You are on page 1of 10

Subscriber access provided by Kaohsiung Medical University

Article
Topochemical Deintercalation of Al from MoAlB: Stepwise Etching
Pathway, Layered Intergrowth Structures, and Two-Dimensional MBene
Lucas T. Alameda, Parivash Moradifar, Zachary Metzger, Nasim Alem, and Raymond E. Schaak
J. Am. Chem. Soc., Just Accepted Manuscript • DOI: 10.1021/jacs.8b04705 • Publication Date (Web): 15 Jun 2018
Downloaded from http://pubs.acs.org on June 16, 2018

Just Accepted

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

is published by the American Chemical Society. 1155 Sixteenth Street N.W.,


Washington, DC 20036
Published by American Chemical Society. Copyright © American Chemical Society.
However, no copyright claim is made to original U.S. Government works, or works
produced by employees of any Commonwealth realm Crown government in the course
of their duties.
Page 1 of 9 Journal of the American Chemical Society

1
2
3
4
5
6
7 Topochemical Deintercalation of Al from MoAlB: Stepwise Etch-
8
9
ing Pathway, Layered Intergrowth Structures, and Two-
10 Dimensional MBene
11
12 Lucas T. Alameda,‡,1 Parivash Moradifar,‡,2 Zachary P. Metzger,1 Nasim Alem,*,2,3 and Raymond E.
13 Schaak*,1,3
14
15 1
Department of Chemistry, 2 Department of Materials Science and Engineering, and 3 Materials Research Institute, The
16 Pennsylvania State University, University Park, PA 16802
17
18
19 ABSTRACT: The synthesis of refractory materials usually relies on high-temperature conditions to drive diffusion-limited solid-
20 state reactions. These reactions result in thermodynamically-stable products that are rarely amenable to low-temperature topochem-
21 ical transformations that post-synthetically modify subtle structural features. Here, we show that topochemical deintercalation of Al
22 from MoAlB single crystals, achieved by room-temperature reaction with NaOH, occurs in a stepwise manner to produce several
23 metastable Mo-Al-B intergrowth phases and a two-dimensional MoB (MBene) monolayer, which is a boride analog to graphene-
24 like MXene carbides and nitrides. A high-resolution microscopic investigation reveals that stacking faults form in MoAlB as Al is
deintercalated and that the stacking faulty density increases as more Al is removed. Within nanoscale regions containing high densi-
25
ties of stacking faults, four previously unreported Mo-Al-B (MAB) intergrowth phases were identified, including Mo2AlB2,
26 Mo3Al2B3, Mo4Al3B4, and Mo6Al5B6. One of these deintercalation products, Mo2AlB2, is identified as the likely MAB-phase pre-
27 cursor that is needed to achieve a high-yield synthesis of 2-D MoB, a highly targeted two-dimensional MBene. Microscopic evi-
28 dence of an isolated MoB monolayer is shown, demonstrating the feasibility of using room-temperature metastable phase engineer-
29 ing and deintercalation to access two-dimensional MBenes.
30
31
32
33 INTRODUCTION retention decreases the extent of atomic rearrangement needed
The rate-limiting step of most solid-state inorganic reactions to convert the reactants into products. It is therefore possible
34
is the diffusion of atoms. High temperatures are typically re- to selectively target more-reactive components of a structure at
35 lower temperatures while leaving intact less-reactive compo-
36 quired to move reactants through a solid-state matrix that is
comprised of microscale grains, which are characteristic of nents, which require higher temperatures to manipulate. Such
37 strategies are analogous to the different reactivity of various
bulk-scale syntheses. Nucleation and crystallization of target-
38 functional groups in the synthesis of organic molecules.
ed phases can only occur when all atoms are fully mixed at the
39 atomic level. However, the consequence of using heat to drive Topochemical deintercalation reactions are particularly in-
40 solid-state diffusion is that upon nucleation, there is a large teresting, because they can transform stable layered com-
41 excess of thermal energy relative to that which is needed to pounds that can be synthesized at high temperatures into new
42 crystallize most phases of a given composition. This limits the and otherwise inaccessible layered or non-layered analogs
43 scope of accessible products to phases that are thermodynami- under mild conditions that preserve key structural features. For
44 cally stable at high temperatures, as metastable and low- example, treatment of b-CaGe2 with HCl at -40 °C results in
45 temperature phases cannot withstand such conditions. the deintercalation of Ca2+ to form a CaCl2 byproduct while
46 Motivated by the possibility of discovering new or im- leaving behind crystals of a new layered GeH phase that re-
47 proved properties in phases that may otherwise be missed by tains the Ge layering motif present in b-CaGe2.4 Similarly,
48 high-temperature methods, several lower-temperature “soft dehydration of acidic layered Ruddlesden-Popper perovskite
chemistry” approaches have been developed.1 Such approach- oxides, H2[An-1BnO3n+1] (A = alkali, alkaline earth, or rare earth
49
es bypass diffusion barriers, thereby lowering reaction temper- metal, B = transition metal), below 500 °C results in deinterca-
50
atures and shifting the rate-limiting step from diffusion to nu- lation of interlayer protons and oxygen (as H2O) to form cubic,
51 A-site defective structural analogs.5 These and other examples
cleation.2 Using this strategy, many metastable and low-
52 temperature phases have been synthesized. Key principles of of topochemical deintercalation demonstrate how the primary
53 soft chemistry include achieving atomic-level mixing through layering motif of the material can be established using ther-
54 low-temperature processes, such as solution-phase dissolution modynamic control (i.e. high temperatures), and then subtle
55 and precipitation, or utilizing topochemical transformations changes to the structure can be achieved using kinetic control
56 where structural motifs of the reactant are preserved in the (low temperatures).
57 product after low-temperature manipulations.3 This structural
58
59
60 ACS Paragon Plus Environment
Journal of the American Chemical Society Page 2 of 9

MAX phases,6 which have the formula Mn+1AXn (M = early phases, including MoAlB (space group Cmcm),10 instead con-
1 transition metal, A = Al or Si, and X = carbon or nitrogen) and tain a zigzag double layer (Figure 1). These differences in both
2 space group P63/mmc, are refractory layered compounds that structure and reactivity underscore the need for an atomic-
3 undergo topochemical deintercalation of Al or Si upon reac- level understanding of the deintercalation process in MoAlB
4 tion with HF and related etchants under mild conditions.7,8 as a model MAB phase. Such understanding may help to iden-
5 Subsequent swelling of the deintercalated MAX phases leads tify and eventually overcome the roadblocks that currently
6 to exfoliation into 2-D nanosheets of MXenes, which are metal preclude the synthesis of MBenes from MAB phases. More
carbide and nitride analogues of graphene that are widely in- broadly, such insights may also help to facilitate the develop-
7
vestigated for applications in supercapacitors, batteries, and ment of topochemical deintercalation routes to metastable and
8 catalysis.9 Their boride analogues, referred to as MAB phases low-temperature MAB phases that would be inaccessible us-
9 and MBenes (M = transition metal; A = aluminum; B = bo- ing the high-temperature routes needed to synthesize them
10 ron), are emerging as potentially important alternatives to directly.
11 MAX phases and MXenes because additional compositional Accordingly, here we present a microscopic investigation
12 and structural variety offered by MAB phases may enable new into the topochemical deintercalation of Al from MoAlB sin-
13 or improved properties in their MBene derivatives relative to gle crystals upon treatment with either NaOH or a solution of
14 MXenes.10,11 Additionally, MBenes may be accessible using LiF in HCl – two commonly used etchants for MAX and
15 alternative, safer etchants than the fluoride-based etchants MAB phases – by using annular dark-field scanning transmis-
16 required for MXene synthesis.12 Unlike MXenes, however, sion electron microscopy (ADF-STEM). We find that deinter-
MBenes have yet to be synthesized. Extending the topochemi- calation of Al from MoAlB is accompanied by the formation
17
cal deintercalation chemistry that has been widely demonstrat- of a high density of (0k0) stacking faults, which helps to ra-
18 ed on MAX phases to MAB phases would open the door to a
19 tionalize the observed differences in etching periodicity in
new class of MXene-like 2D metal borides that may show MAX vs MAB phases and points to alternative approaches for
20 distinct or enhanced properties relative to MXenes. achieving exfoliation of MAB phases into MBenes. We ob-
21 serve that the metastable phase Mo2AlB2 (space goup
22 Cmmm)10 emerges during topochemical deintercalation of Al
23 from MoAlB. Mo2AlB2, which is structurally related to the
24 MAX phases that undergo complete Al deintercalation to form
25 MXenes, has not been experimentally realized in MAB phas-
26 es, but may be the necessary precursor for the synthesis of 2D
27 MBenes. Additionally, we observe that the high density of
28 stacking faults that are generated during the topochemical
29 deintercalation of Al from MoAlB leads to the formation of
several previously unidentified MAB intergrowth phases as
30
nanoscale grains embedded in MoAlB, including Mo2AlB2,
31 Mo3Al2B3, Mo4Al3B4, and Mo6Al5B6. Identification of these
32 MAB phases indicates that low-temperature soft chemistry
33 routes are useful in refractory boride systems for navigating
34 subtle structural differences that would not be tractable using
35 traditional high-temperature reactions. Also, the microscopic
36 insights into topochemical deintercalation of MoAlB indicate
37 that exfoliation of MAB phases to form MBenes, which has
38 remained synthetically elusive, may be feasible through routes
39 that differ from those used to generate MXenes.
Figure 1. Crystal structures of representative MAX phases (left,
40 Ti3AlC2 [100]), MAlB-type MAB phases (middle, MoAlB [001]),
41 and M2AlB2-type MAB phases (right, Fe2AlB2 [100]). EXPERIMENTAL SECTION
42 Synthesis. MoAlB single crystals were prepared using a
43 modification of a reported procedure.13 Briefly, Mo powder
44 Despite their lack of experimental realization, multiple stud- (Sigma Aldrich 99.99%, B powder (Alfa Aesar 98%), and Al
45 ies have predicted that MBenes should be accessible by topo- chips (NOAH Technologies 99.99%) were added to a 2-mL
46 chemical deintercalation of Al from MAB phases.10,11 Partial alumina crucible (Coorstek) in a Mo:Al:B molar ratio of
etching of Al from a prototype MAB phase, MoAlB, was 1:53.3:1. A total of 2 grams of reactants were added to the
47
achieved using NaOH as an etchant.12 In contrast to deinterca- crucible, which was then placed in a mullite tube furnace un-
48 lation of Al from MAX phases, which goes to completion and
49 der an argon flow. The tube was flushed with argon for 10
forms MXenes, Al deintercalates from MoAlB in layers minutes before the temperature was raised to 1400 °C. The
50 spaced every 50 – 300 nm, resulting in MoAlB slabs of na- sample was held at this temperature for 10 hours before cool-
51 noscale thickness. This partial, periodic Al removal observed ing to 900 °C at a rate of 1 °C/min, and then cooled to room
52 in MoAlB is not observed during the etching of MAX phases, temperature at a rate of 3 °C/min. The excess flux media was
53 where Al is instead removed from every layer of the structure. digested in 3M HCl at room temperature to yield MoAlB sin-
54 Interestingly, the Al substructures in the MAX and MAB gle crystals.
55 phases are different, which may contribute to their different
behavior during topochemical deintercalation. Whereas all Etching. MoAlB single crystals were etched with an aque-
56 ous solution of 10% NaOH or a solution of 2 M LiF in 6 M
MAX phases contain a planar single layer of Al, some MAB
57 HCl at room temperature for 2 hours and 24 hours. The
58
59
60 ACS Paragon Plus Environment
Page 3 of 9 Journal of the American Chemical Society

NaOH-etched crystals were then separated from the etchant high-contrast lines proceeding straight down from the crystal
1 with vacuum filtration and rinsed with deionized water. For surface. (Note that the contrast observed at stacking faults is a
2 the LiF/HCl samples, the solution was decanted and the vial result of stacking-fault-induced strain. Thus, changes in con-
3 was refilled with deionized water several times until ~ pH = 7. trast can be used to identify stacking fault distributions in low-
4 The crystals were then separated from the solution with vacu- magnification ADF-STEM images where atoms are not atomi-
5 um filtration and rinsed with deionized water. (Caution: An cally resolved.) At a stacking fault, the zigzag Al double layer
6 acidic solution of LiF will generate dilute hydrofluoric acid of MoAlB is replaced with a single layer of Al in a linear con-
(HF), and therefore should be treated with the same safety figuration, which is analogous to the interlayer Al in the
7
precautions used for HF.) M2AlB2 structure type. The difference between unfaulted and
8 faulted MoAlB is shown in Figures 2b and 2c, respectively.
9 Characterization. ADF-STEM images were collected on a
FEI Titan G2 aberration corrected high-resolution scan- The distance between each stacking fault ranges from tens to
10 hundreds of nanometers.
ning/transmission electron microscope at an accelerating volt-
11 age of 200 kV. Additional imaging details are included in the
12 Supporting Information. TEM sample preparation was done by
13 focused ion beam (FIB) using a FEI Helios NanoLab
14 660FIB/FESEM. Figure S2c shows a SEM image of the crys-
15 tal during the FIB process. A thin slice of the crystal was cut
16 from the bulk sample at the (100) crystal facet and perpendicu-
17 lar to the [001] crystallographic direction, and then lifted out
18 of the crystal. It was attached to a half copper TEM grid and
19 thinned until it was electron transparent. Scanning electron
microscopy (SEM) and energy-dispersive spectroscopy (SEM-
20
EDS) were performed on a FEI NOVA NanoSEM 630 at an
21 accelerating voltage of 5 kV for imaging and 20 kV for EDS,
22 and a working distance of 3 mm for imaging and 5 mm for
23 EDS. Samples were prepared for EDS by coating an SEM stub
24 with carbon tape (TED PELLA) upon which MoAlB single
25 crystals were placed with the basal plane perpendicular to the
26 SEM stub and the [100] orientation facing up. For imaging of
27 stacking-fault-band streaking, the crystals were mounted in
28 epoxy (Loctite) with the basal plane perpendicular to the SEM
29 stub and the [100] orientation facing up. X-ray diffraction was
performed by first grinding multiple crystals with an agate
30
mortar and pestle to prepare a powder of MoAlB. XRD pat-
31 terns were then collected on a Bruker D8 Advance powder X-
32 ray diffractometer.
33
34
35 RESULTS AND DISCUSSION
36 Topochemical Deintercalation of Aluminum. Topochemi-
37 cal deintercalation reactions, including for MAX and MAB
38 phases, are typically carried out using polycrystalline samples.
39 For this study, phase-pure MoAlB single crystals (Fig. S1)
were used so that we could control the extent of Al deinterca-
40
lation, focus on the surface of the crystal, orient the crystal so
41 that the crystallographic direction was known, and use FIB to
42 extract slices of known crystallographic orientation. MoAlB
43 single crystals were treated with 10% NaOH to etch Al (Fig.
44 2). The ADF-STEM images in Figure 2, focus on regions
45 within a few micrometers of the crystal surface. The images
46 were collected in the [001] crystallographic direction, which is
47 parallel to the (0k0) planes, therefore providing an edge-on
48 view of both the MoB layers and Al layers. The insets of Fig-
49 ures 2a and 2d show a schematic representation of the etching
process at different time points.
50
51 Figure 2a shows that prior to etching, (0k0) stacking faults
are sparsely distributed throughout the surface region of the
52
MoAlB crystal. The formation of sparsely distributed stacking
53
faults in MoAlB has been reported for flux growth synthesis
54 and high-temperature powder synthesis.14,15 Hultman and
55 coworkers hypothesized that stacking faults form as a result of Figure 2. ADF-STEM images of MoAlB at various stages of
56 stepwise intercalation of Al into MoB during the formation of etching with 10% NaOH. (a) As-synthesized MoAlB showing a
57 MoAlB.14 At low magnification, the stacking faults appear as few stacking faults as high-contrast lines vertically proceeding
58
59
60 ACS Paragon Plus Environment
Journal of the American Chemical Society Page 4 of 9

down from the crystal surface along the crystal lattice planes. The phase fraction of the four intergrowth structures was deter-
1 inset shows a schematic of the crystal and stacking faults. (b) mined by counting the number of times each phase appeared
2 Atomic-scale-resolution ADF-STEM image of pristine MoAlB. in 49 different imaged regions (85 total data points). Mo2AlB2
3 (c) Atomic-scale-resolution image of MoAlB showing a single was the most abundant (44%), followed by Mo4Al3B4 (35%),
stacking fault in the vertical direction indicated by an arrow below Mo6Al5B6 (14%), and Mo3Al2B3 (7%). Surrounding each or-
4
the image. (d) MoAlB after 2 hours of etching showing the for- dered nanoscale domain are regions containing random distri-
5 mation of several stacking fault bands. The inset shows a sche-
6 butions of stacking faults. Lattice defects have been observed
matic of the crystal and stacking fault bands. (e) Higher- previously in MAB phases, including (0k0) stacking faults in
7 magnification image of the sample with a high density of stacking
MoAlB,15 twist boundaries in Cr2AlB2 due to 90° rotation
8 faults. (f) Higher-magnification image of two stacking fault bands
about the b-axis, and [001] tilt boundaries in Fe2AlB2.14 How-
9 separated by a region of mostly unperturbed MoAlB with no
stacking faults. ever, to our knowledge, there have not been reports of defect
10 ordering or defect engineering, yielding regions of other or-
11 dered phases, in MAB compounds.
12 After 2 hours of treatment with 10% NaOH, regions of high Stacking faults are also well known in MAX phases.16-18
13 stacking fault density (“stacking fault bands”) appear (Fig. They are caused by the insertion of extra MX layers between
14 2d). These regions are ~100 nm wide and separated from each the A layers. For example, Wang and coworkers showed that
15 other by < 100 nm of mostly unperturbed MoAlB. In the de- TaC slabs insert into Ta2AlC to yield Ta4AlC3 and Ta6AlC5.16
16 fect-rich regions, stacking faults occur within nanometers of The insertion of a TaC slab caused the number of Ta layers in
17 each other, such that the frequency of stacking faults (“Al the carbide substructure to increase from 2 to 4 for Ta4AlC3 or
18 single layers”) and normal stacking sites (“Al double layers”) 2 to 6 for Ta6AlC5. These intergrowth structures can be de-
is comparable (Fig 2e). In the regions of unperturbed MoAlB, scribed by the formula M2nAX2n-1, where n represents the num-
19
there are no observable high-contrast lines that are attributed ber of TaC layers in each carbide substructure and is equal to
20 to stacking faults (Fig 2f). An SEM image of the crystal edges 1, 2, and 3 for Ta2AlC, Ta4AlC3, and Ta6AlC5, respectively.
21 after 2 hours of treatment shows that microscale streaking The M2nAX2n-1 tantalum aluminum carbides can be roughly
22 occurs that is in the same plane as the stacking fault bands, considered as MAX-phase structural analogs to Ruddlesden-
23 and is also consistent with the size and periodicity of the Popper layered perovskites, (A2[A’n-1BnO3n+1]), which are peri-
24 bands, suggesting that the bands are observable by SEM (Fig- odic intergrowths of rocksalt AO layers and perovskite ABO3
25 ure S2 of the Supporting Information). The streaks do not exist layers. The Al substructure in M2nAX2n-1 remains unchanged
26 prior to etching and they are present over the entire crystal across each phase and is analogous to the alkaline earth metal
27 edge (which is on the order of millimeters), confirming that (A) in Ruddlesden-Popper phases. In contrast, the stacking
28 the stacking fault bands observed by ADF-STEM occur eve- behavior observed in etched MoAlB is fundamentally different
rywhere that the crystal edge was exposed to NaOH. from that of Ruddlesden-Popper and Ta-Al-C phases. The
29
Collectively, these images demonstrate that upon initial MoB substructure remains unchanged across all phases. In-
30
NaOH treatment, Al partially deintercalates from MoAlB in a stead, the Al substructure changes between single-layer or
31 double-layer states. Thus, the stacking variants that form dur-
manner that was unexpected. As atoms are removed from an
32 ing the topochemical deintercalation of Al from MoAlB repre-
Al double layer, the remaining Al atoms redistribute evenly
33 between the two adjacent MoB layers. Thus, instead of form- sent unique structural motifs for ternary layered borides and
34 ing vacant channels between MoB layers upon Al deintercala- carbides. To understand the structural relationships between
35 tion, as is observed upon Al deintercalation of MAX phases to the stacking variants, the layering can be understood in terms
36 form MXenes, the remaining atoms spread out to form an Al of the sequence of Al layers rather than the sequence of MoB
37 single layer that holds together the boride layers. The stacking layers. Accordingly, each Al substructure in Figure 3 is la-
38 faults assemble as bands, similar to the stacking fault behavior beled with the number of Al layers present, where 1 represents
39 observed in many other materials, including MAX phases. a linear single layer and 2 represents a zigzag double layer.
Each intergrowth phase then contains a unique stacking se-
40
quence of single-layer (1) and double-layer (2) motifs: [1]n,
41 Ordered Stacking Variants Induced by Topochemical De- [211]n, [12]n, and [122]n for Mo2AlB2, Mo3Al2B3, Mo4Al3B4,
42 intercalation. With further NaOH etching beyond 2 hours, the and Mo6Al5B6, respectively. For comparison, MoAlB, which
43 stacking fault bands become continuous as Al deintercalates contains only double layers, would be denoted as [2]n
44 from the unperturbed areas of MoAlB and the stacking fault
Figure 3 presents potential deintercalation pathways to the
45 density concomitantly increases. At this point, stacking faults
four different intergrowth phases, starting with MoAlB as an
46 appear to order into regions containing nanoscale domains of
entryway. Considering only the molecular formulas, it is rea-
47 various other MoB-Al stacking motifs (Figure 3). Four unique
sonable to guess that each phase could be accessed by sequen-
stacking variants emerged, which ranged in composition from
48 tially decreasing the Al content, starting with the conversion of
Mo6Al5B6 to Mo2AlB2 and follow the stoichiometry MoxAlx-
49 MoAlB to Mo6Al5B6, followed by the formation of Mo4Al3B4,
1Bx. With the exception of Mo2AlB2, for which the unit cell
50 then Mo3Al2B3, and finally, Mo2AlB2. However, careful in-
and lattice parameters have already been determined using
51 spection of the stacking sequences shows that two of the con-
density functional theory calculations,10 the proposed chemical
52 versions, Mo6Al5B6 à Mo4Al3B4 and Mo4Al3B4 àMo3Al2B3,
formula for each intergrowth structure was determined by a
are not possible. In these two cases, there is no way to convert
53 simple analysis based on the structures observed by ADF-
an Al double layer to a single layer and obtain the product
54 STEM imaging and the change in the number of Al layers
phase. We assume Al dissolution is irreversible under the con-
55 relative to MoB layers. These ordered regions were observed
ditions of the Al deintercalation reaction, and therefore Al-
56 repeatedly in multiple regions of the crystal, and were also
deficient phases cannot act as intermediates to Al-rich phases.
57 observed when LiF/HCl was used as the etchant. A rough
(Al would have to be inserted, which is not consistent with the
58
59
60 ACS Paragon Plus Environment
Page 5 of 9 Journal of the American Chemical Society

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
Figure 3. ADF-STEM images of stacking variants and their corresponding crystal structures observed after etching MoAlB with 10%
32 NaOH or LiF/HCl for 24 hours. Arrows show possible reaction pathways between phases assuming irreversible dissolution of Al. ‘1’ rep-
33 resents a linear Al single layer, and ‘2’ represents a zigzag Al double layer. Note that the Al atoms in Mo2AlB2 and Mo3Al2B3 are not visi-
34 ble due to a higher level of induced microstructural strain that results in more lattice distortion. Scale bars are 1 nm.
35
36
37 reaction conditions that favor Al removal.) This prohibits the ty, oxidation resistance, and the various mechanical proper-
formation of Mo4Al3B4 from Mo2AlB2, suggesting that ties that are being extensively studied in MAB phases.19-21
38
Mo4Al3B4 must form either directly from MoAlB, or from an
39 additional phase that has not yet been observed. It should be
40 Formation of Etched Cavities. After further treatment of
emphasized that these are proposed pathways based on ex-
41 MoAlB with 10% NaOH for 24 hours (Fig. 4), etched cavi-
situ analysis, and may not be a true representation of the
42 ties form. The cavities are ~10 nm wide at the crystal surface
actual reaction mechanism.
and are distributed periodically, occurring once every 100 –
43 In addition to the structural differences between MAX- 200 nm. This observation is consistent with previous SEM
44 phase stacking faults and MoAlB stacking faults, there are images showing similar periodic etching features.12 The
45 important synthetic differences between the two. The high cross-sectional view of etched MoAlB reveals that the cavi-
46 density of MoAlB stacking faults reported in this study are ties have a range of depths spanning from less than 100 nm,
47 due to the post-synthetic, intentional removal of Al using as is shown in Figure 4a, to over several micrometers. Nota-
48 “soft chemistry,” rather than unintentional insertion of MoB bly, a very high density of stacking faults forms after etching
49 during high temperature synthesis. Thus, there is potential for 24 hours. In the region near the crystal surface, there are
50 for the development of kinetically-controlled topochemical no remaining areas of unperturbed MoAlB; ADF-STEM
deintercalation reactions to selectively target different stack- images from all areas of the crystal surface exhibit signifi-
51
ing variants in the Mo-Al-B system and isolate phase pure cantly increased contrast due to strain from extensive stack-
52 bulk samples, particularly if applied to polycrystalline sam- ing faults. Careful inspection of atomic-resolution images
53 ples having smaller grain sizes and diffusion distances. Such reveals that the density of stacking faults becomes greater
54 capabilities in refractory borides are not well established. We than the density of Al double layers after the 24 hour treat-
55 anticipate that low-temperature structural engineering will ment (Fig. 4c). Of the 15 Al layers shown in Figure 4c, 9 are
56 result in modulation of the properties that are influenced by faulted single layers and 6 are unperturbed double layers.
57 layering motifs, including electrical and thermal conductivi- Similar ratios were observed consistently over large regions
58
59
60 ACS Paragon Plus Environment
Journal of the American Chemical Society Page 6 of 9

of the sample. The stacking faults extend a few micrometers described as a reaction intermediate. Further evidence of the
1 below the surface of the crystal before terminating, at which stacking fault dependency on cavity formation is evident in
2 point unperturbed MoAlB is again observed. SEM-EDS Figure 4b, which shows the point where a cavity terminates.
3 maps of the crystal edge show that the chemical homogenei- In this image, the tip of the cavity appears to preferentially
4 ty of Mo and Al is preserved through the etching process track a stacking fault over an adjacent Al double layer. It
5 (Figure S3 of the Supporting Information). should be noted, though, that due to the high level of strain
6 present at the tips of cavities, only a few sites could be atom-
ically resolved. Thus, we cannot yet generalize this behavior
7
to all cavities. Alternatively, etched cavities may form due to
8 strain caused by contraction of the crystal upon Al deinterca-
9 lation. Strain from the newly formed etched cavities may
10 then contribute to additional stacking fault density in the
11 neighboring regions.
12
13
14
15
16
17
18
19
20
21
22
23
24 Figure 5. (a) ADF-STEM image of MoAlB after 24 hours of
25 etching with a solution of LiF in HCl showing etched cavities,
26 Figure 4. (a) ADF-STEM image of MoAlB after 24 hours of etched “branches,” and high density of stacking faults. As this
27 etching with 10% NaOH showing large etched cavities and image indicates, etching preferentially occurs along [100] and
28 stacking fault saturation. (b) ADF-STEM image of the tip of an [010] crystallographic orientations.
etched cavity showing extensive strain and deformation of the
29
atomic rows at the crack tip, which terminates at a stacking
30 fault. (c) High resolution ADF-STEM image of a representative From a rough contrast analysis of the images presented in
31 surface region indicating a high density of stacking faults. Figures 2a, 2d, and 4a, we propose the following reaction
32 pathway for partial etching of Al from MoAlB (Fig. 6). (1)
33 Al atoms topochemically deintercalate to convert Al double
34 Nearly identical behavior to the NaOH treatment was ob- layers into stacking faults containing Al single layers, which
35 served when the crystals were etched with a solution of 2M occurs in bands that extend a few micrometers into the bulk
LiF in 6M HCl (Figure S4 of the Supporting Information), of the crystal. (2) The stacking fault bands eventually reach a
36
which is a commonly used Al etchant for MAX phases. Af- point of saturation, where the entire surface region is uni-
37 ter 2 hours of treatment with LiF/HCl, there was moderate formly faulted with a defect density greater than 50%. At
38 stacking fault formation, and after 24 hours of treatment a this point, the stoichiometry of the surface region is closer to
39 high density of stacking faults formed along with etched that of Mo2AlB2, a phase that cannot be made by direct high-
40 cavities. The one significant difference between the NaOH temperature synthesis, than it is to MoAlB. (3) Al single
41 and LiF/HCl etchants was that LiF/HCl caused extensive layers are then removed from stacking faults, completely
42 corrosion of the boride layer in addition to Al deintercala- removing the Al between two adjacent MoB layers and re-
43 tion. With LiF/HCl, etched “branches” were observed to sulting in an etched cavity that then swells to ~10 nm wide.
44 propagate in [010], corroding through MoB layers to form a Alternatively, step 3 may occur in reverse order where cavi-
45 network of cavities (Fig. 5). This behavior suggests that the ties form due to strain, which is then followed by dissolution
MoB layer slowly dissolves in the presence of HF at room of the Al atoms that line the inside of the cavity.
46
temperature and therefore that LiF/HCl is not suitable for Al
47
deintercalation from MoAlB, as competing corrosion path-
48 ways dominate. Consistent with this microscopic observa- Implications for the Synthesis of MBenes. Previous at-
49 tion, hollowing was reported previously for HF-treated tempts to exfoliate MoAlB into 2D MoB, a MBene, instead
50 MoAlB crystals.12 resulted in the periodic formation of etched cavities, which
51 was attributed to periodic etching of Al. This microscopic
The sequential formation of stacking faults followed by study now reveals that Al is actually partially etched from
52 etched cavities suggests that the cavities may be dependent the majority of the Al double layers of MoAlB. In most cas-
53 on stacking faults. This relationship is reasonable if one as- es, only one layer of Al is removed from the double layer,
54 sumes that cavities only form when both Al layers between resulting in the formation of a stacking fault having a single
55 two adjacent MoB layers are removed. In that case, a possi- Al layer rather than an etched cavity. Fully exfoliating a
56 ble pathway for Al removal could be a stepwise deintercala- MAB phase to form a MBene requires full etching of Al
57 tion of the two Al layers, where a stacking fault would be between the MoB layers, and therefore pushing the topo-
58
59
60 ACS Paragon Plus Environment
Page 7 of 9 Journal of the American Chemical Society

1
2
3
4
5
6
7
8
9
10 Figure 6. Combined reaction scheme representing the topochemical deintercalation of Al from MoAlB upon treatment with 10% NaOH.
11
12
13 H2 would form, analogous to the process that etches Al from
14 MAX phases using HF. Several MAB-phase structure types
15 are known to form as phase-pure compounds, including
16 M4AlB6, M3AlB4, M2AlB2, and MAlB, where M = transition
metal.22 However, each structure type can only accommo-
17
date certain metals. To our knowledge, there were no exper-
18
imental reports of the Mo-containing Mo2AlB2 phase on
19 which the DFT study was based. However, the results of our
20 microscopic study suggest that Mo2AlB2 may indeed be syn-
21 thetically accessible using low-temperature soft chemistry
22 pathways. Preliminary experimental evidence supporting
23 Sun and coworkers’ conclusion that 2D MBenes are accessi-
24 ble, stable, and can be delaminated into single sheets is
25 shown in Figure 7. In this image, two MoB monolayers split
26 apart as the stacking fault holding them together is etched.
27 The monolayers are partially delaminated from the inner
wall of an etched cavity, and the longer of the two (left side)
28
extends approximately 10 nm before leaving the field of
29 view. Although it has not been observed or achieved on a
30 large scale, this result closes the loop between the micro-
31 Figure 7. Left: ADF-STEM image of two isolated, delaminated scopic insights and experimental implications and suggests
32 MoB (MBene) sheets inside an etched cavity. Middle: Higher- that the formation of stable 2D MoB is feasible using the
33 magnification and contrast-enhanced image of the cavity con- knowledge gained from this study.
taining the MBene sheets, as indicated by the blue outline.
34
Right: Proposed, idealized structure of the delaminated region of
35 the MBene sheets. CONCLUSIONS
36
In conclusion, ADF-STEM imaging revealed important
37
chemical deintercalation of Al to completion across the en- insights into the pathway by which Al is deintercalated from
38 MoAlB, a prototype MAB phase that is of interest as a pre-
39 tire crystal surface. While it is not yet understood why com-
plete etching of an Al double layer doesn’t occur across the cursor for the formation of 2D MBenes. Deintercalation of
40 Al occurs in a stepwise manner, involving the formation of
entire crystal surface, one possible explanation is that the
41 stacking fault bands that grow over time until a high density
formation of certain stacking sequences, i.e. some of the
42 of stacking faults is reached. Aluminum then is removed
intermediate stacking variants described above, causes dein-
43 tercalation to cease. From this microscopic study, we now from stacking faults to form etched cavities. Considering that
44 know that one of the stacking variants formed upon partial cavity formation is preceded by the formation of a high den-
45 Al deintercalation of MoAlB adopts the Mo2AlB2 structure. sity of stacking faults, where the composition of the crystal
46 and the local structure more closely resemble Mo2AlB2 than
Considering that MAX phases are structurally more similar
MoAlB, Mo2AlB2 is proposed as a viable MAB-phase pre-
47 to Mo2AlB2 (single-layer Al) than MoAlB (double-layer Al),
we propose that the synthesis of 2D MoB (MBene) requires cursor that will enable the high-yield synthesis of 2D MoB
48 (MBene). Preliminary progress towards the synthesis of
49 Mo2AlB2 as the starting material, because the intermediate
Mo2AlB2, as well as the other intergrowth phases Mo3Al2B3,
50 stacking variants described above, which may cause deinter-
calation to cease, cannot form from Mo2AlB2. The syntheti- Mo4Al3B4, and Mo6Al5B6, was demonstrated through the
51 observation of regions where ordering of stacking faults oc-
cally-accessible MoAlB phase is therefore unlikely to pro-
52 curred upon topochemical deintercalation of Al in MoAlB.
duce MBenes in high yield, which motivates the develop-
53 ment of synthetic routes to phase-pure Mo2AlB2.
Additionally, an isolated 2D MoB monolayer was observed,
54 suggesting that the synthesis and exfoliation of MBenes is
Consistent with this hypothesis, Sun and coworkers re- indeed feasible. The observation that diverse stacking vari-
55
cently reported DFT studies predicting that 2D molybdenum ants and layered intergrowth compounds form through “soft
56 boride is indeed accessible via HF treatment of Mo2AlB2.10 chemistry” pathways has important implications for achiev-
57 Upon HF treatment, the DFT study predicted that AlF3 and
58
59
60 ACS Paragon Plus Environment
Journal of the American Chemical Society Page 8 of 9

ing synthetic control over subtle structural features in refrac- (6) Barsoum, M. W.; El-Raghy, T. The MAX Phases: Unique
1 tory boride systems, which typically require high- New Carbide and Nitride Materials. Am. Sci. 2001, 89, 334-343.
2 temperature reactions and therefore are limited to phases that (7) Naguib, M.; Kurtoglu, M.; Presser, V.; Lu, J.; Niu, J.; Heon,
M.; Hultman, L.; Gogotsi, Y.; Barsoum, M. W. Two-Dimensional
3 are stable at high temperatures. Optimization, expansion, and
Nanocrystals Produced by Exfoliation of Ti3AlC2. Adv. Mater. 2011,
4 generalization of this topochemical deintercalation approach 23, 4248-4253.
5 beyond MAB-phase compounds would provide a powerful (8) Alhabeb, M.; Maleski, K.; Mathis, T. S.; Sarycheva, A.; Hat-
6 strategy for transcending thermodynamic limitations to the ter, C. B.; Uzun, S.; Levitt, A.; Gogotsi, Y. Selective Etching of
phase accessibility of refractory materials. Silicon from Ti3SiC2 (MAX) to Obtain 2D Titanium Carbide
7
(MXene). Angew. Chem. Int. Ed. 2018, 57, 5444-5448.
8 (9) Anasori, B.; Lukatskaya, M. R.; Gogotsi, Y. 2D Metal Car-
9 bides and Nitrides (MXenes) for Energy Storage. Nat. Rev. Mater.
ASSOCIATED CONTENT
10 2017, 2, 16098.
11 Supporting Information (10) Guo, Z.; Zhou, J.; Sun, Z. New Two-Dimensional Transition
Metal Borides for Li Ion Batteries and Electrocatalysis. J. Mater.
12 XRD data, additional electron microscopy images, and details of Chem. A. 2017, 5, 23530-23535.
13 imaging conditions. (11) Jiang, Z.; Wang, P.; Jiang, X.; Zhao, J. MBene (MnB): a
14 New Type of 2D Metallic Ferromagnet with High Curie Tempera-
The Supporting Information is available free of charge on the ture. Nanoscale Horiz. 2018, 3, 335-341.
15
ACS Publications website. (12) Alameda, L. T.; Holder, C. F.; Fention, J. L.; Schaak, R. E.
16 Partial Etching of Al from MoAlB Single Crystals to Expose Cata-
17 PDF file lytically Active Basal Planes for the Hydrogen Evolution Reaction.
18 Chem. Mater. 2017, 29, 8953-8957.
19 (13) Okada, S.; Atoda, T.; Higashi, I.; Takahashi, Y. Preparation
20 of Single Crystals of MoB2 by the Aluminum-Flux Technique and
AUTHOR INFORMATION some of their Properties. J. Mater. Sci. 1987, 22, 2993-2999.
21 (14) Lu, J.; Kota, S.; Barsoum, M. W.; Hultman, L. Atomic
22 Corresponding Author Structure and Lattice Defects in Nanolaminated Ternary Transition
23 * res20@psu.edu Metal Borides. Mater. Res. Lett. 2017, 5, 235-241.
24 * nalem@psu.edu (15) Kota, S.; Zapata-Solvas, E.; Ly, A.; Lu, J.; Elkassabany, O.;
Huon, A.; Lee, W. E.; Hultman, L.; May, S. J.; Barsoum, M. W.
25 Synthesis and Characterization of an Alumina Forming Nano-
26 Author Contributions laminated Boride: MoAlB. Sci. Rep. 2016, 6, 26475.
27 ‡ These authors contributed equally. (16) Lin, Z.; Zhuo, M.; Zhou, Y.; Li, M.; Wang, J. Microstruc-
28 tures and Theoretical Bulk Modules of Layered Ternary Tantalum
29 ACKNOWLEDGMENT Aluminum Carbides. J. Am. Ceram. Soc. 2006, 89, 3765-3769.
(17) Joulain, A.; Thilly, L.; Rabier, J. Revisiting the Defect
30 L.T.A., Z.P.M., and R.E.S. were supported by the U.S. National
Structure of MAX Phases: the Case of Ti4AlN3. Philos. Mag. 2008,
31 Science Foundation (NSF) Center for Chemical Innovation in 88, 1307-1320.
Solar Fuels (CHE-1305124). P.M. and N.A. were supported by
32 (18) Yu, R.; Zhan, Q.; He, L. L.; Zhou, Y. C.; Ye, H. Q. Stacking
the Penn State Center for Nanoscale Science, a NSF-sponsored Faults and Grain Boundaries of Ti3SiC2. Philos. Mag. Lett. 2003, 83,
33 Materials Science and Engineering Center under award number 325-331.
34 NSF-DMR 1420620. The authors thank Tom Mallouk for help- (19) Kota, S.; Zapata-Solvas, E.; Chen, Y.; Radovic, M.; Lee, W.
35 ful discussions and Ke Wang for electron microscopy support. E.; Barsoum, M. W. Isothermal and Cyclic Oxidation of MoAlB in
36 Air form 1100 °C to 1400 °C. J. Electrochem. Soc. 2017, 164, 930-
37 938.
38 REFERENCES (20) Ali, M. A.; Hadi, M. A.; Hossain, M. M.; Naqib, S.
(1) Stein, A.; Keller, S. W.; Mallouk, T. E. Turning Down the H.; Islam, A. K. M. A. Theoretical Investigation of Structural, Elas-
39 tic, and Electronic Properties of Ternary Boride MoAlB. Phys. Sta-
Heat: Design and Mechanism in Solid-State Synthesis. Science
40 1993, 259, 1558-1564. tus Solidi B 2017, 254, 1700010.
41 (2) Martinolich, A. J.; Kurzman, J. A.; Neilson, J. R. Circum- (21) Li, X.; Cui, H.; Zhang, R. First-Principles Study of the Elec-
42 venting Diffusion in Kinetically Controlled Solid-State Metathesis tronic and Optical Properties of a new Metallic MoAlB. Sci.
Reactions. J. Am. Chem. Soc 2016, 138, 11031-11037. Rep. 2016, 6, 39790.
43 (22) Ade, M.; Hillbrecht, H. Ternary Borides Cr2AlB2, Cr3AlB2,
(3) Gopalakrishnan, J. Chimie Douce Approaches to the Synthe-
44 Cr3AlB4, and Cr4AlB6: The First Members of the Series
sis of Metastable Oxide Materials. Chem. Mater. 1995, 7, 1265-
45 1275. (CrB2)nCrAl with n = 1, 2, 3 and a Unifying Concept for ternary
46 (4) Bianco, E.; Butler, S.; Jiang, S.; Restrepo, O. D.; Windl, W. Borides as MAB-Phases. Inorg. Chem. 2015, 54, 6122-6135.
47 Goldberger, J. E. Stability and Exfoliation of Germanane: A Ger-
48 manium Graphane Analogue. ACS Nano 2013, 7, 4414-4421.
(5) Schaak, R. E.; Mallouk, T. E. Synthesis, Proton Exchange,
49 and Topochemical Dehydration of New Ruddlesden-Popper Tanta-
50 lates and Titanotantalates. J. Solid State Chem. 2000, 155, 46-54.
51
52
53
54
55
56
57
58
59
60 ACS Paragon Plus Environment
Page 9 of 9 Journal of the American Chemical Society

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

You might also like