You are on page 1of 23

Subscriber access provided by CHINESE CULTURE UNIV

Article
Electronic Structure of Core/shell Metal/oxide Aluminium Nanoparticles
Giulia Maidecchi, Chinh Duc Vu, Renato Buzio, Andrea Gerbi,
Gianluca Gemme, Maurizio Canepa, and Francesco Bisio
J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.5b07678 • Publication Date (Web): 04 Nov 2015
Downloaded from http://pubs.acs.org on November 13, 2015

Just Accepted

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

The Journal of Physical Chemistry C 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 22 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8 Electronic Structure of Core/Shell Metal/Oxide
9
10
11
12 Aluminium Nanoparticles
13
14
15
16 Giulia Maidecchi,†,⊥ Chinh Vu Duc,‡,† Renato Buzio,¶ Andrea Gerbi,¶ Gianluca
17
18 Gemme,§ Maurizio Canepa,k,§ and Francesco Bisio∗,¶
19
20
21 †Dipartimento di Fisica, Università di Genova, via Dodecaneso 33, I-16146 Genova, Italy
22
23 ‡Institute of Materials Science - Vietnam Academy of Science and Technology, 18 Hoang
24
25 Quoc Viet Road, Cau Giay District, Ha Noi, Viet Nam
26
27 ¶CNR-SPIN, C.so Perrone 24, I-16152 Genova, Italy
28
29
§INFN, Sezione di Genova, via Dodecaneso 33, I-16146 Genova, Italy
30
31
32
kDipartimento di Fisica, Università di Genova and CNISM, Sede Consorziata di Genova,
33
34
via Dodecaneso 33, I-16146 Genova, Italy
35
36 ⊥Present address: Istituto Italiano di Tecnologia, Via Morego 30, 16163 Genova, Italy
37
38
39 E-mail: francesco.bisio@spin.cnr.it
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
1
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 2 of 22

1
2
3
4
Abstract
5
6 We report a combined X-ray photoelectron spectroscopy and atomic-force microscopy
7
8 study of ultrafine (< 30 nm) Al/Al-oxide core/shell nanoparticles. A complex fine
9
10 structure is observed within O1s and Al2p core-level electronic spectra of the particles,
11
12 indicative of a corresponding complex system morphology. The photoemission spectra
13
14 of the Al metallic core exhibited a low-binding-energy component, ascribed to strongly
15
16 undercoordinated metallic Al atoms in the particle core. Such a fraction of underco-
17
18 ordinated metallic atoms grows larger in relative weight as the particles get smaller,
19
finally becoming completely dominant for nanoparticle size below 20 nm. This fea-
20
21 ture is interpreted as the fingerprint of vacancy-cluster formation within the metallic
22
23 core as an effect of oxidation. For the smallest particle size investigated (< 10 nm),
24
25 vacancy clusters coalesce leading to the core disruption, with a heavy impact on the
26
27 corresponding electronic and plasmonic response.
28
29
30
31
32 1 Introduction
33
34
35 Aluminum nanoparticles (NPs) and nanostructures are extremely appealing due to their po-
36
37 tential application as low-cost plasmonic systems endowed with broadband plasmonic func-
38
39 tionality extending from the visible to the deep-ultraviolet (DUV) photon-energy range. 1–26
40
41 The plasmonic response of aluminium NPs is extremely sensitive to the NP size and shape, 2,3,6,8,17
42
43 much more than the conventional plasmonic metals Au and Ag. This implies that the control
44
45 over the NP morphology and composition is of paramount importance, the more so when
46
47 DUV plasmonics, calling for NP size around or below the 20-nm mark, is involved.
48
49 Aluminum undergoes a fast surface oxidation in atmosphere, leading to the quick for-
50
51 mation of a 2-3 nm thick oxide layer. 27 Such a native oxide has a strong influence on the
52
53 plasmonic response, leading to a significant redshift of the plasmon resonances, that can be
54
55 also positively exploited to extend the application range of Al NPs towards the visible-light
56
57
regime. 1–3,6–9,11 The impact of oxidation clearly becomes increasingly relevant for smaller NP
58
59
60
2
ACS Paragon Plus Environment
Page 3 of 22 The Journal of Physical Chemistry

1
2
3
4
size, where the mass transport associated with the anion/cation migration and their different
5
6 migration velocity may strongly affect the morphology of the metallic core. 27,28 Understand-
7
8 ing the morphology and composition of ultrafine metal/oxide Al core-shell NPs is therefore
9
10 relevant for harnessing the fabrication procedures of Al nanostructures, and understanding
11
12 their chemical and physical response for all the applications in atmosphere. Despite the
13
14 relevance of the subject and the growing interest in Al NPs, however, the objective difficulty
15
16 of fabricating very fine Al particles and of addressing this system have strongly limited the
17
18 studies of Al NPs morphology and composition in the <20 nm size regime. 28
19
20 In this paper, we report a combined X-ray photoelectron spectroscopy (XPS) and atomic-
21
22 force microscopy (AFM) study of ordered arrays of metal/oxide Al core-shell NPs with
23
24 diameter in the 10 − 30 nm range. XPS is indeed extremely powerful for exploring the
25
26 growth of Al oxide and the resulting chemico-physical properties of oxidised Al, 29–39 and can
27
28 be efficiently applied to characterize core-shell NPs. The XPS data are interpreted by means
29
30 of a simple model relying on an ancillary atomic force microscopy (AFM) characterization
31
32 that gives insights in the morphology of such systems. 8,40 The deconvolution of XPS spectra
33
34 demonstrates the existence of O and Al atoms in different ionization states in the NPs. In
35
36 particular, we observed a characteristic low-binding-energy (BE) metallic-Al subcomponent
37
38 that gradually became completely dominant for the smallest NP size investigated. Such
39
40
a feature was interpreted as the spectroscopic fingerprint of vacancy-cluster formation in
41
42
the metallic NP core during the oxidation process, suggesting a major modification of the
43
44
45
morphology and electronic structure of the metallic core below a critical NP size with respect
46
47 to the ideal ”compact-core” situation, with a strong impact on Al-NP electronic and optical
48
49 response.
50
51
52
53
54
55
56
57
58
59
60
3
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 4 of 22

1
2
3
4
2 Results and discussion
5
6
7 2.1 Experimental
8
9
10 The systems consist of ultrafine Al NPs, arranged in 2-dimensional (2D) densely-packed
11
12 arrays supported on an insulating single-crystal substrate. The arrays are fabricated by
13
14 a template-deposition approach, depositing Al on LiF(110) substrates previously nanopat-
15
16 terned by self-organization. The samples were fabricated in a high-vacuum vessel (chamber
17
18 base pressure pbase = 5 · 10−9 mbar) according to a procedure which is described in detail
19
20 in previous work. 8,41 First, self-organized nanopatterns are realized on the surface of a flat
21
22 LiF(110) crystal by homoepitaxial deposition (LiF(110) crystals, 10 × 10 × 1 mm3 , Crystec
23
24 Gmbh). The nanopatterns consist of aligned nanoridges formed by evenly-spaced [100] and
25
26 [010] exposed facets with a periodicity Λ ≈ 25 nm. 41–43 Thereafter aluminium is deposited
27
28 at grazing incidence (60◦ from the normal) on the nanopatterned LiF(110) by molecular
29
30 beam epitaxy at a pressure p < 2.5 · 10−8 mbar, thus preventing the inclusion of impu-
31
32 rities within the metallic core. 44 The grazing-angle deposition promotes the formation of
33
34 Al nanowires, thanks to the shadow effect of the nanoridges. An annealing treatment at
35
36
T=670 K in vacuum for five minutes promotes the nanowire dewetting and consequent for-
37
38
39
mation of disconnected Al NPs. 8,41 After the annealing, the samples are exposed to a flux
40
41
of research-grade oxygen (p = 1 · 10−5 mbar × 200 s) at a temperature T≈ 350 K.
42
43 The NPs are then briefly exposed to atmosphere (typ. few minutes) prior to their inser-
44
45 tion in the XPS apparatus, again under HV conditions. The AFM characterization, involving
46
47 longer permanence in atmosphere, which may affect the oxide composition, was performed
48
49 on sacrificial samples, nominally identical to the ones used for XPS. In this work, we ad-
50
51 dressed Al NPs characterized by a different mean size, that were obtained simply depositing
52
53 a different amount of Al onto the nanopatterned substrate. 8 In order to minimize any spuri-
54
55 ous substrate effect, NPs with different mean size were realized depositing different amounts
56
57 of Al on different zones of the same nanopatterned substrate. These zones are thus labelled,
58
59
60
4
ACS Paragon Plus Environment
Page 5 of 22 The Journal of Physical Chemistry

1
2
3
4
in order of increasing Al coverage, as ”zone” 0, 1, 2 and 3 (z0, z1, z2, z3), corresponding
5
6 to an Al effective coverage (defined as the thickness of the deposited material assuming a
7
8 uniform substrate coverage) of 0.8 nm, 1.7 nm, 2.5 nm and 3.4 nm, respectively.
9
10 The sample characterization has been performed within 24 hours from their fabrication.
11
12 AFM imaging was realized by meas of a Solver P47H-PRO by NT-MDT. The XPS spectra
13
14 were collected using a PHI ESCA 5600 system, equipped with a monochromatized Al source
15
16 Kα (1486.7 eV photon energy). Due to the insulating properties of the LiF substrates, it was
17
18 necessary to use a neutralizer electron gun to compensate for the sample charging effects.
19
20 The neutralizer was applied extremely carefully in order to avoid the deformation of the
21
22 peaks due to the inhomogeneous charge-up of the substrates.
23
24
25
26 2.2 Morphology of Al nanoparticle arrays
27
28
29 In Fig. 1a-d we report representative AFM images for each of the four different zones (z0
30
31 to z3). A graphical sketch of the array morphology is reported in the figure center. The
32
33 nanopatterned LiF surface that supports the NPs exhibits coherently-aligned nanogrooves,
34
35 that effectively induce the formation of Al particles with small size and a reduced size
36
37 dispersion that are aligned along the substrate grooves. 8 The images clearly show these Al
38
39 NPs, with circular in-plane aspect and with a mean size that increases with increasing Al
40
41 coverage. The areal density ρ of the NPs in the arrays was assessed by carefully applying a
42
43 digital recognition algorithm of the ”watershed” type to the images. The algorithm identifies
44
45 each isolated ”bump” above a critical pixel size in the images and assigns it to an individual
46
47 NP. The few coalesced structures, or features that could not be resolved as multiple closely-
48
49 spaced NPs were treated as a single NP. The reliability of the algorithm and its robustness
50
51 were carefully checked by manual inspection of the results. The so-obtained ρ reads 2000±60,
52
53 1750 ± 40, 1300 ± 30, 950 ± 30 NP/µm2 for z0 to z3. Knowing the amount of deposited Al
54
55 for each zone, it is then straightforward to calculate the average number of Al atoms per
56
57 particle NAl for each zone, which reads: NAl =2.5 × 104 , 6 × 104 , 1.2 × 105 and 2.2 × 105 atoms
58
59
60
5
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 6 of 22

1
2
3
4
from z0 to z3, respectively (Table 2). It can be clearly seen that the degree of alignment of
5
6 the NPs with the substrate grooves tend to decrease with increasing Al coverage, as possible
7
8 cross-groove NP coalescence sets in, decreasing the overall order of the NP arrays.
9
10
8 nm
a b 26 nm

11
12
13
14
15 e [1
[110]
0]
00 1
[0
16 ]
[001]
0 nm
0 nm
17
26 nm 26 nm
18
19 Λ

20
21
22
23
24 0 nm c d 0 nm
25
26 Figure 1: (a-d) AFM images of the Al/Al-oxide NPs systems obtained following the deposition of
27 different equivalent thickness of Al: (a) z0: 0.8 nm, (b) z1: 1.7 nm, (b) z2: 2.5 nm, (d)
28
29 z3: 3.4 nm. The area of the images is 1000 x 1000 nm2 . (e) Schematic representation
30 of the nanostructured LiF substrate and the supported Al NPs.
31
32
33 .
34
35
36
37 2.3 X-ray Photoeletron Spectroscopy
38
39
40 In Fig. 2 we report XPS spectra of the Al-NP system. Fig. 2b (c) shows the O1s (Al2p)
41
42 energy region of the XPS spectra measured on the Al NP arrays (z3 to z0 from top to
43
44 bottom). The spectra have been recorded in the geometry depicted in Fig. 3. The cor-
45
46 responding spectra recorded on a flat and continuous 20-nm-thick Al film, measured for
47
48 reference purposes, are reported in Fig. 2a. At a first glance, all the Al2p spectra show two
49
50 main peaks, while the O1s spectra are composed by a single broad feature. In the case of Al,
51
52 the two peaks respectively correspond to atoms in metallic (lower BE) or oxidized (higher
53
54 BE) state. 29,30,32,34,35,39,40 The intensity of the metallic-Al peak monotonously decreases as
55
56 the mean NP size decreases, whereas the Al-oxide peak and the O1s peak almost seem un-
57
58 affected. We believe this effect can be rationalized based on the fact that the volume of
59
60
6
ACS Paragon Plus Environment
Page 7 of 22 The Journal of Physical Chemistry

1
2
3
4
5 a
6
7 Metal
8 O 1s Al 2p
interf.
9 III I
II core
10 II
I III
11
12 Oxide
13
14
15
b c
16 z3

17
18
19 z2
20
21
22
23 z1

24
25
26 z0
27
28
29
30
31
32
33 Figure 2: XPS experimental spectra of O1s and Al2p core-levels. Panel a: experimental spectra
34 of the reference flat-Al film. panels b (c): experimental spectra of O1s (Al2p) for z3,
35 z2, z1, z0 (top to bottom). Experimental data (markers), best fit (solid black line) and
36 deconvolved subcomponents (colored areas) are reported. The roman numbers (I to III)
37 are used to label the various subcomponents. OI (light gray), OII (gray), OIII (black),
38
39
AlIox , AlII III
ox , Alox (shades of gray), Alinterf (blue), Alcore (light blue).
40
41
42
43
44
45 a [1 ]
b
46 00 10
X-Rays ] [0
47 e
48 z
45°
49
50
D
Zmet

51
52 45°
53 Xmet x
54
55 Figure 3: (a) Schematic representation of the XPS measurement configuration (a) and scheme of a
56 nanoparticle, as considered for the XPS intensity calculations.
57
58
59
60
7
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 8 of 22

1
2
3
4
the oxide-shell, responsible for the XPS oxide contributions, scales more slowly for changing
5
6 Al effective thickness with respect to the metallic counterpart. However, the analysis per-
7
8 formed in the following will rely on relative intensities and XPS-peak BE, and is unaffected
9
10 by absolute-intensity issues.
11
12 More detailed information can be however obtained deconvolving these peaks into their
13
14 subcomponents. The Al2p spectra were thus fitted as a superposition of Doniach-Sunjic
15
16 doublets with branching ratio of 0.5 and a spin-orbit splitting of 0.42 eV. 30,32 The O1s
17
18 peaks were instead fitted as a superposition of Lorentian-Voigt components. A Shirley-type
19
20 background was employed in all cases. In the NP case, the width of the various subcompo-
21
22 nents of the Al-oxide peaks were constrained to be equal among themselves, thus effectively
23
24 limiting the number of independent degrees of freedom in the fit, and the same was done
25
26 for the metallic-Al subcomponents and for the O1s subcomponents. This because the peak
27
28 width/shape is mainly determined by the core-hole lifetime, that is expected to be equal
29
30 within each specific atomic environment (metallic or oxide). In the reference flat-Al case,
31
32 the same constraints were applied, although these peaks are narrower with respect to the
33
34 NP case. Indeed, the core-hole screening and lifetime in insulated NPs or in a continuous
35
36 metal film are different, leading to different peak width between the two cases (so-called final
37
38 state effect). 45 While measuring NP spectra, an external charge neutralizer was employed to
39
40
prevent the buildup of sample charge arising due to the insulating nature of the substrate.
41
42
The deconvolution of the spectra was initially performed for the flat-Al case, where the
43
44
45
smaller peak width allowed to better disentangle the various subcomponents. In doing so, we
46
47 adapted the approach of well-established photoemission studies. 30,32 In the Al2p core-level,
48
49 we identified three oxide subcomponents, AlIox (73.4 eV, light-gray area), AlII
ox (74.5eV , gray)
50
51 and AlIII
ox (75.2 eV, dark gray), that can be assigned to different ionization states of Al bound
52
53 to O. 30,32,35 The metal peak was instead fitted as the superimposition of two doublets spaced
54
55 in energy by 0.3 eV, christened Alcore (72.7eV, light blue) and Alinterf (72.4 eV, deep blue). 32
56
57 Please notice that the energy alignment of the BEs was performed setting Alinterf at 72.4
58
59
60
8
ACS Paragon Plus Environment
Page 9 of 22 The Journal of Physical Chemistry

1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16 a
17 III
Alox
18
19
20
21
22 II Alcore Alinterf
23
Alox
I
24 Alox
25
26
27
28
29
30
b II
z0
31 O
32
33 I
34 O
35
36
III
37 O
38
39
40
41
42
43 Figure 4: Binding energy and peak area of each component of the spectra shown in panel c and
44
45 b for Al2p (top) and O1s (bottom). The stick height is proportional to the peak area,
46 while its position on the energy axis corresponds to its BE.
47
48
49
50
51
52
53
54
55
56
57
58
59
60
9
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 10 of 22

1
2
3
4
eV, also for the NP systems. 32
5
6 The main contribution to each Al2p spectrum is represented by AlIII
ox , corresponding
7
8 to the amorphous bulk aluminium oxide resulting from the low temperature oxidation of
9
10 the metal. 28,34,39 The two minor contributions to the oxide peak, AlII I
ox and Alox , correspond
11
12 instead to sub-oxide components. The dominance of the AlIII
ox peak is a typical fingerprint of
13
14 the advanced state of oxide formation that is witnessed after atmospheric exposure. 30 The
15
16 two metallic subcomponents have a more intriguing origin: indeed, while the Alcore peak is
17
18 assigned to bulk metallic Al, the Alinterf state arises upon oxidation, and has been assigned
19
20 to undercoordinated metallic Al atoms within the metallic state. It will be discussed in more
21
22 depth in the following.
23
24 The O1s peak has been fitted as the superposition of three subcomponents, OI (530.6 eV,
25
26 light gray), OII (531.8 eV, gray) and OIII (533.1 eV, dark gray), corresponding to aluminium
27
28 oxide, aluminium hydroxide and water, respectively. 31,37,38
29
30 Table 1: Energy of the subcomponents of the Al2p (top) and O1s (bottom) spectra for the reference
31
flat film and the four nanoparticle arrays. The BEs are expressed in electronvolts.
32
33
34 AlIII
ox AlII
ox AlIox Alcore Alinterf
35
36 Flat 75.2 74.5 73.4 72.7 72.4
37
38 z0 75.4 74.6 73.8 – 72.4
39 z1 75.4 74.8 73.65 – 72.4
40
41 z2 75.5 74.75 73.75 72.9 72.4
42 z3 75.35 74.6 73.7 72.7 72.4
43
44 OIII OII OI
45
46 Flat 533.1 531.8 530.6
47
48 z0 532.8 531.9 530.6
49 z1 532.8 531.7 530.4
50
51 z2 533.0 531.9 530.4
52 z3 533.0 531.9 530.7
53
54
55
56 The NP spectra were fitted in strict analogy with the reference ”flat-Al” case. However,
57
58 since the spectral subcomponents become broader due to final-state effects, especially for
59
60
10
ACS Paragon Plus Environment
Page 11 of 22 The Journal of Physical Chemistry

1
2
3
4
the metallic-Al case, we effectively constrained the energies of the various NP-spectrum
5
6 subcomponents to lie within ±0.4 eV (Al oxide), ±0.2 eV (metallic Al) ±0.3 eV (O1s) with
7
8 respect to the reference values. This was done to prevent the achievement of good fits with
9
10 unphysical energy shifts of the fit subcomponents. The best-fit BEs for each subcomponent
11
12 are reported in Table 1, and graphically represented in Fig. 4. In Fig. 4 it can be clearly seen
13
14 that the energy spread in the subcomponent BEs, which naturally arises from the presence
15
16 of noise in the spectra, and gets larger upon the broadening of the spectral features, is much
17
18 smaller than the energy separation between the different subcomponents, meaning that the
19
20 subcomponent assignment performed for the flat-Al reference still holds for the Al NPs.
21
22 Incidentally, this implies that fitting the metallic peak with a single component would yield
23
24 an unphysical result in comparison with the reference case, supporting the evidence that
25
26 both ”interface” and ”bulk” subcomponents can be observed in NPs.
27
28 Looking at the NP spectra, we notice some clear trends of subcomponent intensity as a
29
30 function of the NP size for both the O1s and the Al2p peaks. For O1s, we witness a gradual
31
32 increase in the relative weight of Al-oxide (OI ) with respect to hydroxide (OII ) for decreasing
33
34 NP size, and the disappearance of the water component. In the smallest-NP case (z0), a
35
36 broad feature (systematically observed) appears at BE lower than the O1s main peak, for
37
38 which however, no clear interpretation can be found in the current literature.
39
40
Speaking of Al2p, whereas its oxide subcomponents show no well-defined trends in in-
41
42
tensity or BE upon NP size, an interesting trend emerges for the intensity of the metallic
43
44
45
counterpart. Indeed, it appears that the Al2p metallic peak can be fitted by the two estab-
46
47 lished Alcore and Alinterf components only in the case of larger NP size (z2 and z3), whereas
48
49 for the smallest size (z0 and z1) the metal peak is compatible with the exclusive presence of
50
51 the ”interface” subcomponent Alinterf . In retrospect, this justifies the choice of referring the
52
53 BE to this peak, since the other obvious choice (Alcore ) is not observable for all the spectra.
54
55
56
57
58
59
60
11
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 12 of 22

1
2
3
4
2.4 Discussion
5
6 The NP size, crucial for discussing the data, cannot be extracted from AFM data alone.
7
8 However, combining XPS and AFM data, the mean NP size can be successfully extracted
9
10 applying the model thoroughly described in Ref. 8. To this end, the NPs were modelled
11
12 as semiellipsoids laid onto the (010) facet of the LiF ripples (Fig. 3a), having circular
13
14
cross-section parallel to the (010) facet and composed by a metallic core surrounded on the
15
16
17
atmosphere side by an oxide shell of thickness D = 2 nm, as appropriate for brief exposure
18
19
to atmosphere. 3 Fixing the total number of Al atoms per NP at the experimental NAl , the
20
21 ratio between the oxide and metallic XPS contributions depends on the NP-core in-plane
22
23 radius (xmet ) and its out-of-plane aspect ratio (zmet /xmet ) only (Fig. 3b). 8 Thus, comparing
24
25 the experimental metal/oxide ratio in Al2p with the ones calculated by the model of Ref.
26
27 8, the mean values of zmet and xmet can be deduced. The mean NP dimensions so obtained
28
29 are reported in table 2. The total NP diameter (core & shell) in the (010)-facet plane reads
30
31 (10 ± 4) nm, (17 ± 5) nm, (22± 8) nm, (29 ± 10) nm for z0 through z3 respectively, with
32
33 particles that tend to get flatter for increasing size. The uncertainty in NP size is mostly to
34
35 be ascribed to the size dispersion of the NPs. 8 The high out-of-plane aspect ratio calculated
36
37 for z0 clearly hints at the breakdown of the model’s assumptions for such a small NP size,
38
39 that will be addressed in detail further on.
40
41 Table 2: Geometrical parameters of the NP arrays: Nanoparticle density ρ and mean in-plane and
42 out-of-plane semiaxis of the ellipsoidal core).
43
44
45 Zone ρ xmet zmet
46 N P s/µm2 nm
47
48
0 2000 ± 60 3.2 ± 1.6 7.7 ± 3.9
49 1 1750 ± 40 6.3 ± 2.6 5.7 ± 2.2
50
51
2 1300 ± 30 9.1 ± 3.9 6.4 ± 2.8
52 3 950 ± 30 12.3 ± 5.1 7.4 ± 3.1
53
54
55
56 Having assessed the NP size, we can now discuss the XPS data in more depth. For the O1s
57
58 peak, strong similarities are present between the flat reference and the NP spectra, suggesting
59
60
12
ACS Paragon Plus Environment
Page 13 of 22 The Journal of Physical Chemistry

1
2
3
4
very similar oxide composition in the two cases. The two main O1s subcomponents (OI
5
6 and OII ) correspond to the O2− anions in non-crystalline alumina and to surface OH−
7
8 groups, respectively, 31,37,38 though the OH− peak may bear a contribution from adventitious
9
10 species. 38,46 The hydroxilation of the NP occurs upon interaction with atmospheric humidity,
11
12 which is also responsible for the relatively weak undissociated-water peak, OIII . 37
13
14 The water-related oxygen subcomponents are simply due to the interaction of the system
15
16 with water vapour during the short exposure to atmosphere during the transfer from the
17
18 preparation chamber to the XPS setup. The small intensity of the OIII peak demonstrates
19
20 that only a very small fraction of adsorbed water is present, while most molecules quickly
21
22 dissociate leading to surface hydroxilation. The OII peak can bear a contribution from
23
24 the Al-related OH− groups and from adventitious molecules adsorbed on the sample during
25
26 the transit in atmosphere. The OH− contribution can be thus calculated subtracting from
27
28 the OII peak the component due to adventitious molecules that can be extracted from the
29
30 deconvolution of the high-resolution C1s core-level spectra (not shown). Such a contribution
31
32 reads between 28−33% in all zones. After this subtraction, the OH− component exceeds
33
34 that of O2− for all the four zones, thus demonstrating a significant hydroxilation of the
35
36 nanoparticle surface. The oxide stoichiometry, calculated through the intensity ratio between
37
38 the oxide-related O1s and Al2p components (XPS intensity yield 4:1), reads ≈ 2 for all the
39
40
four zones, a value that corresponds to an oxyhydroxide stoichiometry, indicating that the
41
42
exposure to atmosphere might induce the formation of AlO(OH), as previously observed in
43
44
45
the case of air-formed Al films. 37
46
47 More intriguing is the analysis of the Al2p line. The oxide-related components, as pre-
48
49 viously mentioned, are dominated by AlIII
ox subcomponent, as expected for the advanced
50
51 degree of oxidation of the systems. We notice that the intensity of the oxide subcomponents
52
53 remains almost unchanged as a function of decreasing NP size. This is mainly a consequence
54
55 of the limited probing depth of XPS and an indication that the oxide thickness does not sig-
56
57 nificantly vary as a function of NP size. Coming to the metallic subcomponents, the crucial
58
59
60
13
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 14 of 22

1
2
3
4
point is the Alinterf contribution. Previous XPS studies on oxidation of single-crystal sys-
5
6 tems had already detected it and vaguely ascribed it to metallic Al atoms in proximity of the
7
8 metal/oxide interface. 32,35 Its origin remains however unclear; in particular, the energy shift
9
10 towards lower BE cannot be rationalized in terms of oxygen-induced electron depletion but
11
12 instead suggest the formation of open and undercoordinated structures within the metallic
13
14 Al. 47 The association of the Alinterf component with phenomena occurring at the metal/oxide
15
16 interface is consistent with its intensity trend vs NP size, that sees the Alinterf contribution
17
18 become dominant as the NP size is reduced. In very fine NPs (diameter<20 nm), our XPS
19
20 analysis assigns the totality of the metallic atoms to such an interface electronic state.
21
22 In order to understand its origin and its implications for ultrafine Al NPs, the microscopic
23
24 dynamics of Al oxidation has to be recalled. According to the Cabrera-Mott theory, indeed,
25
26 Al oxidation is characterized by a faster outward migration of the Al cations with respect
27
28 to the oxygen anions through the forming oxide layer. 27 Such a different diffusion velocity
29
30 induces the formation of vacancies in the metal at the metal-oxide interface that coalesce to
31
32 form vacancy clusters leading to the formation of open, undercoordinated metallic structures
33
34 near the interface. We remark that such Al-interface state, originally observed upon con-
35
36 trolled oxidation in ultra-high vacuum of Al single crystal, appears as a remarkably general
37
38 effect, observable also under the less-ideal conditions in which we performed the oxidation,
39
40
and therefore of wide impact in all the atmospheric applications of Al NPs.
41
42
In NPs, these microscopic effects migth have a potentially disruptive impact on the metal
43
44
45
core, as indeed shown in a recent transmission-electron microscopy study where metallic Al
46
47 NPs with diameter below a critical value (d <8 nm) were so heavily affected by oxidation
48
49 to completely lose their metallic core (Kirkendall effect), giving rise to hollow-core NP mor-
50
51 phologies. 28,48 Based on this, we therefore suggest that the Alinterf is the spectral fingerprint
52
53 of vacancy clusters at the metal/oxide interface. This phenomenon is clearly associated with
54
55 an alteration of the electronic properties of the NP core with respect to its idealized bulk
56
57 counterpart. Its increasing relevance for smaller NP surely implies a substantial difference of
58
59
60
14
ACS Paragon Plus Environment
Page 15 of 22 The Journal of Physical Chemistry

1
2
3
4
the overall electronic (hence plasmonic) response of the these systems with respect to expec-
5
6 tations based on oversimplified NP-oxidation models. This incidentally promotes XPS as an
7
8 effective method for assessing nanometric characteristics of NPs having a complex chemical
9
10 structure.
11
12 In this respect, we have to mention that recent measurements of DUV plasmonic response
13
14 of Al core/shell NPs revealed that NPs as small as those of z1 still support a clearly-detectable
15
16 localized-surface plasmon resonance, 8 whereas for smaller NPs (z0) no such response was
17
18 observed. This sets the threshold for plasmonic functionality in between the pertinent mean
19
20 NP size for z0 and z1 systems, where z0 NPs, while still retaining a fraction of metallic
21
22 atoms, are likely characterized by a disrupted core morphology unable to sustain detectable
23
24 free-electron plasma oscillations.
25
26
27
28
29 3 Conclusions
30
31
32 We reported a combined AFM and XPS study of the electronic properties of ultrafine Al/Al-
33
34 oxide core/shell NPs supported on nanopatterned insulating substrates. The photoemission
35
36 spectra of Al NPs are deconvolved as the superposition of several chemically-shifted sub-
37
38 components, indicative of a complex chemistry of the oxidized shell and the metallic core.
39
40 We identified a characteristic low-binding-energy valence state of metallic Al, that could
41
42 be ascribed to undercoordinated metallic atoms within the core. For decreasing NP size,
43
44 the undercoordinated metallic structures coalesce in vacancy clusters within the NP inner
45
46 metallic core, possibly leading to the disruption of the homogeneous Al metallic core for NP
47
48 diameters below 10 nm, with heavy implications for the electronic and plasmonic response
49
50 of ultrafine Al particles.
51
52
53
54
55
56
57
58
59
60
15
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 16 of 22

1
2
3
4
4 Acknowledgments
5
6
7 Financial support from the Fondazione Carige, from Sincrotrone Elettra SCpA (PIK Ex.Pro.Rel.)
8
9 and from the Ministero dell’Istruzione, dell’Università e della Ricerca (Project no. PRIN
10
11 2008AKZSXY 002) is acknowledged. The image analysis was performed using the open-
12
13 source software Gwyddion. 49
14
15
16
17 References
18
19
20
(1) Chan, G. H.; Zhao, J.; Schatz, G. C.; van Duyne, R. P. Localized Surface Plasmon
21
22
23
Resonance Spectroscopy of Triangular Aluminum Nanoparticles. J. Phys. Chem. C
24
25 2008, 13958.
26
27
(2) Ekinci, Y.; Solak, H. H.; Löffler, J. F. Plasmon Resonances of Aluminum Nanoparticles
28
29
30
and Nanorods. J. Appl. Phys. 2008, 104, 083107.
31
32
(3) Langhammer, C.; Schwind, M.; Kasemo, B.; Zorić, I. Localized Surface Plasmon Reso-
33
34
nances in Aluminum Nanodisks. Nano Lett. 2008, 8, 1461–1471.
35
36
37 (4) Stratakis, E.; Barberoglou, M.; Fotakis, C.; Viau, G.; Garcia, C.; Shafeev, G. A. Gen-
38
39
eration of Al Nanoparticles via Ablation of Bulk Al in Liquids with Short Laser Pulses.
40
41
Opt. Express 2009, 17, 12650–12659.
42
43
44 (5) Blaber, M. G.; Arnold, M. D.; Ford, M. J. Search for the Ideal Plasmonic Nanoshell:
45
46
The Effects of Surface Scattering and Alternatives to Gold and Silver. J. Phys. Chem.
47
48
C 2009, 113, 3041–3045.
49
50
51 (6) Hu, J.; Chen, L.; Lian, Z.; Cao, M.; Li, H.; Sun, W.; Tong, N.; Zeng, H. Deep-Ultraviolet
52
53
- Blue-Light Surface Plasmon Resonance of Alcore and Al2 O3shell in Spherical and Cylin-
54
55
56
drical Nanostructures. J. Phys. Chem. B 2012, 116, 15584–15590.
57
58
59
60
16
ACS Paragon Plus Environment
Page 17 of 22 The Journal of Physical Chemistry

1
2
3
4
(7) Taguchi, A.; Saito, Y.; Watanabe, K.; Yijian, S.; Kawata, S. Tailoring Plasmon Reso-
5
6 nances in the Deep-Ultraviolet by Size-Tunable Fabrication of Aluminum Nanostruc-
7
8 tures. Appl. Phys. Lett. 2012, 101, 081110.
9
10
11
(8) Maidecchi, G.; Gonella, G.; Proietti Zaccaria, R.; Moroni, R.; Anghinolfi, L.; Bisio, F.
12
13 Deep Ultraviolet Plasmon Resonance in Aluminum Nanoparticle Arrays. ACS Nano
14
15 2013, 7, 5834–5841.
16
17
18
(9) Knight, M. W.; King, N. S.; Liu, L.; Everitt, H. O.; Nordlander, P.; Halas, N. J.
19
20 Aluminum for Plasmonics. ACS Nano 2014, 8, 834–840.
21
22
23
(10) Jiao, X.; Blair, S. Polarization Multiplexed Optical Bullseye Antennas. Plasmonics
24
25
2011, 7, 39–46.
26
27
(11) Knight, M. W.; Liu, L.; Wang, Y.; Brown, L.; Mukherjee, S.; King, N. S.; Everitt, H. O.;
28
29
30
Nordlander, P.; Halas, N. J. Aluminum Plasmonic Nanoantennas. Nano Lett. 2012, 12,
31
32
6000–6004.
33
34
(12) Akimov, Y. A.; Koh, W. S. Design of Plasmonic Nanoparticles for Efficient Subwave-
35
36
37
length Light Trapping in Thin-Film Solar Cells. Plasmonics 2010, 6, 155–161.
38
39 (13) Kochergin, V.; Neely, L.; Jao, C.-Y.; Robinson, H. D. Aluminum Plasmonic Nanos-
40
41
tructures for Improved Absorption in Organic Photovoltaic Devices. Appl. Phys. Lett.
42
43
44
2011, 98, 133305.
45
46 (14) Ferry, V. E.; Polman, A.; Atwater, H. A. Modeling Light Trapping in Nanostructured
47
48
Solar Cells. ACS Nano 2011, 5, 10055–10064.
49
50
51 (15) Zhang, Y.; Ouyang, Z.; Stokes, N.; Jia, B.; Shi, Z.; Gu, M. Low Cost and High Perfor-
52
53 mance Al Nanoparticles for Broadband Light Trapping in Si Wafer Solar Cells. Appl.
54
55
Phys. Lett. 2012, 100, 151101.
56
57
58
59
60
17
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 18 of 22

1
2
3
4
(16) Hylton, N. P.; Li, X. F.; Giannini, V.; Lee, K.-H.; Ekins-Daukes, N. J.; Loo, J.; Ver-
5
6 cruysse, D.; Van Dorpe, P.; Sodabanlu, H.; Sugiyama, M. et al. Loss Mitigation in
7
8 Plasmonic Solar Cells: Aluminium Nanoparticles for Broadband Photocurrent En-
9
10 hancements in GaAs Photodiodes. Sci. Rep. 2013, 3, 2874.
11
12
13 (17) Zorić, I.; Kasemo, B.; Langhammer, C.; Zäch, M. Gold, Platinum and Aluminum
14
15 Nanodisk Plasmons : Material Damping Mechanisms. ACS Nano 2011, 5, 2535–2546.
16
17
18
(18) Lu, G.; Li, W.; Zhang, T.; Yue, S.; Liu, J.; Hou, L.; Li, Z.; Gong, Q. Plasmonic-
19
20 Enhanced Molecular Fluorescence within Isolated Bowtie Nano-Apertures. ACS Nano
21
22 2012, 6, 1438–1448.
23
24
25
(19) Ray, K.; Chowdhury, M. H.; Lakowicz, J. R. Aluminum Nanostructured Films as Sub-
26
27 strates for Enhanced Fluorescence in the Ultraviolet-Blue Spectral Region. Anal. Chem.
28
29 2007, 79, 6480–6487.
30
31
32
(20) Taguchi, A.; Hayazawa, N.; Furusawa, K.; Ishitobi, H.; Kawata, S. Deep-UV Tip-
33
34 Enhanced Raman Scattering. J. Raman Spectrosc. 2009, 40, 1324–1330.
35
36
37
(21) Jha, S. K.; Ahmed, Z.; Agio, M.; Ekinci, Y.; Löffler, J. F. Deep-UV Surface-Enhanced
38
39
Resonance Raman Scattering of Adenine on Aluminum Nanoparticle Arrays. J. Am.
40
41 Chem. Soc. 2012, 134, 1966–1969.
42
43
44
(22) Castro-Lopez, M.; Brinks, D.; Sapienza, R.; van Hulst, N. F. Aluminum for Nonlinear
45
46
Plasmonics: Resonance-Driven Polarized Luminescence of Al, Ag, and Au Nanoanten-
47
48 nas. Nano Lett. 2011, 11, 4674–4678.
49
50
51
(23) Thyagarajan, K.; Rivier, S.; Lovera, A.; Martin, O. J. F. Enhanced Second-Harmonic
52
53
Generation from Double Resonant Plasmonic Antennae. Opt. Express 2012, 20, 12860–
54
55 12865.
56
57
58
59
60
18
ACS Paragon Plus Environment
Page 19 of 22 The Journal of Physical Chemistry

1
2
3
4
(24) Bisio, F.; Proietti Zaccaria, R.; Moroni, R.; Maidecchi, G.; Alabastri, A.; Gonella, G.;
5
6 Giglia, A.; Andolfi, L.; Nannarone, S.; Mattera, L. et al. Pushing the High-Energy
7
8 Limit of Plasmonics. ACS Nano 2014, 8, 9239–9247.
9
10
11
(25) Martin, J.; Kociak, M.; Mahfoud, Z.; Proust, J.; Gérard, D.; Plain, J. High-Resolution
12
13 Imaging and Spectroscopy of Multipolar Plasmonic Resonances in Aluminum Nanoan-
14
15 tennas. Nano Lett. 2014, 14, 5517–5523.
16
17
18
(26) Bisio, F.; Gonella, G.; Maidecchi, G.; Buzio, R.; Gerbi, A.; Moroni, R.; Giglia, A.;
19
20 Canepa, M. Broadband Plasmonic Response of Self-Organized Aluminium Nanowire
21
22 Arrays. J. Phys. D: Appl. Phys. 2015, 48, 184003.
23
24
25
(27) Cabrera, N.; Mott, N. F. Theory of the Oxidation of Metals. Rep. Prog. Phys. 1948,
26
27 12, 163–184.
28
29
30
(28) Nakamura, R.; Tokozakura, D.; Nakajima, H.; Lee, J.-G.; Mori, H. Hollow Oxide For-
31
32
mation by Oxidation of Al and Cu Nanoparticles. J. Appl. Phys. 2007, 101, 074303.
33
34
(29) Hagstrom, S. B. M.; Bachrach, R. Z.; Bauer, R. S. Oxidation of Aluminum Surfaces
35
36
37
Studied by Synchrotron Radiation Photoelectron Spectroscopy. Phys. Scr. 1977, 16,
38
39
414–419.
40
41
(30) McConville, C. F.; Seymour, D. L.; Woodruff, D. P.; Bao, S. Synchrotron Radiation
42
43
44
Photoemission Investigation of the Initial Stages of Oxidation of Al(111). Surf. Sci.
45
46
1987, 188, 1–14.
47
48
(31) Strohmeier, B. R. The Effects of O2 Plasma Treatments on the Surface Composition
49
50
51
and Wettability of Cold-Rolled Aluminum Foil. Am. Vac. Soc. 1989, 7, 3238–3245.
52
53 (32) Berg, C.; Raaen, S.; Borg, A.; Andersen, J. S.; Lundgren, E.; Nyholm, R. Observation of
54
55
a Low-Binding-Energy Peak in the 2p Core-Level Photoemission from Oxidized Al(111).
56
57
58
Phys. Rev. B 1993, 47, 13063–13066.
59
60
19
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 20 of 22

1
2
3
4
(33) Jeurgens, L. P. H.; Sloof, W. G.; Borsboom, C. G.; Tichelaar, F. D. Determination
5
6 of Total Primary Zero Loss Intensities in Measured Electron Emission Spectra of Bare
7
8 and Oxidised Metals. Application to Aluminium Oxide Films on Aluminium Substrates.
9
10 Appl. Surf. Sci. 2000, 161, 139–148.
11
12
13 (34) Jeurgens, L. P. H.; Sloof, W. G. Composition and Chemical State of the Ions of
14
15 Aluminium-Oxide Films Formed by Thermal Oxidation of Aluminium. Surf. Sci. 2002,
16
17 506, 313–332.
18
19
20 (35) Schouborg, J.; Raarup, M. K.; Balling, P. Adsorbate Reactivity and Thermal Mo-
21
22 bility from Simple Modeling of High-Resolution Core-Level Spectra: Application to
23
24 O/Al(111). J. Phys. Condens. Matter 2009, 21, 265003.
25
26
27 (36) Do, T.; McIntyre, N. S.; Harshman, R. A.; Lundy, M. E.; Splinter, S. J. Surf. Interf.
28
29 Anal. 618–628.
30
31
32
(37) Alexander, M. R.; Thompson, G. E.; Beamson, G. Characterization of the Ox-
33
34 ide/Hydroxide Surface of Aluminium Using X-ray Photoelectron Spectroscopy : a Pro-
35
36 cedure for Curve Fitting the O 1s Core Level. Surf. Interface Anal. 2000, 29, 468–477.
37
38
39
(38) van den Brand, J.; Sloof, W. G.; Terryn, H.; de Wit, J. H. W. Correlation between
40
41 Hydroxyl Fraction and O/Al Atomic Ratio as Determined from XPS Spectra of Alu-
42
43 minium Oxide Layers. Surf. Interface Anal. 2004, 36, 81–88.
44
45
46
(39) Frerichs, M.; Voigts, F.; Maus-Friedrichs, W. Fundamental Processes of Aluminium
47
48 Corrosion Studied under Ultra High Vacuum Conditions. Appl. Surf. Sci. 2006, 253,
49
50 950–958.
51
52
53
(40) Sánchez-Lopez, J. C.; Fernández, A. Photoelectron Spectroscopy versus Absorption
54
55 Spectroscopy for Quantitative Characterization of Nanometric Powders Coated with
56
57 an Overlayer. Surf. Interface Anal. 1998, 26, 1016–1026.
58
59
60
20
ACS Paragon Plus Environment
Page 21 of 22 The Journal of Physical Chemistry

1
2
3
4
(41) Anghinolfi, L.; Moroni, R.; Mattera, L.; Canepa, M.; Bisio, F. Flexible Tuning of
5
6 Shape and Arrangement of Au Nanoparticles in 2-Dimensional Self-Organized Arrays:
7
8 Morphology and Plasmonic Response. J. Phys. Chem. C 2011, 115, 14036–14043.
9
10
11
(42) Sugawara, A.; Mae, K. Surface Morphology of Epitaxial LiF(110) and CaF2 (110) Lay-
12
13 ers. J. Vac. Sci. Technol. B 2005, 23, 443–448.
14
15
16
(43) Kitahara, T.; Sugawara, A.; Sano, H.; Mizutani, G. Optical Second-Harmonic Spec-
17
18
troscopy of Au Nanowires. J. Appl. Phys. 2004, 95, 5002–5005.
19
20
(44) Knight, M. W.; King, N. S.; Liu, L.; Everitt, H. O.; Nordlander, P.; Halas, N. J.
21
22
23
Aluminum for Plasmonics. ACS Nano 2014, 8, 834–840.
24
25 (45) Mason, M. G. Electronic Structure of Supported Small Metal Clusters. Phys. Rev. B
26
27
1983, 27, 748–762.
28
29
30 (46) Briggs, D.; Beamson, XPS Studies of the Oxygen 1s and 2s Levels in a Wide Range of
31
32 Functional Polymers. Anal. Chem. 1993, 65, 1517–1523.
33
34
35 (47) Borg, M.; Birgersson, M.; Smedh, M.; Mikkelsen, A.; Adams, D.; Nyholm, R.; Alm-
36
37 bladh, C.-O.; Andersen, J. Experimental and Theoretical Surface Core-Level Shifts of
38
39 Aluminum (100) and (111). Phys. Rev. B 2004, 69, 235418.
40
41
42 (48) Anderson, B. D.; Tracy, J. B. Nanoparticle Conversion Chemistry: Kirkendall Effect,
43
44 Galvanic Exchange, and Anion Exchange. Nanoscale 2014, 6, 12195–12216.
45
46
47 (49) Necas, D.; Klapetek, P. Gwyddion: an Open-Source Software for SPM Data Analysis.
48
49 Cent. Eur. J. Phys. 2012, 10, 181–188.
50
51
52
53
54
55
56
57
58
59
60
21
ACS Paragon Plus Environment
The Journal of Physical Chemistry Page 22 of 22

1
2
3 >10 nm - <10 nm
4 e Al core
5 X-Rays Al interface
6 Al oxide
7
8
9
10 XPS
11
12
13
14
15 Figure 5: Table of Content graphics.
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
22
ACS Paragon Plus Environment

You might also like