI and CSI intensity patterns. The derived
patterns bear closer resemblance to experiment than what is expected from skin depth considerations
alone, or from conventional specifications of the voxel= x y z (see for e.g. Figs. 2, S4, and S7).
Moreover, the analysis presented here may throw light on previous bulk metal MRI studies. For
instance, consider the MRI in k orientation in an earlier study:14 with c = 10mm and y = 31m,
Eq.(5) yields S ac /S ab = 161, rendering the contribution from the ab faces all but invisible in those
images. It would be interesting to see future studies involving EM simulations,14,25,26 B1 maps,27
etc., to gain additional insight about EM field distribution around bulk metals and help understand
the remaining gap between the observed and derived intensity ratios, and to see to what extent the
effective voxel size emerges.
On a practical note, these formulae can guide experimental strategies to relatively enhance MRI
and CSI signals from different regions of the bulk metal.

V. CONCLUSIONS
In conclusion, the unexpected findings presented here may impact bulk metal MRI and CSI
studies in general, via fresh insights for data collection, analysis and interpretation. The bulk metal
MRI and CSI results in this study have the noninvasive diagnostic potential for structure of metals and
alloys,28,29 metallurgy (metal fatigue, fracture, strain),3032 catalysis,33,34 bulk metal surface science
and surface chemistry,3537 metallic medical implants, dielectric MRI in the vicinity of bulk metals
(section S1).
Given the basic nature of the findings, the approach presented here is likely to benefit from
advances made in the mainstream (medical) MRI field.

SUPPLEMENTARY MATERIAL
See supplementary material for supporting text and figures.

ACKNOWLEDGMENTS
A portion of this work was performed at the National High Magnetic Field Laboratory, which is
supported by National Science Foundation Cooperative Agreement No. DMR1157490 and the State
of Florida. SC thanks Dr. Daniel T. Hallinan Jr. and Dr. WilliamW. Brey, of FSU, and Dr. Hailong
Chen of Georgia Tech., for helpful discussions and support. CSC recognized the anomaly between
observed and expected intensity patterns in the images, prepared figures, did literature search. AS
prepared the phantoms. SC conceptualized and designed the project, conducted experiments, derived
intensity ratio formulae. CSC and SC analyzed and interpreted results. All authors discussed results,
helped write and review the manuscript.
The authors declare no competing financial interests.
025310-7 Chandrashekar et al. AIP Advances 7, 025310 (2017)

1 V. Perrin, MRI Techniques (John Wiley & Sons, 2013).


2 T. E. Yankeelov, D. R. Pickens, and R. R. Price, Quantitative MRI in Cancer (CRC Press, 2011).
3 P. T. Callaghan, Principles of Nuclear Magnetic Resonance Microscopy (Oxford University Press, 2003).
4 B. Blumich, NMR Imaging of Materials (Oxford University Press, 2003).
5 V. S. Bajaj, J. Paulsen, E. Harel, and A. Pines, Zooming in on microscopic flow by remotely detected MRI, Science 330,

10781081 (2010).
6 D. Rugar, H. J. Mamin, M. H. Sherwood, M. Kim, C. T. Rettner, K. Ohno, and D. D. Awschalom, Proton magnetic

resonance imaging using a nitrogen-vacancy spin sensor, Nature Nanotechnology 10, 120124 (2015).
7 A. J. Davenport, M. Forsyth, and M. M. Britton, Visualisation of chemical processes during corrosion of zinc using magnetic

resonance imaging, Electrochem. Commun. 12, 4447 (2010).


8 A. J. Ilott, N. M. Trease, C. P. Grey, and A. Jerschow, Multinuclear in situ magnetic resonance imaging of electrochemical

double-layer capacitors, Nature Communications (2014).


9 J. D. Jackson, Classical Electrodynamics, 3rd edition (John Wiley & Sons, Inc., 1990).
10 D. J. Griffiths, Introduction to Electrodynamics, 3rd edition (Pearson Education Inc., Upper Saddle River, New Jersey,

07458, USA, 1999).


11 F. T. Ulaby, Fundamentals of Applied Electromagnetics, 2001 media edition (Prentice-Hall Inc., Upper Saddle River, New

Jersey, 07458, USA, 1997).


12 R. Bhattacharyya, B. Key, H. Chen, A. S. Best, A. F. Hollenkamp, and C. P. Grey, In situ NMR observation of the formation

of metallic lithium microstructures in lithium batteries, Nature Materials 9, 504510 (2010).


13 S. Chandrashekar, N. M. Trease, H. J. Chang, L.-S. Du, C. P. Grey, and A. Jerschow, 7 Li MRI of Li batteries reveals

location of microstructural lithium, Nature Materials 11, 311315 (2012).


14 A. J. Ilott, S. Chandrashekar, A. Kloeckner, H. J. Chang, N. M. Trease, C. P. Grey, L. Greengard, and A. Jerschow,

Visualizing skin effects in conductors with MRI: 7 Li MRI experiments and calculations, J. Magn. Reson. 245, 143149
(2014).
15 K. Romanenko, M. Forsyth, and L. A. ODell, New opportunities for quantitative and time efficient 3D MRI of liquid and

solid electrochemical cell components: Sectoral fast spin echo and SPRITE, J. Magn. Reson. 248, 96104 (2014).
16 M. M. Britton, Magnetic resonance imaging of electrochemical cells containing bulk metal, Chem. Phys. Chem. 15,

17311736 (2014).
17 H. J. Chang, A. J. Ilott, N. M. Trease, M. Mohammadi, A. Jerschow, and C. P. Grey, Correlating microstructural lithium

metal growth with electrolyte salt depletion in lithium batteries using 7 Li MRI, J. Am. Chem. Soc. 137, 1520915216
(2015).
18 S. Chandrashekar, O. Oparaji, G. Yang, and D. Hallinan, Jr., 7 Li MRI unveils concentration dependent diffusion in polymer

electrolyte batteries, J. Electrochem. Soc. 163, A2988A2990 (2016).


19 E. Mark Haacke, R. E. Brown, M. R. Thompson, and R. Venkatesan, Magnetic Resonance Imaging: Physical Principles

and Sequence Design (John Wiley & Sons, Inc., 1999).


20 N. Spengler, J. Hoefflin, A. Moazenzadeh, D. Mager, N. MacKinnon, V. Badilita, U. Wallrabe, and J. G. Korvink, Het-

eronuclear micro-Helmholtz coil facilitates m-range spatial and sub-Hz spectral resolution NMR of nL-volume samples
on customisable microfluidic chips, PLoS ONE 11 (2016).
21 K. W. Feindel, Spatially resolved chemical reaction monitoring using magnetic resonance imaging, Magn. Reson. Chem.

(2015).
22 R. E. Hoffman, Measurement of magnetic susceptibility and calculation of shape factor of NMR samples, J. Magn. Reson

178, 237247 (2006).


23 L. Zhou, M. Leskes, A. J. Ilott, N. M. Trease, and C. P. Grey, Paramagnetic electrodes and bulk magnetic susceptibil-

ity effects in the in situ NMR studies of batteries: application to Li1.08 Mn1.92 O4 spinels, J. Magn. Reson 234, 4457
(2013).
24 H. J. Chang, N. M. Trease, A. J. Ilott, D. Zeng, L.-S. Du, A. Jerschow, and C. P. Grey, Investigating Li microstructure

formation on Li anodes for lithium batteries by in situ 6 Li/7 Li NMR and SEM, J. Phys. Chem. C 119, 1644316451 (2015).
25 R. R. Mett and J. S. Hyde, Aqueous flat cells perpendicular to the electric field for use in electron paramagnetic resonance

spectroscopy, J. Magn. Reson. 165, 137152 (2003).


26 J. W. Sidabras, R. R. Mett, and J. S. Hyde, Aqueous flat cells perpendicular to the electric field for use in electron

paramagnetic resonance spectroscopy, II: Design, J. Magn. Reson. 172, 333341 (2005).
27 M. A. Cloos, F. Knoll, T. Zhao, K. T. Block, M. Bruno, G. C. Wiggins, and D. K. Sodickson, Multiparametric imaging

with heterogeneous radiofrequency fields, Nature Communications (2016).


28 J. A. Rodriguez and D. Wayne Goodman, The nature of the metal-metal bond in bimetallic surfaces, Science 257, 897903

(1992).
29 C. P. Flynn, Point defect reactions at surfaces and in bulk metals, Physical Review B 71, 085422 (2005).
30 L. Gireaud, S. Grugeon, S. Laruelle, B. Yrieix, and J. M. Tarascon, Lithium metal stripping/plating mechanisms studies:

A metallurgical approach, Electrochem. Commun. 8, 16391649 (2006).


31 B. P. P. A. Gouveia, J. M. C. Rodrigues, and P. A. F. Martins, Fracture predicting in bulk metal forming, International

Journal of Mechanical Sciences 38, 361372 (1996).


32 M. Mavrikakis, B. Hammer, and J. K. Norskov, Effect of strain on the reactivity of metal surfaces, Physical Review Letters

81, 28192822 (1998).


33 C.-J. Zhong, J. Luo, B. Fang, B. N. Wanjala, P. N. Njoki, R. Loukrakpam, and J. Yin, Nanostructured catalysts in fuel

cells, Nanotechnology 21, 062001 (2010).


34 F.-W. Yuan, H.-J. Yang, and H.-Y. Tuan, Seeded silicon nanowire growth catalyzed by commercially available bulk metals:

Broad selection of metal catalysts, superior field emission performance, and versatile nanowire/metal architectures, Journal
of Materials Chemistry 21, 1379313800 (2011).
025310-8 Chandrashekar et al. AIP Advances 7, 025310 (2017)

35 S. Talapatra, S. Kar, S. K. Pal, R. Vajtai, L. Ci, P. Victor, M. M. Shaijumon, S. Kaur, O. Nalamasu, and P. M. Ajayan, Direct

growth of aligned carbon nanotubes on bulk metals, Nature Nanotechnology 1, 112116 (2006).
36 J.L. Whitten and H. Yang, Theory of chemisorption and reactions on metal surfaces, Surface Science Reports 24, 5557
(1996).
37 R. G. Greenler, Infrared study of adsorbed molecules on metal surfaces by reflection techniques, The Journal of Chemical

Physics 44, 310315 (1966).

Das könnte Ihnen auch gefallen